rotating trapped bose-einstein condensatespjh/teaching/phz7429/fetter.pdfii. physics of...

45
Rotating trapped Bose-Einstein condensates Alexander L. Fetter * Geballe Laboratory for Advanced Materials and Departments of Physics and Applied Physics, Stanford University, Stanford, California 94305-4045, USA Published 18 May 2009 After reviewing the ideal Bose-Einstein gas in a box and in a harmonic trap, the effect of interactions on the formation of a Bose-Einstein condensate are discussed, along with the dynamics of small-amplitude perturbations the Bogoliubov equations. When the condensate rotates with angular velocity , one or several vortices nucleate, leading to many observable consequences. With more rapid rotation, the vortices form a dense triangular array, and the collective behavior of these vortices has additional experimental implications. For near the radial trap frequency , the lowest-Landau-level approximation becomes applicable, providing a simple picture of such rapidly rotating condensates. Eventually, as , the rotating dilute gas is expected to undergo a quantum phase transition from a superfluid to various highly correlated nonsuperfluid states analogous to those familiar from the fractional quantum Hall effect for electrons in a strong perpendicular magnetic field. DOI: 10.1103/RevModPhys.81.647 PACS numbers: 03.75.Hh, 05.30.Jp, 67.10.Fj CONTENTS I. Introduction 648 II. Physics of Bose-Einstein Condensates in Dilute Trapped Gases 648 A. Ideal Bose gas 649 1. Ideal BEC in a box with periodic boundary conditions 649 2. Ideal BEC in a harmonic trap 650 B. Laser trapping and optical lattices 650 C. Inclusion of interparticle interactions 651 D. Bogoliubov equations 654 III. Physics of One Vortex a Few Vortices in a Trap 655 A. One vortex in unbounded condensate 655 B. One vortex in a trapped condensate 656 1. Gross-Pitaevskii energy in a trap 656 2. Dynamical motion of a trapped vortex 657 a. Lagrangian functional 657 b. Matched asymptotic expansion 658 c. Bogoliubov equations and stability of a vortex 659 C. Small vortex arrays in a trap 659 D. Experimental and theoretical studies in a trap 660 1. Vortex behavior in two-component BEC mixtures JILA 660 2. Vortex behavior in one-component BECs Ecole Normale Supérieure ENS and subsequently others 662 3. Normal modes of an axisymmetric condensate with one central vortex 664 4. Bent vortex configurations 665 5. Kelvin axial waves on a vortex 665 6. Other ways to create vortices 666 a. Transfer of orbital angular momentum from a laser 666 b. Merging of multiple BECs 666 c. Imposition of topological phase 666 E. Berezinskii-Kosterlitz-Thouless transition in dilute BECs 667 IV. Vortex Arrays in Mean-Field Thomas-Fermi TF Regime 668 A. Physics of BEC in axisymmetric harmonic traps for rapid rotation 668 B. Experimental and theoretical studies of vortex lattices in the TF regime 670 1. Collective modes of rotating condensates 670 2. Uniformity of the vortex array 671 3. Vortex core size for large rotation speeds 672 4. Tkachenko oscillations of the vortex lattice 673 5. Rotating two-component BECs 674 6. Rotating gas of paired fermions in a tightly bound BEC limit 674 C. Effect of additional quartic confinement 676 V. Vortex Arrays in Mean-Field LLL Regime 678 A. Physics of LLL one-body states 678 B. Direct treatment of LLL states 679 C. LLL condensate wave functions 680 1. Unrestricted minimization 680 2. Inclusion of the vortices 681 a. Renormalization of g 2D 681 b. Nonuniformity of the vortex array 682 3. Rapidly rotating anisotropic traps 682 VI. Quantum Phase Transition to Highly Correlated States 683 VII. Outlook 684 A. Successes 684 B. Challenges 685 1. Rapid rotation of both Bose and Fermi gases in the quantum Hall regime at small filling fraction 685 2. Rotation of ultracold dilute gases with long-range dipole-dipole interactions 685 * [email protected] REVIEWS OF MODERN PHYSICS, VOLUME 81, APRIL–JUNE 2009 0034-6861/2009/812/64745 ©2009 The American Physical Society 647

Upload: others

Post on 05-Jun-2020

5 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Rotating trapped Bose-Einstein condensatespjh/teaching/phz7429/fetter.pdfII. Physics of Bose-Einstein Condensates in Dilute Trapped Gases 648 A. Ideal Bose gas 649 1. Ideal BEC in

Rotating trapped Bose-Einstein condensates

Alexander L. Fetter*

Geballe Laboratory for Advanced Materials and Departments of Physics and AppliedPhysics, Stanford University, Stanford, California 94305-4045, USA

�Published 18 May 2009�

After reviewing the ideal Bose-Einstein gas in a box and in a harmonic trap, the effect of interactionson the formation of a Bose-Einstein condensate are discussed, along with the dynamics ofsmall-amplitude perturbations �the Bogoliubov equations�. When the condensate rotates with angularvelocity �, one or several vortices nucleate, leading to many observable consequences. With morerapid rotation, the vortices form a dense triangular array, and the collective behavior of these vorticeshas additional experimental implications. For � near the radial trap frequency ��, thelowest-Landau-level approximation becomes applicable, providing a simple picture of such rapidlyrotating condensates. Eventually, as �→��, the rotating dilute gas is expected to undergo a quantumphase transition from a superfluid to various highly correlated �nonsuperfluid� states analogous tothose familiar from the fractional quantum Hall effect for electrons in a strong perpendicular magneticfield.

DOI: 10.1103/RevModPhys.81.647 PACS number�s�: 03.75.Hh, 05.30.Jp, 67.10.Fj

CONTENTS

I. Introduction 648II. Physics of Bose-Einstein Condensates in Dilute

Trapped Gases 648A. Ideal Bose gas 649

1. Ideal BEC in a box with periodic boundaryconditions 649

2. Ideal BEC in a harmonic trap 650B. Laser trapping and optical lattices 650C. Inclusion of interparticle interactions 651D. Bogoliubov equations 654

III. Physics of One Vortex �a Few Vortices� in a Trap 655A. One vortex in unbounded condensate 655B. One vortex in a trapped condensate 656

1. Gross-Pitaevskii energy in a trap 6562. Dynamical motion of a trapped vortex 657

a. Lagrangian functional 657b. Matched asymptotic expansion 658c. Bogoliubov equations and stability of a

vortex 659C. Small vortex arrays in a trap 659D. Experimental and theoretical studies in a trap 660

1. Vortex behavior in two-component BECmixtures �JILA� 660

2. Vortex behavior in one-component BECs�Ecole Normale Supérieure �ENS� andsubsequently others� 662

3. Normal modes of an axisymmetriccondensate with one central vortex 664

4. Bent vortex configurations 6655. Kelvin �axial� waves on a vortex 6656. Other ways to create vortices 666

a. Transfer of orbital angular momentum

from a laser 666

b. Merging of multiple BECs 666

c. Imposition of topological phase 666

E. Berezinskii-Kosterlitz-Thouless transition in dilute

BECs 667

IV. Vortex Arrays in Mean-Field Thomas-Fermi �TF�Regime 668

A. Physics of BEC in axisymmetric harmonic traps forrapid rotation 668

B. Experimental and theoretical studies of vortexlattices in the TF regime 670

1. Collective modes of rotating condensates 6702. Uniformity of the vortex array 6713. Vortex core size for large rotation speeds 6724. Tkachenko oscillations of the vortex lattice 6735. Rotating two-component BECs 6746. Rotating gas of paired fermions in a tightly

bound �BEC� limit 674C. Effect of additional quartic confinement 676

V. Vortex Arrays in Mean-Field LLL Regime 678A. Physics of LLL one-body states 678B. Direct treatment of LLL states 679C. LLL condensate wave functions 680

1. Unrestricted minimization 6802. Inclusion of the vortices 681

a. Renormalization of g2D 681b. Nonuniformity of the vortex array 682

3. Rapidly rotating anisotropic traps 682VI. Quantum Phase Transition to Highly Correlated

States 683VII. Outlook 684

A. Successes 684B. Challenges 685

1. Rapid rotation of both Bose and Fermigases in the quantum Hall regime at smallfilling fraction 685

2. Rotation of ultracold dilute gases withlong-range dipole-dipole interactions 685*[email protected]

REVIEWS OF MODERN PHYSICS, VOLUME 81, APRIL–JUNE 2009

0034-6861/2009/81�2�/647�45� ©2009 The American Physical Society647

Page 2: Rotating trapped Bose-Einstein condensatespjh/teaching/phz7429/fetter.pdfII. Physics of Bose-Einstein Condensates in Dilute Trapped Gases 648 A. Ideal Bose gas 649 1. Ideal BEC in

3. Rotating spinor condensates 686Acknowledgments 687References 687

I. INTRODUCTION

The remarkable creation of Bose-Einstein conden-sates �BECs� in cold dilute alkali-metal gases �Andersonet al., 1995; Bradley et al., 1995; Davis et al., 1995� hasproduced a wholly new exciting field that continues tothrive. Typically, these systems have a macroscopic con-densate wave function ��r , t� that characterizes thestatic and dynamic behavior of the BEC �for a generalbackground, see, for example, Dalfovo et al. �1999�; In-guscio et al. �1999�; Pethick and Smith �2002�; Pitaevskiiand Stringari �2003��. In the usual limit of a dilute gas, �obeys a nonlinear Schrödinger equation known as theGross-Pitaevskii equation �Gross, 1961; Pitaevskii,1961�.

Subsequent experiments formed rotating condensates,initially with a small number of vortices �Matthews et al.,1999a; Madison et al., 2000a�, and then with large vortexarrays �Abo-Shaeer et al., 2001; Engels et al., 2002�. Asin the case of rotating superfluid 4He �Donnelly, 1991�,the vortices have quantized circulation, arising from thesingle-valued nature of the condensate wave function �.Nevertheless, the new dilute-gas BECs differ consider-ably from superfluid 4He because the Gross-Pitaevskiiequation provides a remarkably detailed description ofthe physics, allowing a careful comparison of theory andexperiment.

This article focuses on the physics of quantized vorti-ces in ultracold dilute trapped quantum gases. I first re-view the dilute Bose-Einstein gas, both in a box wherethe density is uniform and in harmonic traps where thedensity is nonuniform �Sec. II�. The presence of �repul-sive� interactions has a dramatic effect, leading to �1� acollective mode with a linear long-wavelength dispersionrelation and �2� superfluidity of macroscopic flow. Sec-tion III treats the behavior of one vortex �or a smallnumber of vortices� in a rotating condensate. For morerapid rotations, the vortices form a regular array thathas many analogies with rotating 4He and with the flux-line lattice in type-II superconductors �Sec. IV�. As therotation rate � continues to increase and approaches theradial trapping frequency ��, the effective radial trap-ping potential weakens, and the condensate expands. Inthis limit, the system enters a new two-dimensional su-perfluid regime that has close analogies with the lowest-Landau-level description of an electron in a uniformmagnetic field �Sec. V�. Ultimately, in the limit � /��

�0.999, the rotating Bose gas is predicted to make aquantum phase transition to one of several highly corre-lated ground states that are not superfluid �Sec. VI�. Thisregime involves a sequence of many-body states that aresimilar to those developed for the quantum Hall effect�a two-dimensional electron gas in a strong magneticfield�.

Throughout this article, I follow the lead of Blochet al. �2007�, focusing on models and predictions withexperimental basis or confirmation. As a result, I treatonly briefly two areas that have been studied in someexperimental detail as nonrotating systems: �1� dipolecondensates that interact with long-range electric ormagnetic dipole forces �see, for example, Griesmaier etal. �2005�; Lahaye et al. �2007�, and references therein�and �2� spinor condensates �see, for example, Stenger etal. �1998�; Sadler et al. �2006�; Vengalattore et al. �2008�,and references therein�. In both cases, however, rotationand the corresponding vortex structures remain an un-explored experimental capability �see, however, Sec.VII.B for an introductory discussion�.

II. PHYSICS OF BOSE-EINSTEIN CONDENSATES INDILUTE TRAPPED GASES

The general topic of Bose-Einstein condensates in di-lute trapped gases has been the focus of several reviewsand books �Dalfovo et al., 1999; Inguscio et al., 1999;Pethick and Smith, 2002; Pitaevskii and Stringari, 2003�.The current article will emphasize those aspects mostrelevant to rotating trapped gases and the associatedquantized vortices �Fetter and Svidzinsky, 2001; Aftal-ion, 2006; Parker et al., 2007; Kasamatsu and Tsubota,2008�. In particular, I will emphasize the regime of regu-lar vortex arrays, both the “mean-field Thomas-Fermi”limit when the vortex cores remain small and the“lowest-Landau-level” limit when the cores are compa-rable to the intervortex spacing �Ho, 2001; Baym, 2005�.

To review the basic facts of Bose-Einstein condensa-tion, consider an ideal �noninteracting� uniform three-dimensional gas with N particles in a cubical box �vol-ume V=L3� with mean density n=N /V. If the gas has atemperature T, the mean momentum per particle is pT

��MkBT, ignoring numerical factors of order unity,where M is the atomic mass and kB is Boltzmann’s con-stant. The de Broglie relation then yields the thermalwavelength �T�h /pT�h /�MkBT. The other relevantlength is the interparticle spacing n−1/3. The short-wavelength limit �T�n−1/3 holds when �→0 or T islarge; it describes the classical regime when quantuminterference and diffraction are negligible �the analog ofray optics�. As T falls, however, this inequality eventu-ally fails when �T�n−1/3, or, equivalently, when the“phase-space density” defined as n�T

3 becomes of orderunity. This condition signals the onset of quantum de-generacy; whenever n�T

3 �1, quantum mechanics plays acentral role in the physics.

The preceding description relies on an atomic-physicsperspective, but a condensed-matter view yields a simi-lar picture. In an ideal gas, each atom is effectively con-fined to a volume n−1 by its neighbors, leading to a zero-point energy �zp��2n2/3 /M. At high T or low density�kBT�zp�, the system behaves classically, but whenkBT��zp �because of either reduced temperature or in-creased density�, quantum effects again become crucial.

648 Alexander L. Fetter: Rotating trapped Bose-Einstein condensates

Rev. Mod. Phys., Vol. 81, No. 2, April–June 2009

Page 3: Rotating trapped Bose-Einstein condensatespjh/teaching/phz7429/fetter.pdfII. Physics of Bose-Einstein Condensates in Dilute Trapped Gases 648 A. Ideal Bose gas 649 1. Ideal BEC in

These two characterizations of the onset of quantum de-generacy are essentially equivalent.

For ideal bosons, the critical temperature Tc for theonset of quantum degeneracy leads to Bose-Einsteincondensation. For ideal Fermi gases, in contrast, thecomparable temperature is known as the Fermi tem-perature TF; it denotes the onset of a crossover to adegenerate quantum regime. For liquid helium with n�1022 cm−3, the appropriate temperature is of order 1 Kfor both isotopes �4He is a boson and 3He is a fermion�.A typical bosonic dilute alkali-metal gas �such as 7Li,23Na, and 87Rb� has a much lower Tc of order100–1000 nK, because of the larger atomic mass and re-duced number density �n�1013 cm−3 is a typical valuefor dilute cold gases�. This article will focus on suchbosonic systems.

The presence of Bose-Einstein condensation meansthat a single quantum state has macroscopic occupation.This particular quantum state acts like a particle reser-voir that can absorb or emit excited particles with neg-ligible change in its own properties. Thus it is natural touse the grand canonical ensemble in which the system ofinterest is assumed to be in equilibrium with a reservoirat temperature T and chemical potential that deter-mine the mean total energy and mean total number ofparticles.

A. Ideal Bose gas

Consider an ideal Bose gas in an external trap poten-tial Vtr, with a complete set of quantum-mechanicalsingle-particle energies �j, and assume that the system isin equilibrium at temperature T and chemical potential. In the grand canonical ensemble, the mean occupa-tion of the jth state is

nj =1

exp����j − �� − 1� f��j� , �2.1�

where �= �kBT�−1 is proportional to the inverse tem-perature. Here f���= �exp����−��−1−1 is the usualBose-Einstein distribution function �it depends explicitlyon T and �, as shown, for example, in Sec. V of Fetterand Walecka �2003�, Chap. 5 of Lifshitz and Pitaevskii�1980a�, or Chap. 7 of Pathria �1996�. A detailed analysisyields the mean total number of particles N�T ,� andthe mean total energy E�T ,�,

N�T,� = j

f��j� , �2.2�

E�T,� = j�jf��j� . �2.3�

In principle, the first relation �2.2� can be inverted togive �T ,N�, and substitution into Eq. �2.3� then yieldsthe mean total energy E�T ,N� as a function of tempera-ture and the total number of particles.

It is useful to introduce the density of states

g��� = j��� − �j� , �2.4�

so that Eqs. �2.2� and �2.3� reduce to

N�T,� =� d�g���f��� , �2.5�

E�T,� =� d�g����f��� . �2.6�

Although g��� is formally singular, a smoothed �coarse-grained� version would be well defined.

In the classical limit, the chemical potential cl

=kBT ln�n�T3 � is large and negative since n�T

3 �1. As thetemperature decreases �or the density increases�, how-ever, increases toward positive values. The criticaltemperature Tc for the onset of BEC is defined implic-itly through the relation �Tc ,N�=�0, where �0 is theground-state energy of the single-particle Hamiltonianwith the potential Vtr. For T Tc, the chemical potentialremains fixed at this value, and the ground state has amacroscopic occupation N0�T�, whose temperature de-pendence follows from the conservation of total par-ticles N=N0�T�+N��T�, with

N��T� = ��0

d�g���

exp���� − �0�� − 1�2.7�

defining the total number of particles not in the conden-sate. If g��0� vanishes, then this integral is finite and welldefined, and the number of noncondensed particles de-creases with decreasing temperature. In contrast, if g��0�is finite or singular, then this integral for the number ofnoncondensed particles diverges, and the system cannotform a BEC. It is helpful to examine two specific ex-amples.

1. Ideal BEC in a box with periodic boundary conditions

The textbook example of a BEC is an ideal gasin a three-dimensional cubical box of volume V=L3

with periodic boundary conditions. The single-particleeigenfunctions are plane waves �k�r�=V−1/2 exp�ik ·r�with single-particle energy �k=�2k2 / �2M�, where k= �2� /L��nx ,ny ,nz� with nj any integer. It is not hard toobtain the corresponding density of states

g��� =V

4�2�2M

�2 3/2

�1/2. �2.8�

In a box with periodic boundary conditions, the lowestsingle-particle energy is �0=0, which corresponds to k=0 �a uniform state with density ��0�2=V−1�. A standardanalysis with the conventional definition �T

2

=2��2 /MkBT yields the condition

649Alexander L. Fetter: Rotating trapped Bose-Einstein condensates

Rev. Mod. Phys., Vol. 81, No. 2, April–June 2009

Page 4: Rotating trapped Bose-Einstein condensatespjh/teaching/phz7429/fetter.pdfII. Physics of Bose-Einstein Condensates in Dilute Trapped Gases 648 A. Ideal Bose gas 649 1. Ideal BEC in

n�Tc

3 = ��32 � 2.612, �2.9�

which is the critical phase-space density for the onset ofBEC. Alternatively, the critical temperature Tc is givenby

kBTc =2�

���3

2 �2/3

�2n2/3

M� 3.31

�2n2/3

M, �2.10�

as anticipated from the previous qualitative discussion.Below Tc, the �uniform� condensate with zero momen-tum has a macroscopic occupation number, which is thesignature of BEC. In particular, it is easy to verify that

N��T�N

= � T

Tc 3/2

�2.11�

for T Tc, so that the macroscopic occupation of theuniform ground state has the temperature dependence

N0�T�N

= 1 − � T

Tc 3/2

. �2.12�

It is worth generalizing these results to a uniform idealBose gas in a d-dimensional hypercubical box; the cor-responding density of states g��� is proportional to �d/2−1.This alteration means that the integral in Eq. �2.7� di-verges for d�2. As a result, the limit →�0 requires amore careful analysis, leading to the conclusion thatsuch a uniform gas in two or one dimension cannot forma BEC; for a detailed treatment of the two-dimensionaluniform Bose gas, see Pethick and Smith �2002�.

2. Ideal BEC in a harmonic trap

In three dimensions, an anisotropic harmonic trap po-tential

Vtr�r� = 12M��x

2x2 + �y2y2 + �z

2z2� �2.13�

yields single-particle energies that are labeled by a trip-let of non-negative integers nx, ny, nz,

�nx,ny,nz= ��nx�x + ny�y + nz�z� + �0, �2.14�

where the second term is the zero-point energy �0

= 12���x+�y+�z�. The ground-state wave function is a

product of three one-dimensional Gaussians with meanwidths dj=�� /M�j �here j=x, y, or z�. If the sums in thedensity of states are approximated by integrals �whichholds for large ��, the corresponding density of states is

g��� =�2

2�3�03 , �2.15�

where �03=�x�y�z defines a geometric-mean trap fre-

quency.The onset of BEC in a harmonic trap occurs at =�0,

with the transition temperature

kBTc = ���3��−1/3��0N1/3 � 0.94��0N1/3, �2.16�

where ��3��1.212. Below Tc, a macroscopic number ofparticles occupies the anisotropic Gaussian ground stateof this harmonic trap, which has a macroscopic occupa-tion with temperature dependence

N0�T�N

= 1 − � T

Tc 3

. �2.17�

Images of this temperature-dependent Gaussian con-densate density profile provided the first clear evidencefor the creation of BEC in 87Rb �Anderson et al., 1995�.Typical traps have d0=�� /M�0 approximately a few mand N�106, confirming the previous value Tc�100–1000 nK �depending on the atomic mass�.

For a harmonic trap in d dimensions, the density ofstates g��� is proportional to �d−1 �Bagnato and Klepp-ner, 1991�, which differs from the d dependence for ahypercubical box. As a result, a two-dimensional Bosegas in a trap can form a BEC with a finite transitiontemperature Tc, in contrast to a uniform two-dimensional gas in a box. This behavior is important inthe limit of rapidly rotating Bose-Einstein condensates,when the centrifugal forces flatten the atomic system,producing an effectively two-dimensional trapped gas.Pethick and Smith �2002� discussed this situation insome detail, especially the behavior in a one-dimensional trap where the order of limits N→� andT→0 becomes crucial.

B. Laser trapping and optical lattices

The response of an atom to electromagnetic radiationis an old subject �see, for example, Allen and Eberly�1987� and Cohen-Tannoudji et al. �1998��, yet it hasmany important applications to the interaction betweenultracold dilute atoms and laser beams. To understandthe basic physics, place an electric dipole with moment pin an external electric field E. Standard electrostaticsyields the interaction energy Uint=−p ·E, where the mi-nus sign means that the lowest-energy configuration hasthe dipole moment oriented along E.

This expression holds for a fixed dipole moment p, butan atom typically acquires its electric dipole moment asa result of induced polarization. In such a case, the di-pole moment is proportional to the applied field withp=��0E in SI units, where �0�8.85�10−12 F/m is thepermittivity of empty space and � is the atomic polariz-ability. As a very simple classical model, a groundedconducting sphere of radius b has a polarizability �=4�b3 �three times its volume�, and this picture providesa useful order-of-magnitude estimate for atomic polariz-abilities. The total interaction energy of the atom withthe external field then builds up from Ei=0, and integra-tion yields the final expression

Uint = − 12��0E2. �2.18�

650 Alexander L. Fetter: Rotating trapped Bose-Einstein condensates

Rev. Mod. Phys., Vol. 81, No. 2, April–June 2009

Page 5: Rotating trapped Bose-Einstein condensatespjh/teaching/phz7429/fetter.pdfII. Physics of Bose-Einstein Condensates in Dilute Trapped Gases 648 A. Ideal Bose gas 649 1. Ideal BEC in

If the electric field E�r� is static and nonuniform, thenUint�r�=− 1

2��0E2�r� is also nonuniform. Thus an atomexperiences a force that seeks to minimize the interac-tion energy, placing the atom at the maximum of E2 be-cause the static polarizability is positive. This result alsoholds for low frequencies, but the situation eventuallychanges because of induced transitions among atomicenergy levels. The Drude model of a bound electronwith resonant frequency �0 and damping time � subjectto an oscillatory driving field yields the frequency-dependent polarizability

���� =��0��0

2

�02 − �2 − i�/�

, �2.19�

where ��0� is the �positive� static polarizability. Notethat the real part of ���� changes from positive to nega-tive when the applied frequency � exceeds the resonantfrequency �0. The combined relation Uint��� from Eqs.�2.18� and �2.19� is sometimes called the ac Stark effect.

These ideas apply directly to the optical trapping ofatoms by a focused laser. If the frequency � of the laseris less than that of the atomic resonance �“red detun-ing”� with Re �����0, then Eqs. �2.18� and �2.19� indi-cate that the atoms preferentially move toward thesquared-field maxima, whereas if the laser frequency ishigher than the resonant frequency �“blue detuning”�with Re ���� 0, then the atoms move toward thesquared-field minima. In practice, it is necessary to usefar-off-resonance lasers to avoid dissipation. For thealkali-metal atoms of interest here �Li, Na, K, Rb, Cs�,the lowest s to p transition of the valence electron pre-dominates, as in the familiar yellow emission spectrumof Na. Stamper-Kurn et al. �1998� successfully trap a pre-existing 23Na BEC with a focused infrared laser beam.Unlike magnetic traps, which confine only certain hyper-fine states, these optical traps confine all such states.This experimental technique allows the study of“spinor” condensates �see Sec. VII.B.3� that involvecoupled BECs, one for each hyperfine component of agiven F multiplet �Ho, 1998; Ohmi and Machida, 1998�.

A more subtle example is the formation of an opticalstanding wave to trap atoms in a periodic array.Consider a laser beam propagating along the z axiswith electric field E�z , t�=E0 cos�kz−�t�. If the laserbeam reflects normally off a mirror at z=0, theresulting standing wave E0�cos�kz−�t�−cos�kz+�t��=2E0 sin kz sin �t yields the corresponding squaredelectric field �E�z , t��2=4�E0�2 sin2 kz sin2 �t. This peri-odic lattice has a spatial period 1

2�, where �=2� /k is thewavelength of the original laser beam. For either sign ofdetuning, the atoms assume a spatially periodic configu-ration in this optical lattice, as emphasized by Jaksch etal. �1998�. This seminal idea has wide applications in thestudy of ultracold dilute gases, as discussed, for example,by Pitaevskii and Stringari �2003� and Bloch et al. �2008�.The strength of the laser field can be controlled exter-nally, so that the height of the barrier for tunneling be-tween adjacent layers can vary from small to large. This

possibility leads to experiments that demonstrate thetransition from superfluid to insulating behavior inthree-dimensional optical lattices �Greiner et al., 2002�.In the present context of one-dimensional lattices,Anderson and Kasevich �1998� were the first to demon-strate such atom trapping. Their vertical lattice experi-ences a gravitational potential difference Mg� /2 be-tween adjacent layers, with associated interlayertunneling at a measured ac Josephson frequency �J�Mg� / �2�� ��Josephson, 1962�; for a full account of thevarious Josephson effects, see Tinkham �1996��.

C. Inclusion of interparticle interactions

Roughly 60 years ago, Bogoliubov �1947� introducedwhat is now seen as the essential physical approximationfor a dilute Bose gas with repulsive interactions. To un-derstand his idea, note that the repulsive interparticleinteraction can be characterized by the positive s-wavescattering length a, which is typically a few nanometersfor the dilute alkali-metal atoms of interest �see Pethickand Smith �2002��. In the low-density limit �a�n−1/3�, thegas parameter na3 is small and provides a natural expan-sion parameter. Bogoliubov noted that the ground stateof the dilute gas will then be close to that of the idealgas. Thus the number of particles N0 in the condensatewill be close to the total number N, and the differencecan be ignored in first approximation. The macroscopicoccupation of the single-particle mode for the conden-sate means that the creation and annihilation operatorsfor this particular mode can be approximated as “classi-cal fields” similar to the electric fields for a laser mode.Bogoliubov’s physical picture underlies almost all subse-quent developments in the field and is basic for under-standing the behavior of a low-temperature BEC.

For a uniform Bose gas at rest, the condensate is thezero-momentum state, and the excited states are labeledby wave number k�0, as discussed, for example, by Lif-shitz and Pitaevskii �1980b�, Pethick and Smith �2002�,and Fetter and Walecka �2003�. In the present case of atrapped condensate, however, the analysis is somewhatmore intricate and relies on the macroscopic condensatewave function � �sometimes called the “order param-eter”�. This condensate wave function describes thesingle-particle mode that has macroscopic occupation,with its squared absolute value ���r��2=n�r� giving thenonuniform condensate particle density n�r�. For a di-lute gas at low temperature with N0�N, the normaliza-tion requires

� dV���r��2 = N0 � N . �2.20�

In the limit of a nearly ideal Bose gas at T=0 K, thespatial form of this condensate wave function follows byminimizing the total energy E, along with the constraintthat the total number of particles N is conserved.Equivalently, one can simply minimize the grand-canonical thermodynamic potential E−N, where isthe chemical potential. To understand this approach, I

651Alexander L. Fetter: Rotating trapped Bose-Einstein condensates

Rev. Mod. Phys., Vol. 81, No. 2, April–June 2009

Page 6: Rotating trapped Bose-Einstein condensatespjh/teaching/phz7429/fetter.pdfII. Physics of Bose-Einstein Condensates in Dilute Trapped Gases 648 A. Ideal Bose gas 649 1. Ideal BEC in

rely on physical arguments, but many-body quantum-field theory provides a rigorous foundation for the sameresults �Hohenberg and Martin, 1965; Fetter, 1972,1996�. In a trap with confining potential Vtr, the Gross-Pitaevskii �GP� �Gross, 1961; Pitaevskii, 1961� energyfunctional for the condensate has the form

EGP��� =� dV��2����2

2Mkinetic

+ Vtr���2

trap

+12

g���4

interaction��2.21�

containing the kinetic energy, the trap energy �see Eq.�2.13��, and the interaction energy, respectively. The firsttwo terms �those quadratic in �� are just the one-bodyenergy for an ideal Bose gas in a �usually� harmonictrap. In contrast, the quartic term �the two-body energy�describes the effect of interactions, where the interpar-ticle potential has been approximated by a short-range “pseudopotential” V�r−r���g��3��r−r��, with g=4�a�2 /M a coupling constant fixed by the s-wave scat-tering length a �this result follows from standard two-body quantum-mechanical scattering theory; see, forexample, Pethick and Smith �2002� and Fetter and Wa-lecka �2003��.

Comparison of the kinetic and trap energies for a har-monic potential yields the familiar oscillator length

d0 =� �

M�0�2.22�

that characterizes the mean size of the noninteractingcondensate, with �0= ��x�y�z�1/3. Similarly, comparisonof the kinetic energy and the interaction energy for auniform condensate yields the “healing length”

� =�

�2Mgn=

1�8�an

, �2.23�

which characterizes the length scale over which a local-ized alteration in the condensate density heals back toits uniform value n. In this context, recall the Bogoliu-bov approximation of small quantum depletion N��N,which requires na3�1. As a result, Eq. �2.23� has theimportant corollary

n2/3�2 =1

8�n1/3a 1, �2.24�

namely, the healing length must exceed the interparticlespacing in any GP description.

Variation of EGP��� in Eq. �2.21� with respect to �* atfixed normalization yields the time-independent GPequation �Gross, 1961; Pitaevskii, 1961�

�−�2�2

2M+ Vtr�r� + g���r��2���r� = ��r� . �2.25�

Here the chemical potential can be considered either aLagrange multiplier or a parameter in the zero-temperature grand-canonical thermodynamic potentialE−N. In the present context, this GP equation is es-

sentially a nonlinear Schrödinger equation that includesan additional quadratic self-coupling g���2; such a termcan be interpreted as a Hartree potential VH�r�=gn�r�that represents the interaction with the local nonuniformcondensate density.

Compared to a uniform BEC, a trapped condensateinvolves an additional characteristic length d0 �the non-interacting condensate size�. A simple scaling argument�see Pethick and Smith �2002�� yields a new dimension-less parameter Na /d0 that characterizes the importanceof the interaction in a trapped condensate. As a result,the three terms in Eq. �2.21� can have very differentmagnitudes depending on the parameters. In the usualsituation �N�106, a approximately a few nm, and d0 ap-proximately a few m�, this dimensionless parameterNa /d0 is large, and the resulting regime is known as theThomas-Fermi limit �Baym and Pethick, 1996�1 In thiscase, the repulsive interactions dominate and expand thecondensate to a mean radius R0 that greatly exceeds themean oscillator length d0 �a factor of 10 is typical�. Thisexpansion dramatically reduces the radial gradient ofthe density, and the associated kinetic energy thus be-comes negligible relative to the trap energy and the in-teraction energy. Equation �2.21� then reduces to

ETF��� � � dV�Vtr���2 +12

g���4 , �2.26�

which involves only ���2 and ���4. Minimization with re-spect to ���2 at fixed normalization gives the Thomas-Fermi �TF� approximation

Vtr�r� + g���r��2 = , �2.27�

which also follows by omitting the kinetic energy term inEq. �2.25�. This algebraic equation can be solved directlyfor the equilibrium density

n�r� = − Vtr�r�

g�� − Vtr�r�� , �2.28�

where ��x� is the unit positive step function. Althoughthis expression holds for any reasonable trap potential,the usual case of a quadratic harmonic trap �2.13� yieldsthe simple and explicit expression

n�r� = n�0��1 −x2

Rx2 −

y2

Ry2 −

z2

Rz2 , �2.29�

where the right-hand side is positive and zero otherwise.Here n�0�= /g is the central density and Rj

2=2 /M�j2

are the squared condensate radii in the three coordinatedirections. This TF density has a parabolic cross sectionand fills the interior of an ellipsoid.

Application of the normalization condition �2.20� tothe TF density �2.29� yields N=8�n�0�R0

3 /15, where R03

�RxRyRz depends on the chemical potential . Some

1The limit Na /d01 means that the condensate density iseffectively uniform on relevant length scales, which explainsthe name Thomas-Fermi.

652 Alexander L. Fetter: Rotating trapped Bose-Einstein condensates

Rev. Mod. Phys., Vol. 81, No. 2, April–June 2009

Page 7: Rotating trapped Bose-Einstein condensatespjh/teaching/phz7429/fetter.pdfII. Physics of Bose-Einstein Condensates in Dilute Trapped Gases 648 A. Ideal Bose gas 649 1. Ideal BEC in

algebra shows that n�0�= /g=R02 /8�ad0

4, and a combi-nation of these results yields the important dimension-less relation

R05

d05 = 15

Na

d0, �2.30�

which is large in the present TF limit. Correspondingly,the TF chemical potential becomes

TF =12

M�02R0

2 =12��0

R02

d02 , �2.31�

so that TF��0 in the TF limit. Since TF is propor-tional to N2/5, the thermodynamic relation =�E /�Nyields the TF energy for a trapped condensate

ETF = 57TFN , �2.32�

which also follows by direct integration of Eq. �2.26� forthe TF density profile.

It is conventional to use the central density n�0� todefine the healing length �Eq. �2.23�� in a nonuniformtrapped condensate, which gives

�2 =1

8�an�0�. �2.33�

These relations lead to the conclusion that �2=d04 /R0

2;equivalently,

d0=

d0

R0. �2.34�

This TF limit holds when the mean condensate radius R0is large compared to the mean oscillator length d0. Thusthe TF oscillator length d0 is the geometric mean of �and R0, and Eq. �2.34� then yields a clear separation ofTF length scales, with ��d0�R0.

Equation �2.25� has the expected time-dependent gen-eralization �known as the time-dependent GP equation�

i����r,t�

�t= �−

�2�2

2M+ Vtr�r� + g���r,t��2���r,t� ,

�2.35�

where � now depends on t as well as on r. Although Iwill not discuss the field-theoretical derivation of thetime-dependent GP equation, it is worth noting that thecondensate wave function is interpreted as an ensemble

average ��r , t�= ���r , t�� of the field operator ��r , t��Gross, 1961; Pitaevskii, 1961; Hohenberg and Martin,1965; Fetter, 1972, 1996�. Comparison of Eqs. �2.25� and�2.35� implies that a stationary solution has the time de-pendence exp�−it /��. This result also follows directlyfrom the field-theoretical definition because the stateson the left-hand side of the ensemble average have oneless particle than those on the right-hand side.

This nonlinear field equation �2.35� can be recast in anintuitive hydrodynamic form by writing the condensatewave function

��r,t� = ���r,t��exp�iS�r,t�� �2.36�

in terms of its magnitude ��� and phase S. In this picture,the condensate �particle� density is n�r , t�= ���r , t��2,whereas the particle-current density becomes

j =�

2Mi��* � � − � � �*� = ���2

� � S

M. �2.37�

The usual hydrodynamic relation j=nv identifies the lo-cal velocity as

v�r,t� =�

M� S�r,t� = ���r,t� , �2.38�

where ���S /M is the velocity potential, as proposedby Feynman �1955� in his discussion of the quantum me-chanics of vortices in superfluid He II. Note that ��vvanishes wherever S is not singular, so that the “super-fluid velocity” v is irrotational, as assumed by Landau�1941� in his phenomenological two-fluid hydrodynamicsfor superfluid He II. For more complete discussions, see,for example, Lifshitz and Pitaevskii �1980b� and Landauand Lifshitz �1987�.

The representation of the velocity v in terms of thegradient of the phase has one major implication. Con-sider the “circulation” �, defined at a given instant oftime in terms of a line integral �=�Cdl ·v around aclosed path C. Substitution of Eq. �2.38� yields

� =�

M�C

dl · �S =�

M��S�C, �2.39�

where ��S�C is the net change in the phase on followingC. The single valuedness of the condensate wave func-tion requires that this quantity must be an integral mul-tiple of 2�, which shows that the circulation in a diluteBEC must be quantized in units of 2�� /M �Onsager,1949; Feynman, 1955�. For accounts of this remarkablequantum-mechanical relation, see London �1954� andDonnelly �1991�.

Substitute Eq. �2.36� into the time-dependent GPequation �2.35�. The imaginary part yields the familiarcontinuity equation for a compressible fluid

�n

�t+ � · �nv� = 0. �2.40�

Correspondingly, the real part provides a generalizedBernoulli equation for this quantum fluid

12

Mv2 + Vtr −�2

2M�n�2�n + gn + M

��

�t= 0, �2.41�

where the explicitly quantum-mechanical term contain-ing �2 is sometimes called a “quantum pressure.” Thefull structure of this Bernoulli equation �including thequantum pressure� can be understood as that for an ir-rotational compressible isentropic fluid �see Landau andLifshitz �1987�� with an enthalpy density obtained fromEq. �2.21� �Fetter, 2002�. Such an isentropic picture isclearly appropriate for a low-temperature superfluid.

653Alexander L. Fetter: Rotating trapped Bose-Einstein condensates

Rev. Mod. Phys., Vol. 81, No. 2, April–June 2009

Page 8: Rotating trapped Bose-Einstein condensatespjh/teaching/phz7429/fetter.pdfII. Physics of Bose-Einstein Condensates in Dilute Trapped Gases 648 A. Ideal Bose gas 649 1. Ideal BEC in

The existence of a Bernoulli equation for the time-dependent GP equation has one basic consequence inthe present context of vortices in dilute quantum gases.All classical results about vortex dynamics in an irrota-tional fluid �such as Kelvin’s theorem on the conserva-tion of circulation, proved, for example, by Landau andLifshitz �1987�� automatically hold for any vortex con-figuration in a dilute BEC. Nevertheless, the presence ofa nonuniform trapping potential Vtr�r� and the resultingnonuniform density in Eq. �2.28� significantly affect thedynamics of vortices, so that qualitative classical picturesbased on uniform incompressible fluids may not alwaysapply.

D. Bogoliubov equations

The time-dependent GP equation �2.35� describes thedynamics of the condensate, and it is natural to considerthe linearized behavior of small perturbations aroundsuch solutions ��r , t�. Although a fully quantum-mechanical analysis is feasible �Bogoliubov, 1947; Pit-aevskii, 1961; Fetter, 1972�, I rely on a simpler classicalapproach �Dalfovo et al., 1999� �see also Pethick andSmith �2002�� that yields the same eigenvalue equations.Assume a stationary nonuniform condensate with wavefunction ��r�e−it/�. Since the nonlinear Hartree terminvolves ���2, the perturbations must, for consistency, in-clude e�i�t with both signs of the frequency. Specifically,take

��r,t� = e−it/����r� + u�r�e−i�t − v*�r�ei�t� , �2.42�

where the Bogoliubov amplitudes u�r� and v�r� aretreated as small.

Substitute Eq. �2.42� into Eq. �2.35� and collect first-order terms proportional separately to e�i�t. The resultis a pair of linear eigenvalue equations:

Luj − g���2vj = ��juj, �2.43�

Lvj − g��*�2uj = − ��jvj, �2.44�

where L=−�2�2 / �2M�+Vtr−+2g���2 is a Hermitianoperator. These coupled equations are known as the Bo-goliubov equations, although Pitaevskii �1961� was thefirst to consider the application to nonuniform systems�sometimes called the Bogoliubov–de Gennes equationsbecause of their close correspondence to similar equa-tions in the theory of superconductivity, as discussed byde Gennes �1966� and Tinkham �1996��. In many ways,the Bogoliubov equations are analogous to a nonrelativ-istic version of the Dirac equation, with u and v as theparticle and hole amplitudes, including the ��,�� metricseen in the minus sign on the right-hand side of Eq.�2.44� compared to that of Eq. �2.43� �see, for example,Schiff �1968�, Shankar �1994�, and Abers �2004��.

For a uniform condensate, the appropriate eigenfunc-tions are plane waves labeled by k. The Bogoliubov en-ergy eigenvalue is determined to be

Ek = �gn�2k2

M+ ��2k2

2M 2�1/2

, �2.45�

where n is the condensate density and �gn for thisuniform gas. For long wavelengths �k��1�, where � isthe healing length from Eq. �2.23�, the Bogoliubov en-ergy reduces to a phonon spectrum

Ek � �sk , �2.46�

with s=�gn /M the speed of compressional sound. Forshort wavelengths �k�1�, in contrast, the Bogoliubovspectrum becomes that of a free particle. The existenceof a linear �phonon� spectrum at long wavelengths iscrucial for superfluidity because the Landau �1941� criti-cal velocity vc for onset of dissipation is the minimumvalue of Ek /�k with respect to k. In the present case, theBogoliubov excitation spectrum �2.45� yields vc=s. Notethat s��g is real and positive only because of the repul-sive interactions �g�0�. Furthermore, vc=s vanishes foran ideal Bose gas with g=0. Hence the ideal Bose gas isnot strictly superfluid even at zero temperature, al-though it does indeed have a condensate with k=0. Inthis well-defined sense, the existence of superfluidity in auniform dilute Bose gas arises from the repulsive inter-actions.

Manipulation of the Bogoliubov equations shows that�j�dV��uj�2− �vj�2� is real. In most cases, the integral isnonzero, so that �j itself is real. To understand the basicissue involved in the normalization of the Bogoliuboveigenfunctions, it helps to recall the usual particle cre-ation and annihilation operators aj

† and aj fromquantum-field theory. These particle and hole operatorsobey the familiar Bose-Einstein commutation relations�aj ,aj�

† �=�j,j�. The solution of the Bogoliubov equationsconstitutes a canonical transformation to a new set of“quasiparticle” operators �j

† and �j that are linear com-binations of the original particle and hole operators. Thesituation is similar to that for BCS superconductors, asshown, for example, by de Gennes �1966� and Tinkham�1996�. Almost by definition, any canonical transforma-tion preserves the commutation relations so that the Bo-goliubov quasiparticle operators must obey the sameBose-Einstein commutation relations ��j ,�j�

† �=�j,j�. Adetailed analysis shows that physically relevant Bogoliu-bov eigenfunctions must satisfy the following positivenormalization condition:

� dV��uj�2 − �vj�2� = 1. �2.47�

Under this canonical transformation, the Bogoliobovapproximate grand-canonical Hamiltonian assumes thesimple form

654 Alexander L. Fetter: Rotating trapped Bose-Einstein condensates

Rev. Mod. Phys., Vol. 81, No. 2, April–June 2009

Page 9: Rotating trapped Bose-Einstein condensatespjh/teaching/phz7429/fetter.pdfII. Physics of Bose-Einstein Condensates in Dilute Trapped Gases 648 A. Ideal Bose gas 649 1. Ideal BEC in

H − N � const + j

���j�j†�j, �2.48�

where the sum runs over all states with positivenormalization.2 It is clear that the sign of the eigenfre-quencies is crucial for the stability of the system. If allthe frequencies �j are positive, then creation of quasi-particles raises the energy, and the system is manifestlystable. If, however, any of the positive-norm solutionshas a negative frequency, then creation of those particu-lar quasiparticles lowers the energy, and the system issaid to display an “energy instability.”

A simple example of this situation is a uniform con-densate that moves with velocity v=�q /M. The conden-sate wave function now has a nonuniform complexstructure ��r�=�neiq·r, and the solution with positivenormalization has the excitation energy �Pollock, 1967;Fetter, 1972; Rokhsar, 1997a�

Ek�v� = �k · v + Ek, �2.49�

where Ek is the Bogoliubov energy �2.45� for a station-ary condensate. For small v, this excitation energy ispositive for any k. When v exceeds the speed of sound s,however, this eigenvalue will be negative in certain di-rections, which represents the onset of the Landau insta-bility associated with spontaneous creation of phonons�Landau, 1941�. Although the energy instability impliesthat the system is formally unstable, the absence of dis-sipation means that no mechanism exists to support thegrowth of the unstable process.

For some special condensed systems �typically whenBEC occurs in an excited single-particle state�, the lin-earized Bogoliubov equations also have zero-norm solu-tions with �dV��u�2− �v�2�=0. In these cases, the frequen-cies can be �and usually are� complex. Such systems aresaid to exhibit a “dynamical instability,” and the un-stable mode�s� will grow exponentially until various non-linearities limit the evolution. Such dynamically unstablebehavior occurs in a condensate with a doubly quantizedvortex �Pu et al., 1999; Shin et al., 2004�, in certain one-dimensional solitons �Fedichev et al., 1999; Muryshev etal., 1999; Feder et al., 2000� and in one-dimensional op-tical lattices �Wu and Niu, 2001�.

III. PHYSICS OF ONE VORTEX (A FEW VORTICES) IN ATRAP

When the GP equation was developed in 1961 to pro-vide a tractable model for a vortex line in a BEC, rotat-ing superfluid 4He was the only system of experimentalinterest �even though 4He is a dense highly correlatedsystem�. Since 1995, however, dilute ultracold trappedalkali-metal gases have provided remarkable physicalsystems that indeed fit the GP picture in considerabledetail.

A. One vortex in unbounded condensate

Gross �1961� and Pitaevskii �1961� introduced the GPequation specifically to describe a single straight vortexin an otherwise uniform dilute Bose-Einstein conden-sate with bulk particle density n. The condensate wavefunction has the form ��r�=�n �r�, with

�r� = ei!f�r/�� , �3.1�

where r, ! are two-dimensional plane-polar coordinatesand f→1 for large r. Equation �2.38� gives the circulat-ing velocity of a singly quantized long straight vortexline

v�r� =�

Mr�; �3.2�

this hydrodynamic flow has circular streamlines and amagnitude that diverges as r→0. The circulation �= �dl ·v=2�� /M can be transformed with Stokes’s theo-rem as �dS ·��v=2�� /M, implying a singular localizedvorticity at the center of the vortex core,

� � v =2��

M��2��r�z . �3.3�

The kinetic energy per unit length of the vortex is

�2

2M� d2r����2 =

�2n

2M� d2r��df

dr 2

+f2

r2� , �3.4�

where the first term arises from the density variationnear the vortex core and the second term�= 1

2M�d2rn�f�r��2v2� is the kinetic energy of the circulat-ing flow. With this kinetic-energy functional, the Euler-Lagrange equation yields the nonlinear GP equation forthe radial amplitude f�r�. The resulting centrifugal bar-rier forces the amplitude f�r� to vanish linearly for r"�,which characterizes the vortex core, and f�r��1 for r�. In the present dilute limit, Eq. �2.24� shows that thevortex core is significantly larger than the interparticlespacing. In addition, the particle-current density �j�r��=n�f�r��2� /Mr vanishes both far from the vortex and atthe center of the vortex, with a maximum near r��.Finally, the speed of sound in a dilute Bose gas can berewritten as s=�gn /M=� /�2M�, so that the circulatingflow in Eq. �3.2� becomes supersonic for r"�; in thissense, local acoustic cavitation can be considered thesource of the vortex core.

2The criterion for inclusion in this sum has been a source ofconfusion because every acceptable solution of the Bogoliubovequation �uj ,vj with frequency �j and positive normalizationhas a second “antiparticle” solution �vj

* ,uj* with frequency −�j

and negative normalization. The original study of these rela-tions �Fetter, 1972� incorrectly assumed that the positive-normsolution always had positive energy. This result indeed de-scribes the simplest case of a uniform stationary condensatewith plane-wave eigenfunctions. In contrast, nonuniform con-densates, such as one in uniform motion or one containing avortex �see Sec. III.B.2 and Rokhsar �1997a�� can have physicalstates with positive normalization and negative frequency. Fora more general discussion, see Fetter and Svidzinsky �2001�.

655Alexander L. Fetter: Rotating trapped Bose-Einstein condensates

Rev. Mod. Phys., Vol. 81, No. 2, April–June 2009

Page 10: Rotating trapped Bose-Einstein condensatespjh/teaching/phz7429/fetter.pdfII. Physics of Bose-Einstein Condensates in Dilute Trapped Gases 648 A. Ideal Bose gas 649 1. Ideal BEC in

In terms of the dimensionless scaled variable x=r /�,the variational approximation fvar�x�=x /�x2+2 �Fetter,1969� provides a good fit to the numerical solution of theradial GP equation �see, for example, Fig. 9.1 of Pethickand Smith �2002��. The vortex produces an additionalGP energy per unit length Ev����2n /M�ln�1.46R /��,where R is a large-distance cutoff. Apart from the addi-tive numerical constant, this result is essentially the in-tegral of 1

2Mv2n over a large circle of radius R with ashort-distance cutoff of order � �Ginzburg and Pit-aevskii, 1958�.

B. One vortex in a trapped condensate

A classical viscous fluid in a container that rotateswith angular velocity � rapidly acquires the same angu-lar velocity because of the microscopically rough walls.From a theoretical view, the laboratory is no longer theappropriate reference frame because the rotating wallsare a time-dependent external potential that can dowork on the system. Thus it is essential to transform tothe frame rotating with the walls, in which case the ex-ternal potential becomes stationary. Consider an origi-nally static Hamiltonian H�r ,p� with an external trappotential Vtr�x ,y ,z� that in general is anisotropic about

�. When the potential rotates with angular velocity �, adirect transformation of the Lagrangian from the labo-ratory frame to the rotating frame eventually yields theHamiltonian H��r� ,p�� in the rotating frame

H��r�,p�� = H�r�,p�� − � · L�r�,p�� , �3.5�

where r� and p� are the coordinates and canonical mo-menta in the rotating frame. Here the right-hand sideinvolves the nonrotating Hamiltonian H�r� ,p�� and an-gular momentum L�r� ,p��=r��p�, with the originalvariables replaced by those in the rotating frame �seeLandau and Lifshitz �1960� and Lifshitz and Pitaevskii�1980a��.3

1. Gross-Pitaevskii energy in a trap

In the context of the GP equation, this transformationto a rotating frame yields a modified energy functional

E���� =� dV� �2

2M����r��2 + Vtr�r����r��2 +

12

g���r��4�− � ·� dV�*�r�r � p��r� , �3.6�

where all variables here are in the rotating frame �writ-ten without primes for notational simplicity� �compareEq. �3.5��. The minimization of E���� now explicitly in-volves the angular velocity, and the last term �−� ·L�clearly favors states with nonzero �positive� angular mo-mentum.

Most experiments on vortices in rotating dilute BECsoccur in the TF regime. Thus I concentrate on this situ-ation, which has several simplifying features. Specifically,the dominant repulsive interactions imply that the meancondensate radius R0 is large relative to the originalmean oscillator length d0. Equation �2.34� then impliesthat the vortex core size ���� is smaller than d0 by thesame ratio d0 /R0. Thus the presence of a few vorticesdoes not significantly affect the number density, whichcan be taken as the unperturbed TF shape in Eq. �2.29�.In most cases, the condensates are axisymmetric withradial and axial trap frequencies �� and �z, so that thenonrotating TF density depends only on r=�x2+y2 andz.

If an axisymmetric condensate contains a vortex at thecenter, the general wave function has the form �Dalfovoand Stringari, 1996; Lundh et al., 1997; Sinha, 1997�

��r� = ei!���r,z�� , �3.7�

with the phase ! appropriate for a singly quantized vor-tex. The hydrodynamic velocity is again v=�� /Mr, as inEq. �3.2� for a singly quantized vortex in an asymptoti-cally uniform fluid. The resulting centrifugal barrier nearthe axis of symmetry creates an axial node in the con-densate wave function; thus the presence of one vortexfundamentally alters the topology of the condensatedensity from ellipsoidal to toroidal because of the holealong the vortex core.

It is important to generalize this wave function to anoff-center vortex. For simplicity, consider a three-dimensional disk-shaped condensate ��z��� with ra-dial condensate dimension R�. In this case, a vortex atr0= �x0 ,y0� will remain essentially straight, and the prin-cipal effect of the lateral translation is the altered phase

S�x,y� = arctan�y − y0

x − x0 , �3.8�

which is the azimuthal angle around the shifted origin r0.In the TF regime, the density is taken to be unchanged�the logarithmically divergent circulating kinetic energyis cut off at the vortex core radius ��.

Use the energy functional E���� in Eq. �3.6� and letE0� denote the energy of a vortex-free state �it is inde-pendent of ��. If E1��r0 ,�� is the total GP energy of asingle off-center vortex in the rotating frame, then thedifference of the two energies �E��r0 ,��=E1��r0 ,��−E0� represents the formation energy of the vortex. Astraightforward analysis with this model TF wave func-tion �Svidzinsky and Fetter, 2000a� yields

�E��r0,���E��0,0�

= �1 −r0

2

R�2 3/2

−�

�c�1 −

r02

R�2 5/2

, �3.9�

where

�c =52

MR�2 ln�R�

� , �3.10�

and the detailed form of �E��0,0� is unimportant here.For fixed �, this expression gives the energy in the ro-

3Note that H�r� ,p�� is not simply the original Hamiltonian inthe laboratory frame because of the primed variables.

656 Alexander L. Fetter: Rotating trapped Bose-Einstein condensates

Rev. Mod. Phys., Vol. 81, No. 2, April–June 2009

Page 11: Rotating trapped Bose-Einstein condensatespjh/teaching/phz7429/fetter.pdfII. Physics of Bose-Einstein Condensates in Dilute Trapped Gases 648 A. Ideal Bose gas 649 1. Ideal BEC in

tating frame �E��r0 ,�� as the vortex moves from thecenter of the condensate �r0=0� to the edge �r0=R��,shown in Fig. 1 for various fixed values of � �Packardand Sanders �1972� made a similar analysis for a vortexin an incompressible uniform superfluid; Castin andDum �1999� obtained comparable results numerically fora vortex in a trapped condensate�.

�a� Curve �a� shows the energy �E��r0 ,0� of a vortex ina nonrotating condensate as it moves from the cen-ter to the outer edge. This energy decreases mono-tonically as r0 increases, with negative curvaturefor r0�R� /�2. If there is no dissipation, the dis-tance r0 determines the energy, and the vortex fol-lows a circular trajectory at fixed r0. At low butfinite temperature, however, weak dissipationslowly reduces the energy of the vortex, which thusmoves outward along curve �a�, executing a spiraltrajectory toward the edge of the condensate�Rokhsar, 1997a, 1997b�.

�b� With increasing �, the function �E��r0 ,�� for fixed� flattens and the �negative� central curvature de-creases. Curve �b� shows the case of zero centralcurvature, which occurs for

�m =32

MR�2 ln�R�

� =

35�c. �3.11�

This angular velocity �m represents the onset ofmetastability for a vortex near the center of a ro-tating condensate. For � �m, the negative centralcurvature means that weak dissipation moves thevortex away from the center, and it eventually spi-rals out to the edge. For ���m, however, the cen-tral curvature is now positive, and weak dissipationwill move the vortex back to the center. In thislatter regime, the surrounding barrier makes thevortex locally stable near the trap center, eventhough it is not globally stable because the energy

at the edge is definitely lower than the energy atthe center.

�c� Eventually, at �c defined in Eq. �3.10� the centralenergy �E��0,�c� vanishes �and thus equalsE��R� ,�c��. For ���c, the central vortex is bothlocally and globally stable. This value �c is inter-preted as the thermodynamic critical angular veloc-ity for the creation of a singly quantized vortex inthe TF limit.4 As discussed later, experiments onrotating trapped gases generally find a different�larger� critical angular velocity for the creation ofa vortex, but similar analyses for incompressiblefluids in cylindrical containers do provide a reason-able description of vortex formation in rotating su-perfluid 4He �Packard and Sanders, 1972; Yarm-chuk and Packard, 1982� �see also Donnelly�1991��.

2. Dynamical motion of a trapped vortex

Although the GP energy functional �3.6� describes theequilibrium of a GP condensate with one vortex �ormore vortices�, it also provides a useful dynamical pic-ture of the vortex motion in a trapped condensate.

a. Lagrangian functional

The most direct approach relies on the basic observa-tion that the time-dependent GP equation �2.35� is theEuler-Lagrange equation for the time-dependent La-grangian functional

L��� =� dVi�

2��*

��

�t−

��*

�t� − E���� �3.12�

under variation of �*. If the condensate wave function� depends on one or more parameters, the resultingLagrangian functional provides approximate Lagrangianequations of motion for these parameters. In the contextof BECs, this approach was first applied with great suc-cess to the low-lying normal modes of a trapped station-ary �nonrotating� condensate �Pérez-García et al., 1996�.Subsequently, its application to a single straight vortex ina disk-shaped condensate yields an explicit predictionfor the angular precession of such a trapped vortex

4My treatment generally assumes a dilute gas, with na3�1. Inaddition, the discussion here is in the Thomas-Fermi limitwhere Na /d01 �so that interactions predominate�. For com-pleteness, the corresponding thermodynamic critical angularvelocity for a near-ideal Bose-Einstein gas �Na /d0�1� is�c /���1−Na /�8�dz �Butts and Rokhsar, 1999; Linn andFetter, 1999�. Note that this result also holds for attractive in-teractions �a 0�, but the resulting inequality �c��� impliesthat the radial confinement disappears before the vortex wouldbe stable, in agreement with earlier conclusions �Dalfovo andStringari, 1996; Wilkin et al., 1998�. Indeed, attractive interac-tions quite generally tend to suppress BEC in favor of “frag-mented” states with macroscopic occupation in more than onesingle-particle state �Pollock, 1967; Nozières and Saint James,1982�.

FIG. 1. Dimensionless scaled energy �E��r0 ,�� /�E��0,0� �Eq.�3.9�� of a singly quantized straight vortex in a rotating disk-shaped TF trap as a function of the displacement r0 from thesymmetry axis. Different curves represent different fixed val-ues of the angular velocity �: �a� �=0; �b� �=�m, defined inEq. �3.11�; and �c� �=�c, defined in Eq. �3.10�.

657Alexander L. Fetter: Rotating trapped Bose-Einstein condensates

Rev. Mod. Phys., Vol. 81, No. 2, April–June 2009

Page 12: Rotating trapped Bose-Einstein condensatespjh/teaching/phz7429/fetter.pdfII. Physics of Bose-Einstein Condensates in Dilute Trapped Gases 648 A. Ideal Bose gas 649 1. Ideal BEC in

�Lundh and Ao, 2000; Svidzinsky and Fetter, 2000a; Mc-Gee and Holland, 2001�. In this case, the position r0 ofthe off-center vortex serves as an appropriate parameter,and the Lagrangian formalism shows that the precessionrate is proportional to �the negative of� the slope of theappropriate curve in Fig. 1. For a nonrotating disk-shaped TF condensate, this analysis gives

! =�m

1 − r02/R�

2 , �3.13�

where �m= 32� ln�R� /�� /MR�

2 is the critical rotation fre-quency for the onset of metastability given in Eq. �3.11�.Note that a vortex in a nonrotating condensate precessesin the positive sense, which is the same sense as thecirculating velocity field of the vortex.

In this Lagrangian approach, the factor 1−r02 /R�

2 inthe denominator arises from the parabolic radial TFdensity profile. It means that a vortex near the outerboundary precesses more rapidly than one near the cen-ter, although this simple picture omits bending of thevortex and thus fails as r0→R�.

It is interesting to compare this TF result �Eq. �3.13��to the precession rate for a long straight classical vortexin an incompressible fluid inside a circular cylinder ofradius R. In this latter case, the boundary condition ofzero radial flow velocity at the radius R requires an ex-ternal image vortex, yielding the result �see, for ex-ample, Pethick and Smith �2002��

!cl =�

MR2

1

1 − r02/R2 . �3.14�

Although the two expressions for the precession ratehave the same denominators, the precession here arisesfrom the motion induced by the image �and there is nolarge logarithmic factor�. Since the TF density necessar-ily vanishes at the condensate radius, image vortices aregenerally omitted for trapped condensates, although thisquestion remains partially unresolved �Anglin, 2002;Khawaja, 2005; Mason et al., 2006�

b. Matched asymptotic expansion

An alternative approach to the dynamics of a vortexin a trapped condensate is to study the time-dependentGP equation �2.35� itself; this analysis relies on themethod of matched asymptotic expansions �Pismen andRubinstein, 1991; Rubinstein and Pismen, 1994; see alsoPismen, 1999�. In the TF limit, the vortex core is smallcompared to other length scales, and it is possible to usethe separation of length scales as follows �Svidzinskyand Fetter, 2000b�. Assume that the straight vortex �lo-cated at r0� moves with a locally uniform velocity V�r0�that lies in the x-y plane. Transform to a locally comov-ing reference frame in which the vortex is stationary.The problem is solved approximately in two differentregimes.

�1� Near the vortex core, the short-distance solutionshows that the trap potential exerts a force propor-tional to ��Vtr evaluated at r0. This solution in-

cludes both the detailed nonuniform structure of thevortex core and the asymptotic region �� �r−r0�.

�2� Far from the vortex, the core can be approximatedas a line singularity, and the resulting solution alsoincludes the region �� �r−r0�.

The two solutions must agree in the common regionof overlap, and a lengthy analysis determines the trans-lational velocity V�r0� of the vortex line near the centerof the three-dimensional disk-shaped trap

V�r0� =3�

4Mln�R�

� z � ��Vtr�r0� . �3.15�

To understand the structure of this result, note that thetrap exerts a force −��Vtr on the vortex. Since the vor-tex itself has intrinsic angular momentum, it acts gyro-scopically and hence moves at right angles to this ap-plied force. As a result, the motion follows anequipotential line along the direction z���Vtr and thusconserves energy, as expected for the GP equation atzero temperature.

In fact, this result is essentially the same as that in Eq.�3.13� for r0 near the trap center. The gradient of theaxisymmetric potential Vtr= 1

2M���2 r2+�z

2z2� and thesquared TF condensate radius R�

2 =2 /M��2 readily

yield the equivalent form for Eq. �3.15�,

V�r0� = �mr� , �3.16�

involving the critical angular velocity �m for the onset ofmetastability given in Eq. �3.11�. These expressions forthe translational velocity V can be generalized to treat arotating anisotropic harmonic trap with �x��y, and toinclude the three-dimensional effects of vortex curva-ture �Svidzinsky and Fetter, 2000a, 2000b�.

Although the final result in Eq. �3.15� is simple andphysical, the details of the method of matchedasymptotic expansions are not very transparent. A morephysical picture starts from the conservation of particles� · �nv�=0 in the locally comoving coordinate system. Inthe TF limit, the particle density is explicitly known interms of the trap potential �compare Eq. �2.29��. Sincev��S, this equation can be rewritten as �n ·�S+n�2S=0, which determines the phase S of the condensatewave function in terms of the known density. A pertur-bation expansion with the appropriate azimuthal angle�see Eq. �3.8�� as the leading term yields an additionalphase proportional to ln�r−r0� �Svidzinsky and Fetter,2000b; Sheehy and Radzihovsky, 2004a, 2004b�. The gra-dient of this additional phase evaluated at r0 then fixesthe precessional velocity of the vortex �3.15�. In this lat-ter perspective, the precessional motion can be thoughtto arise from the gradient of the condensate density n�r�evaluated at the position of the vortex �Sheehy andRadzihovsky, 2004a, 2000b�

658 Alexander L. Fetter: Rotating trapped Bose-Einstein condensates

Rev. Mod. Phys., Vol. 81, No. 2, April–June 2009

Page 13: Rotating trapped Bose-Einstein condensatespjh/teaching/phz7429/fetter.pdfII. Physics of Bose-Einstein Condensates in Dilute Trapped Gases 648 A. Ideal Bose gas 649 1. Ideal BEC in

V�r0� =�

M

z � �− �n�2n

ln�R�

� , �3.17�

where the minus sign appears because the density de-creases with increasing r �this detailed result describes atwo-dimensional condensate�. Apart from a numericalfactor that reflects the difference between two and threedimensions, Eqs. �3.15� and �3.17� describe the samephysics because the gradient of the TF density �n isproportional to −�Vtr �see Eq. �2.28��.

c. Bogoliubov equations and stability of a vortex

Soon after the creation of BECs in dilute 87Rb�Anderson et al., 1995�, theorists studied the collective-mode spectrum of a nonrotating stationary vortex-freetrapped condensate using the GP equation �2.25� for thecondensate and the Bogoliubov equations �2.43� and�2.44� for the small-amplitude normal modes. As Ed-wards et al. �1996� reported, the agreement between thetheoretical predictions and the experimental observa-tions was impressive �see also Dalfovo et al. �1999��.

This success rapidly led to a similar numerical study�Dodd et al., 1997� of the collective modes of a nonro-tating trapped condensate containing a singly quantizedvortex. If the vortex is on the axis of symmetry, the con-densate wave function is similar that in Eq. �3.7�, and theconservation of the z component of angular momentumgreatly simplifies the analysis. Remarkably, the spectrumof quasiparticle excitations for the vortex turns out tocontain a very unusual mode �now called “anomalous”�with positive normalization integral �Eq. �2.47��, nega-tive angular momentum ma=−1 relative to that of thevortex, and negative frequency �a �Rokhsar �1997b� hasemphasized this particular identification; see also Fetter�1998��. As noted in Sec. II.D �see Eq. �2.48��, such anegative eigenfrequency implies that a vortex in a non-rotating condensate is formally �energetically� unstable,but the vortex cannot spiral out of the condensate in theabsence of dissipation �compare the discussion concern-ing curve �a� in Fig. 1�. Detailed studies �Linn and Fetter,1999; Svidzinsky and Fetter, 2000a� showed that �a�−�� in the near-ideal limit and �a�−�m in the TFlimit, so that �a remains negative for all couplingstrengths.

The physics underlying the anomalous mode is espe-cially transparent in the weak-coupling �near-ideal� limit,when the condensate with a singly quantized vortex atthe origin has the radial wave function �1�r ,!�� �x+ iy�exp�− 1

2r2�=rei! exp�− 12r2�, with one unit of angular

momentum and an excitation energy ���. Note that �1differs from the true radial ground-state wave function�0�exp�− 1

2r2�, which instead has zero angular momen-tum and zero excitation energy. In principle, the par-ticles in the macroscopic vortex condensate could makea transition to the nonrotating ground state, with achange in angular momentum m=−1 and a change inenergy −���; this transition is just the anomalous modein an ideal Bose-Einstein gas. Such anomalous modestypically occur whenever the condensate occupies a

single-particle state that is not the true ground state�compare the discussion of a moving condensate belowEq. �2.49��.

In the present case of a condensate with a single vor-tex, the anomalous mode has two important conse-quences.

�1� The linearized solution for the jth mode of the time-dependent GP equation in Eq. �2.42� immediatelyyields the corresponding perturbation in the density�nj= ��*uj−�vj�e−i�jt. In particular, a singly quan-tized vortex with � given in Eq. �3.7� has ananomalous-mode density perturbation �na�r , t��exp�i�ma!−�at��, with ma=−1 and negative exci-tation frequency �a=−��a�. This density perturba-tion �na�exp�i���a�t−!�� precesses in the positivesense around the symmetry axis of the condensate ata rate −�a= ��a�. In the TF limit when ��a�=�m�Svidzinsky and Fetter, 2000a�, this rate agrees withthat in Eq. �3.13� and with experiments �Anderson etal., 2000� discussed in Sec. III.D.1.

�2� When the condensate rotates, Eq. �3.5� shows thatthe frequency �j of the jth mode in the laboratoryframe shifts to

�j���� = �j − mj� �3.18�

in the rotating frame, where mj is the correspondingangular quantum number. For the anomalous mode,this result implies that �a���=−��a�+�, which be-comes positive and thus energetically stable when �exceeds ��a�. This conclusion confirms the criticalrotation frequency �m for the onset of metastabilityin the TF regime, as seen in curve �b� of Fig. 1.

C. Small vortex arrays in a trap

In the TF limit, Eq. �3.10� shows that the thermody-namic critical angular velocity for vortex formation ismuch smaller that the trap frequency, with �c /�0

�d02 /R�

2 �apart from logarithmic factors�. The corre-sponding condensate wave function has a 2�-phase sin-gularity associated with a vortex at r0 �see Eq. �3.8��.More generally, a vortex is simply a node in the conden-sate wave function �Feynman, 1955�. In the context ofweakly interacting ultracold BECs with Na /d0�1, Buttsand Rokhsar �1999� exploited this property in their stud-ies of small vortex arrays �Castin and Dum �1999� ob-tained similar vortex configurations with a somewhatdifferent variational approach�.

To understand their procedure, assume that the con-densate is tightly confined in the z direction with a smallaxial oscillator length dz�d�. The remaining two-dimensional isotropic harmonic trap has the one-bodyHamiltonian H0= 1

2 �p2 /M+M��2 r2�. For positive angular

momentum, the low-lying normalized one-body eigen-functions are

659Alexander L. Fetter: Rotating trapped Bose-Einstein condensates

Rev. Mod. Phys., Vol. 81, No. 2, April–June 2009

Page 14: Rotating trapped Bose-Einstein condensatespjh/teaching/phz7429/fetter.pdfII. Physics of Bose-Einstein Condensates in Dilute Trapped Gases 648 A. Ideal Bose gas 649 1. Ideal BEC in

�m�r� =umeim!

��m!�1/2d�

exp�−u2

2

=�m

��m!�1/2d�

exp�−u2

2 , �3.19�

where u=r /d�, �= �x+ iy� /d�=uei!, and m#0. Note that�m is also an eigenfunction of the dimensionless angular-momentum operator l=−i� /�! with eigenvalue m. As asimple example of a possible trial state, consider the lin-ear combination �=c0�0+c1�1, which vanishes at �0��x0+ iy0� /d�=−c0 /c1 �the only node�. The point �0 isthe position of the vortex, since the phase of � increasesby 2� on going around �0 in the positive sense.

For a rotating condensate with positive angular mo-mentum, Butts and Rokhsar �1999� used a general linearcombination of such states �=m#0cm�m with normal-ization m#0�cm�2=1.5 The expectation value of the an-gular momentum per particle is �l=�m#0m�cm�2. Fur-thermore, the expectation value of the one-bodyHamiltonian �H0�=���m#0m�cm�2+���=����l+1� isessentially that of the angular momentum. Thus allstates with the same value of l have the same one-bodyenergy, independent of the specific coefficients cm. To liftthis degeneracy, it is necessary to consider the interac-tion energy Eint=

12g�d3r���4 �assumed to be small�. In

the limit of tight z confinement, this energy reduces to atwo-dimensional integral, which they evaluated numeri-cally with the linear-combination trial state �. For agiven angular momentum l, the equilibrium state mini-mizes the total energy, which also means the interactionenergy because of the degeneracy. In this way, Butts andRokhsar �1999� determined the coefficients cm and thusthe position of the vortices �the nodes of �� as a func-tion of l.

Given the equilibrium energy Etot�l�, the thermody-namic relation ��=�Etot /�l fixes the corresponding an-gular velocity �. Extensive numerical studies �Butts andRokhsar, 1999� then determined the resulting physicalsequence of vortex states with increasing �, as shown inFig. 2. The state with l=1 arises from a single vortex onthe symmetry axis of the condensate, but configurationsof two or more vortices vary with l, so that the corre-sponding angular momentum does not generally take in-teger values.

D. Experimental and theoretical studies in a trap

Two basic experimental methods have been most ef-fective in creating vortices in a BEC. The first approach�Matthews et al., 1999b� manipulates two hyperfine com-ponents of 87Rb, spinning up one component with anexternal coherent electromagnetic coupling beam �Will-iams and Holland, 1999�. This approach relies heavily ontechniques from atomic physics, especially the optical

Bloch equations for a two-level system �see, for ex-ample, Feynman �1965�, Cohen-Tannoudji et al. �1977,1977� and Allen and Eberly �1987��. The second nearlysimultaneous approach �Madison et al., 2000a� is similarto the “rotating-bucket” method of conventional low-temperature physics �see, for example, Donnelly �1991��.In the present context of BECs, a driven rotating defor-mation of the confining trap serves to spin up the wholecondensate.

1. Vortex behavior in two-component BEC mixtures (JILA)

The search for quantized vortices in dilute BECsstarted soon after the creation of trapped condensates.Since a mixture of two hyperfine components of 87Rbplays an essential role in the first reported vortex �Mat-thews et al., 1999b�, I first reviewed the physics of suchsystems. The GP energy functional in Eq. �2.21� isreadily generalized for a mixture of two BECs with con-densate wave functions �1 and �2. The total energycontains �1� the two separate energies Ej��j� of eachcomponent j=1,2, with appropriate self-interactionparameters gj �I assume repulsive interactions withgj�0� and �2� the interaction energy E12��1 ,�2�=�dV g12��1�2��2�2 between species 1 and 2, where g12 isthe interspecies interaction parameter �see Pethick andSmith �2002� and Pitaevskii and Stringari �2003��. Forthe simplest case of two uniform BECs in a box, if g12

2

�g1g2, the two components are immiscible and separateinto nonoverlapping phases. Otherwise, if g12

2 g1g2,

5Similar solutions play a crucial role in the study of rapidlyrotating condensates in Sec. V.

FIG. 2. Angular momentum l vs dimensionless angular veloc-ity � /�� for stable �black lines� and metastable �gray lines�states. Here $= �2/��1/2Na /dz is assumed small, so that theinteractions are simply a weak perturbation. There are nostable states with 0 l 1, and l=1 represents a single vortexon the axis of symmetry. For larger l, the angular momentumchanges smoothly because the positions of the two or morevortices vary continuously with � �for each stable branch, theinteger denotes the rotational symmetry�. Three-dimensionalplots of the density for two and seven vortices are remarkablysimilar to experimental images of Yarmchuk et al. �1979� forsuperfluid 4He �see Fig. 8� and of Madison et al. �2000a� forultracold dilute 87Rb gas �see Fig. 7�. From Butts and Rokhsar,1999.

660 Alexander L. Fetter: Rotating trapped Bose-Einstein condensates

Rev. Mod. Phys., Vol. 81, No. 2, April–June 2009

Page 15: Rotating trapped Bose-Einstein condensatespjh/teaching/phz7429/fetter.pdfII. Physics of Bose-Einstein Condensates in Dilute Trapped Gases 648 A. Ideal Bose gas 649 1. Ideal BEC in

they form two uniform interpenetrating BECs �in thecontext of dilute BECs, this result apparently was firstobtained by Nepomnyashchi �1974�; see also Colson andFetter �1978��.

For trapped dilute BECs, the situation is more com-plicated because the nonuniform trap potentials Vj in-duce both kinetic and potential energies that affect theoverall energy functional. The first such experiment�Myatt et al., 1997� at JILA �Boulder, CO� uses two dif-ferent hyperfine states of 87Rb. This isotope has nuclearspin I= 3

2 , which combines with the single valence elec-tron �S= 1

2 � to give two hyperfine manifolds, a lower onewith F=1 and an upper one with F=2. The typical traphas a local minimum in the magnetic field, and onlysome of these hyperfine states are weak-field seekingand hence stable �see Pethick and Smith �2002� and Pi-taevskii and Stringari �2003� for a description of mag-netic and optical traps�. In the common notation �F ,mF�,the stable weak-field seeking states are �1,−1�, �2, 1�, and�2, 2�.

The experiment of Myatt et al. �1997� cools state �1,−1� evaporatively and the interspecies interaction coolsstate �2, 2� through sympathetic cooling. Although theinteraction parameters approximately satisfy g12

2 "g1g2�Cornell et al., 1998�, implying that two such uniformBECs would overlap, the trapped condensates in factseparate because of the differences in the two trap po-tentials and interaction parameters �Myatt et al., 1997�.A careful theoretical analysis �Esry et al., 1997� that in-cludes all relevant effects finds a reasonable fit to theexperiments.

To avoid the problem of different trap potentials, sub-sequent experiments on these mixtures use the pair ofstates �1���1,−1� and �2���2,1� that differ by an angu-lar momentum �m=2. Each component can be imagedselectively with appropriately tuned lasers. Furthermore,the state can be quickly changed from �1� to �2� and backby a two-photon transition �microwave at �6.8 GHz andrf at �2 MHz�. To understand the physics of this system,it is helpful to examine the coupled time-dependent GPequations for the two components �Hall et al., 1998�:

i��

�t��1

�2 =�T + V1 + VH1 +

12��

12���t�ei�emt

12���t�e−i�emt T + V2 + VH2 −

12���

���1

�2 , �3.20�

where T=−�2�2 / �2M� is the kinetic energy, Vj is the trappotential, and VHj is the Hartree energy including theinteraction with both densities ��1�2 and ��2�2. The cru-cial new feature here is the applied electromagnetic �mi-crowave and rf� fields with combined frequency �em;they produce an off-diagonal coupling that involves thephase of each condensate wave function. The effectivecoupling ��t� is known as the “Rabi” frequency, whichincreases with the strength of the applied electromag-

netic fields �see Allen and Eberly �1987��. In general, thiscoupling depends explicitly on time �it vanishes whenthe fields are turned off�. Finally, � is the detuning be-tween the combined frequencies of the driving fields andthat of the atomic hyperfine transition �1�↔ �2�.

This electromagnetically coupled two-component sys-tem provides a remarkable example of topological con-trol of a quantum-mechanical state through an externalparameter. Specifically, when the near-resonant electro-magnetic coupling ��t� is turned off, the interaction be-tween the condensates involves only the densities�through VHj� with no phase information. In this case,each condensate has its own complex order parameter�j. The associated U�1� symmetry �see, for example,Schiff �1968�� has the topology of a circle or a cylinder.For T�Tc, the magnitude ��j� is fixed, and each com-plex order parameter has its own phase angle, yielding aquantized circulation that remains a topological invari-ant.

The topology is very different when the electromag-netic coupling is turned on because the coupled conden-sates in Eq. �3.20� now have a single two-componentSU�2� order parameter with the topology of a sphere,like that of a particle with spin 1

2 �see Schiff �1968� andHall et al. �1998��. Two spherical-polar angles suffice tocharacterize the coupled system, whose dynamics obeysthe optical Bloch equations �Allen and Eberly, 1987�,similar to those of nuclear magnetic resonance �NMR�.Specifically, the transformation of �1� into �2� is analo-gous to a � pulse in NMR that rotates a spin from up todown. If the pulse is twice as long, the resulting 2� ro-tation essentially reproduces the initial �1� state apartfrom a phase factor.

An experiment �Hall et al., 1998� showed the ability tomanipulate the coupled condensates as a single SU�2�order parameter. Application of a � /2 pulse to a purestate �1� produces a coherent mixture with equal popu-lation in both states that are spatially separated with asmall region of overlap. After a variable delay, a second� /2 pulse measures the relative quantum phase betweenthem. A subsequent experiment demonstrated thethree-dimensional character of the SU�2� topology ex-plicitly �Matthews et al., 1999a; Williams et al., 2000� asfollows: An external coupling field with a spatial gradi-ent produces a differential twist along the length of thecondensate. With increasing time, the increasing phasedifference between the two ends of the condensate re-laxes periodically by a three-dimensional evolution ofthe coupled SU�2� order parameter, leading to a collapseand recurrence of the structure. Such behavior wouldnot occur for two uncoupled U�1� condensate wavefunctions.

To create a quantized vortex, Williams and Holland�1999� proposed applying an external laser beam in ad-dition to the external electromagnetic fields that pro-duce the two-photon transition. This external laser beamwith frequency � �see Fig. 3� in effect adds a rotatinggradient to the coupling fields. Such a process startsfrom the ground state �1� and yields the state �2� with one

661Alexander L. Fetter: Rotating trapped Bose-Einstein condensates

Rev. Mod. Phys., Vol. 81, No. 2, April–June 2009

Page 16: Rotating trapped Bose-Einstein condensatespjh/teaching/phz7429/fetter.pdfII. Physics of Bose-Einstein Condensates in Dilute Trapped Gases 648 A. Ideal Bose gas 649 1. Ideal BEC in

unit of angular momentum. Specifically, the theory pre-dicts that the angular momentum of the state �2� oscil-lates and increases with time. Turning off the coupling atthe appropriate time leaves the state �2� with a persistentsingly quantized vortex because the associated order pa-rameter for this state �2� now has U�1� symmetry.

In the experiment �Matthews et al., 1999b�, the vortexin state �2� surrounds a nonrotating core consisting ofstate �1�. Furthermore, a � pulse can interchange �1� and�2�, so that the vortex can occur in either state. There isone important difference between the states �1� and �2�because the state �2� with mF=1 is not maximallyaligned. Consequently it can decay by spin relaxationwith a measured lifetime �1 s. As expected, a vortex instate �1� surrounding a core of state �2� is more stablebecause the state �2� core eventually decays, leaving asimple one-component vortex in state �1�. The experi-ment can image either the filled core or the surroundingvortex, but these images by themselves do not demon-strate the presence of quantized circulation. Fortunately,the interference procedure described above allows aclear measurement of the relative phase between thetwo condensates at various positions, demonstrating thesinusoidal variation predicted for a vortex �see Fig. 4�.

One important advantage of this two-component ex-periment is the ability to follow the motion of the vortexcore with nondestructive time-lapse images �Andersonet al., 2000�. Figure 5 shows the uniform precession ofthe filled vortex core in state �2� around the trap center,determining the angular velocity of the circular motion.Note the gradual shrinkage of the core radius as state �2�decays through spin-flip transitions �Fig. 5�d��. The ex-perimenters can also remove the core component with abrief resonant laser pulse, leaving a one-component vor-tex whose empty core �of order �� is too small to imagein the trap. To study the precession of such a one-component vortex, the experimenters first image thetwo-component vortex, then remove the core, wait avariable time, and finally turn off the trap, imaging the

expanded vortex core. In this way, they measure the pre-cession rate of a one-component empty-core vortex,yielding values in reasonable agreement with theoreticalpredictions �Fetter and Svidzinsky, 2001�. These datashow no evidence that the vortex spirals outward, sug-gesting that thermal damping is negligible on the timescale of �1 s.

2. Vortex behavior in one-component BECs [Ecole NormaleSupérieure (ENS) and subsequently others]

Most other experimental methods of creating vorticesin dilute BECs rely on the rotation of an anisotropicpotential, and I briefly review the physics of such a sys-tem. Consider a classical incompressible fluid in a longcylinder with elliptical cross section �x��a and �y��b.When the container rotates with angular velocity �=�z, the moving walls �in the laboratory frame� exertforces on the fluid. As a result, the fluid executes apurely irrotational motion with velocity virr=��irr �see

FIG. 3. Vortex creation in a trapped two-component BEC. �a�Schematic of the technique used to create a vortex: An off-resonant laser provides a rotating force on the atoms acrossthe condensate with a simultaneous microwave drive of detun-ing �. �b� Level diagram showing the microwave transition tothe sublevels of the state �2�. When the rotation frequency ofthe laser beam � is close to the detuning �, the perturbationselectively modulates the transition from the l=0 state to thel=1 state. From Matthews et al., 1999b.

FIG. 4. Cosine of the relative phase around the vortex, show-ing the sinusoidal variation of the azimuthal angle. From Mat-thews et al., 1999b.

FIG. 5. Vortex precession in a trapped two-component BEC.�a� Successive images of a two-component condensate with avortex in state �1� surrounding a core in state �2�. �b� The imageis then fit with a smooth Thomas-Fermi distribution. �c� Theazimuthal angle of the vortex core plotted against the time,giving a precession frequency of 1.3�1� Hz. �d� The gradualdecrease in the core radius r �measured in units of the coher-ence length �� because of spin relaxation. From Anderson etal., 2000.

662 Alexander L. Fetter: Rotating trapped Bose-Einstein condensates

Rev. Mod. Phys., Vol. 81, No. 2, April–June 2009

Page 17: Rotating trapped Bose-Einstein condensatespjh/teaching/phz7429/fetter.pdfII. Physics of Bose-Einstein Condensates in Dilute Trapped Gases 648 A. Ideal Bose gas 649 1. Ideal BEC in

Eq. �2.38��, where the associated velocity potential �irr

=�xy�a2−b2� / �a2+b2� is proportional to xy �see Lamb�1945��. Remarkably, this irrotational velocity field in-duces a nonzero angular momentum per unit lengthLirr=Lsb�a2−b2�2 / �a2+b2�2, where Lsb is the correspond-ing angular momentum per unit length for a fluid thatexecutes solid-body rotation with vsb=��r. Note thatthis irrotational angular momentum occurs withoutquantized vortices. Such vortex-free angular momentumincreases the critical angular velocity �c for vortex for-mation in superfluid 4He �Fetter, 1974�, as DeConde andPackard �1975� verified by experiments on rotating ellip-tical and rectangular cylinders.

The same irrotational flow occurs in rotating aniso-tropic trapped BECs, where it appears in the phase ofthe condensate wave function S=M�irr /� �see Eq.�2.38�� and leads to unexpected phenomena. For ex-ample, a sudden small rotation of the anisotropic trappotential induces an unusual oscillatory motion knownas the “scissors mode” �see Pitaevskii and Stringari�2003� for a detailed description of this and related ex-periments�.

The experimental group at the ENS in Paris used sucha direct approach �see Fig. 6� to create one or morevortices in a one-component 87Rb cigar-shaped conden-sate �Madison et al., 2000a�. Starting with an axisymmet-ric condensate, they applied a toggled off-center stirringlaser beam that slightly deforms the trap potential; theplane of deformation then rotates about the long axiswith an applied angular velocity � / �2��"200 Hz �some-what earlier, Onofrio et al. �2000� used a similar tech-nique to excite stationary and rotating surface modeswith twofold and fourfold symmetry�. Physically, the la-ser beam acts on each polarizable atom in the BEC witha “dipole” force6 proportional to the gradient of thesquared electric field of the laser �for the details of thestirring procedure, see Madison et al. �2000b��. Thisdipole-force mechanism is the basis for optical trappingof atoms, as discussed in Sec. II.B in connection withoptical lattices �for a more detailed account, see Pethickand Smith �2002� and Pitaevskii and Stringari �2003��.

Since the vortex core radius ���0.2–0.4 m is toosmall for direct visual detection, they turn off the trapand allow the condensate �including the vortex core� toexpand. The initial tight radial confinement means thatthis expansion is mostly in the radial direction, and theexpanded condensate usually becomes disk shaped. Theresulting images of the expanded condensate show thepresence of one or a few vortices �see Fig. 7�. Theselatter images are remarkably similar to those in Fig. 8 ofvortices in rotating superfluid 4He �Yarmchuk et al.,

6Here, “dipole” means the induced atomic electric dipole; theactual deformation of the originally axisymmetric condensateusually has quadrupole symmetry, but Onofrio et al. �2000� andRaman et al. �2001� also induced deformations with highersymmetry.

FIG. 6. A cigar-shaped condensate is confined by an axisym-metric magnetic trap and stirred by an off-center far-detunedlaser beam. The laser beam propagates along the long axis andcreates an anisotropic quadrupole potential that rotates at anangular velocity �. From Madison et al., 2000b.

FIG. 7. Absorption images of a BEC stirred with a laser beam;in all images the condensate has about 105 atoms, the tempera-ture is below 80 nK, and the rotation rate � / �2�� increasesfrom 145 Hz for �a� and �c� to 168 Hz for �g�. Images �a� and�b� show the optical density for images �c� and �d�, with theclear appearance of the vortex core. Images �e�–�g� show stateswith two, three, and four vortices. From Madison et al., 2000a.

FIG. 8. Photographs of stable vortex configurations in super-fluid 4He, with up to four vortices. Adapted from Yarmchuket al., 1979.

663Alexander L. Fetter: Rotating trapped Bose-Einstein condensates

Rev. Mod. Phys., Vol. 81, No. 2, April–June 2009

Page 18: Rotating trapped Bose-Einstein condensatespjh/teaching/phz7429/fetter.pdfII. Physics of Bose-Einstein Condensates in Dilute Trapped Gases 648 A. Ideal Bose gas 649 1. Ideal BEC in

1979� obtained by trapping charged particles on the vor-tex cores �see Donnelly �1991��. Hodby et al. �2002� �Ox-ford� observed similar vortex patterns with a rotatingmagnetic deformation of the trap potential. Their ex-perimental arrangement allows them to explore thephysics of vortex nucleation for relatively large quadru-pole deformations.

How the vortices nucleate is not a simple question. Insuperfluid 4He, the general picture is that a finite rela-tive velocity between the fluid and the rotating wall cannucleate a vortex at local regions of high surface rough-ness �see, for example, Packard and Sanders �1972��. Thevortex then migrates toward the bulk, although this pro-cess is not well understood. Packard and Sanders �1972�and Yarmchuk and Packard �1982� studied the appear-ance �disappearance� of vortices as the angular velocityincreases �decreases�. They found moderate hysteresis in�, as expected from this “cool, then rotate” scenario.

The situation in a trapped BEC should be quite dif-ferent, for the trap itself has a smooth surface. Dalfovoand Stringari �2000� developed a detailed picture basedon surface modes of the condensate, where the TF hy-drodynamic theory �Stringari, 1996� predicts that a sur-face mode with angular dependence e±im! has an angularfrequency �m=��m���. The corresponding criticalLandau angular frequency is �c=min��m / �m��=min��� /��m��, which clearly depends on symmetry mof the surface mode.

Assume first that the stirring laser beam has quadru-pole symmetry with �m�=2 and a width comparable tothe condensate diameter. In this case �Chevy et al., 2000;Madison et al., 2000a, 2000b, 2001�, the measured criticalangular velocity �c for the appearance of the first vortexis �0.7��, close to prediction �� /�2 based on quadru-pole surface excitations. Note, however, that this value istypically much larger than the thermodynamic value�Eq. �3.10�� for the relevant TF limit. Sinha and Castin�2001� and Pitaevskii and Stringari �2003� discussed thisquadrupole mode in detail.

In contrast, if the width of the stirring laser beam issmall compared to the condensate diameter, then theperturbation can create collective surface waves withhigher values of �m� and correspondingly smaller Landaucritical angular velocities �c��� /��m�. Indeed, in thissituation, the observed critical frequency for vortex cre-ation can be as low as the theoretical thermodynamiccritical frequency �Raman et al., 2001; Chevy et al.,2002�. Haljan, Coddington, et al. �2001� found a compa-rably lower critical rotation frequency when they rotatea normal cloud and then cool into the superfluid state.This latter observation is consistent with the expectationthat a “rotate, then cool” scenario would attain thermo-dynamic equilibrium at the particular �, although I amunaware of comparable experiments that rotate 4He inthe normal state and then cool into the superfluid state.

3. Normal modes of an axisymmetric condensate with onecentral vortex

A static condensate without vortices has numerouslow-lying small-amplitude normal modes characterizedby coupled density and velocity perturbations. Many ex-perimental studies have confirmed the various theoreti-cal predictions in considerable detail �see Pethick andSmith �2002� and Pitaevskii and Stringari �2003��. A di-rect treatment of the hydrodynamic equations �Stringari,1996� for a nonrotating axisymmetric condensate in theTF limit yields the frequency of quadrupole modes withl=2, m=0 and l=2, m= ±2. The last two modes with m= ±2 are degenerate and together produce an oscillatingquadrupole distortion of the condensate in the planeperpendicular to the original axis of symmetry. In theabsence of any perturbation that breaks time-reversalsymmetry, the axes of the deformation remain fixed inspace �see Fig. 9�a��.

When the condensate has a vortex, however, its circu-lating velocity breaks time-reversal symmetry and splitsthe previously degenerate states with ±m �Svidzinskyand Fetter, 1998; Zambelli and Stringari, 1998�. This ef-fect is analogous to the Zeeman splitting of magneticsublevels. In the usual TF limit, the fractional splitting issmall �of order �m�d�

2 /R�2 �. One dramatic consequence is

that the quadrupole distortion now precesses slowly inthe direction of the circulating vortex flow �Chevy et al.,2000; Haljan, Anderson, et al., 2001�, as seen in Figs. 9�b�and 9�c�.

Zambelli and Stringari �1998� noted that the magni-tude of the splitting provides a direct measure of theangular momentum of the condensate. The ENS group�Chevy et al., 2000� used this approach to detect the firstappearance of a vortex in the condensate and to verifythat its angular momentum per particle is ��. It is re-

FIG. 9. Nondestructive images of quadrupole mode of an axi-symmetric condensate. Condensate �a� is vortex-free, whereascondensates in �b� and �c� have straight vortices �with oppositesense of circulation� aligned perpendicular to the plane of thefigure; �d� shows a linear fit to the precession angle. From Hal-jan, Anderson, et al., 2001.

664 Alexander L. Fetter: Rotating trapped Bose-Einstein condensates

Rev. Mod. Phys., Vol. 81, No. 2, April–June 2009

Page 19: Rotating trapped Bose-Einstein condensatespjh/teaching/phz7429/fetter.pdfII. Physics of Bose-Einstein Condensates in Dilute Trapped Gases 648 A. Ideal Bose gas 649 1. Ideal BEC in

markable that this macroscopic experiment with N�106 atoms measures a single-particle fundamental con-stant; it demonstrates the coherence of the macroscopicwave function �see Leggett �2006� for a valuable per-spective on this and related experiments�. Donnelly�1991� discussed similar earlier macroscopic experimentson rotating superfluid 4He.

The precession of the quadrupole mode provides asensitive vortex detector �sometimes called “surface-wave spectroscopy”� that complements the direct visibil-ity of the vortex core. The JILA group �Haljan, Ander-son, et al., 2001� used this method to track the large-amplitude tilting motion of a vortex in an axisymmetriccondensate. As a tilted vortex rotates about the axis ofsymmetry, its empty core initially moves away from theline of sight. Periodically, however, the vortex returns toits original orientation, when the optical visibility of thecore and the quadrupole precession both attain their ini-tial values. The experiment indeed observes two recur-rences, yielding good agreement with the predicted pre-cession frequency �Svidzinsky and Fetter, 2000b�.

4. Bent vortex configurations

When a disk-shaped condensate rotates faster thanthe critical velocity �m for creation of a metastable vor-tex �see Eq. �3.11��, the initial equilibrium vortex is es-sentially straight. For condensates with larger aspect ra-tios Rz /R�, however, the equilibrium vortex shape canbe curved, depending on the angular velocity and thespecific aspect ratio. Predictions of this phenomenonarise both from linearized stability analysis for a straightvortex �Svidzinsky and Fetter, 2000b; Feder et al., 2001�and from direct calculation of the state that minimizesthe energy in the rotating frame �Aftalion and Rivière,2001; García-Ripoll and Pérez-García, 2001a, 2001b�. Ina long cigar-shaped condensate, for example, the initialvortex can reduce its total energy by forming a smallsymmetrical loop near the equator. As the angular ve-locity � increases, the vortex loop expands toward theaxis of symmetry. When these vortex loops becomelarge, ENS experiments �Rosenbusch et al., 2002�showed that they can have both symmetric configura-tions �denoted U� and antisymmetric configurations �de-noted �unfolded� N although S or � might be more ap-propriate�. Modugno et al. �2003� gave a physical pictureof how curvature can lower the total energy of the vor-tex.

5. Kelvin (axial) waves on a vortex

A long straight classical vortex line in an incompress-ible fluid represents a dynamical equilibrium betweenthe kinetic-energy density and the pressure, as followsfrom Bernoulli’s equation. In an influential paper,Thomson �1880� �subsequently, he became Lord Kelvin�studied the small oscillations of such a vortex; he founda characteristic long-wavelength ��k���1� dispersion re-lation

�k ��k2

2Mln� 1

�k�� , �3.21�

where k is the axial wave number, � is the vortex coreradius, and I assume a positive quantized circulation �=2�� /M. For general k, these classical Kelvin modesare left circularly polarized helical waves with velocitypotential �exp�i�kz−!−�kt��K1�kr�, where �r ,! ,z� arecylindrical polar coordinates and K1�u� is a Bessel func-tion of imaginary argument that vanishes for large u.Each element of the vortex core executes a circular orbitwith a sense opposite to the circulating velocity of thevortex line. In a quantum-mechanical version, thesemodes have quantized angular momentum with m=−1.Note that this negative helicity �m=−1� is independentof the direction of the wave propagation �±k�. The over-all helical behavior reflects the gyroscopic character ofthe vortex line, and the physics thus differs significantlyfrom that of a stretched string.7

Unfortunately, Kelvin’s treatment is rather intricate,for he allowed the unperturbed axisymmetric flow tohave an arbitrary radial dependence �including bothsolid-body rotation and irrotational vortex flow�. For ac-cessible discussions of these Kelvin modes, see Fetter�1969�, Lifshitz and Pitaevskii �1980b�, Sonin �1987�,Donnelly �1991�, and Fetter and Walecka �2006�.

In a modern quantum-mechanical context, Pitaevskii�1961� used the Bogoliubov equations �2.43� and �2.44�to show that a singly quantized vortex in a dilute BECalso has this Kelvin mode with the same long-wavelength dispersion relation �Eq. �3.21��. Althoughthe existence of Kelvin waves are widely accepted, directevidence for their presence in low-temperature superflu-ids is relatively meager. Sonin �1987� and Donnelly�1991� discussed the interpretation of relevant superfluid4He experiments in some detail.

The evidence is much more direct in dilute BECs.Bretin et al. �2003� reported a clear signature of suchquantized Kelvin modes. They drive the condensatewith a near-resonant quadrupole excitation �m= ±2� andobserve preferential damping of the mode with m=−2.Recall that the Kelvin modes have quantized angularmomentum m=−1 relative to the circulation around thevortex, independent of the direction of k. They sug-gested that this particular excited state with m=−2 candecay by the creation of two Kelvin-wave excitations,each with m=−1 and opposite momenta ±�k along theaxis determined by the conservation of energy. In con-trast, the other quadrupole mode m= +2 has no suchopen channel for decay. A theoretical analysis �Mi-zushima et al., 2003� confirmed this interpretation. Fur-thermore, transverse images of the expanded conden-sate excited with the quadrupole mode m=−2 exhibitthe expected periodic spatial structure associated withthe Kelvin-mode deformation, whereas those excited

7For a detailed comparison, see Fetter and Harvey �1971�.

665Alexander L. Fetter: Rotating trapped Bose-Einstein condensates

Rev. Mod. Phys., Vol. 81, No. 2, April–June 2009

Page 20: Rotating trapped Bose-Einstein condensatespjh/teaching/phz7429/fetter.pdfII. Physics of Bose-Einstein Condensates in Dilute Trapped Gases 648 A. Ideal Bose gas 649 1. Ideal BEC in

with m= +2 have no periodic structure �Bretin et al.,2003�.

A particularly interesting situation arises when a ro-tating condensate is subsequently confined in a one-dimensional optical lattice �see Sec. II.B� oriented alongthe rotation axis. Consider an axisymmetric condensatewith a single vortex on the axis of symmetry. Then adia-batically turn on a standing-wave laser field. If the laserbeams propagate along the symmetry axis, the opticallattice slices the rotating condensate into many circulardisks. The resulting periodic atomic array forms a set ofeffectively two-dimensional condensates that arecoupled by tunneling between adjacent layers. Eachtwo-dimensional condensate contains a segment of thepreviously continuous vortex line. Thus it becomes asingle pancake vortex, and the position of each vortexcenter can fluctuate more-or-less independently, depend-ing on the strength of the laser field and the height ofresulting potential barriers. Martikainen and Stoof�2003� used a variational trial function to study the re-sulting axial modes in this discrete system, which are theanalog of the Kelvin modes in a cigar-shaped conden-sate. At present, no direct experimental evidence forsuch behavior exists.

6. Other ways to create vortices

As seen throughout Sec. III.D, experimentalists havedevised many procedures that successfully create one ormore vortices in a dilute BEC. Section III.D.1 discussesdynamical phase-imprinting in a two-component con-densate, and Sec. III.D.2 considers the more commonapproach of rotating an anisotropic potential that spinsup the condensate. A related technique �discussed inSec. IV.B in connection with vortex arrays� is to spin upthe normal gas and then cool into the superfluid state�Haljan, Coddington, et al., 2001�. One-dimensional soli-tons provide an additional unusual alternative, for theygenerally decay to form three-dimensional vortex rings�Anderson et al., 2001; Dutton et al., 2001�.

a. Transfer of orbital angular momentum from a laser

A Gaussian laser beam can be decomposed into com-ponents with orbital angular momenta m relative to thepropagation vector �the Laguerre-Gaussian modes�.Thus a laser beam can, in principle, create a vortex bycoherently transferring orbital angular momentum tothe condensate. In a recent experiment, Andersen et al.�2006� used a two-photon stimulated Raman processinvolving a counter-propagating linearly polarizedLaguerre-Gaussian mode with m=1 �Allen et al., 1999,2003� and a linearly polarized Gaussian mode with m=0 �the linear polarization ensures that “spin” angularmomentum does not cause the observed effect�. Some ofthe atoms in an initial single 23Na condensate absorb aphoton with one unit of angular momentum from theLaguerre-Gaussian beam and emit a photon to theGaussian beam, acquiring both angular momentum andlinear momentum. Consequently, these atoms form anew coherent moving condensate with a singly quan-

tized vortex. After release from the trap, the resultingspatial separation between the new and original conden-sates permits clear imaging of the vortex core.

Andersen et al. �2006� also used a sequence of suchpulses to make vortices with higher quantized circula-tion; as noted below, however, such vortices are essen-tially unstable with respect to decay into singly quan-tized vortices. Recently, similar techniques createpersistent quantized flow with one unit of circulation ina toroidal trap �Ryu et al., 2007�; the �10 s lifetime ofthis persistent flow arises from experimental factors andis not intrinsic.

b. Merging of multiple BECs

Another experiment creates a vortex by mergingthree 87Rb condensates in a single disk-shaped trap.Scherer et al. �2007� applied a repulsive �blue-detuned�laser beam shaped as a threefold optical barrier, dividingthe noncondensed gas into three nonoverlapping regions�see Sec. II.B for the effect of external laser beams�. Asthe gas cools below Tc, three independent and uncorre-lated BECs form. The optical barrier is then slowly re-moved, and the three BECs merge. After a variable de-lay time, they remove the trap and image the expandedcondensate. Depending on the initial relative phases ofthe three condensates, a net circulating current canform, producing a visible vortex core in the expandedimage. For random phase differences, a vortex shouldappear with probability 0.25, in reasonable agreementwith the experimental observations for repeated trials.

c. Imposition of topological phase

As a final topic in this discussion of forming a singlevortex, I consider the creation of a doubly quantizedvortex through imposition of topological phase �Lean-hardt et al., 2002� �they also create fourfold vortices�.The experiment follows suggestions by Isoshima et al.�2000�, Nakahara et al. �2000�, and Ogawa et al. �2002�. Itrelies on a special form of magnetic confinement knownas the Ioffe-Pritchard trap �for a detailed description,see Pethick and Smith �2002��. Here the important fea-ture is that the x and y components of the magnetic fieldhave a quadrupole form, whereas the z component canbe taken as uniform.

This MIT experiment studies 23Na in the hyperfinestates �F=1,mF=−1� and �F=2,mF=2�. Initially, the zcomponent of magnetic field is large and positive, whichaligns the hyperfine spin F vector with the direction ofB�Bzz. Thus this spin texture is uniform and oriented,with variable number density induced by the nonuni-form trap potential. The experiment then slowly re-verses the z component of magnetic field, and the hyper-fine spin vector F adiabatically follows the localmagnetic field. In particular, when Bz vanishes, the hy-perfine spin vector F takes the �nonuniform� quadrupoleform of the remaining x and y components of B. As Bzthen becomes negative, the topological winding of thespin texture continues adiabatically, leading to a non-trivial final texture. Remarkably, at the end of the Bz

666 Alexander L. Fetter: Rotating trapped Bose-Einstein condensates

Rev. Mod. Phys., Vol. 81, No. 2, April–June 2009

Page 21: Rotating trapped Bose-Einstein condensatespjh/teaching/phz7429/fetter.pdfII. Physics of Bose-Einstein Condensates in Dilute Trapped Gases 648 A. Ideal Bose gas 649 1. Ideal BEC in

inversion, a detailed analysis shows that the local spintexture has an additional factor e−i2mF!, where mF=−1 or2. Consequently, the final state represents a doublyquantized vortex with quantum number two or a four-fold vortex with quantum number four, in each case hav-ing a reversed sense of rotation. Leanhardt et al. �2002�used surface-wave spectroscopy �precession of the quad-rupole oscillation mode, see Sec. III.D.3� to verify thatthese vortices indeed have angular momentum �2� or�4� per particle; their expansion images also illustratedthat a second adiabatic reversal of the magnetic fieldreturns the system to its original state.

The energy of a vortex arises principally from the to-tal kinetic energy �dV 1

2Mv2���2, which is proportional tothe square of the circulation. In a multiply connectedgeometry like an annulus, such flow represents a macro-scopic quantized circulation, and the presence of the in-ner wall ensures that the hydrodynamic flow generallyremains stable �a persistent current� as long as the tem-perature is well below Tc. The situation is very differentfor a free-standing multiply quantized vortex, whichshould be unstable with respect to splitting into the ap-propriate number of singly quantized vortices. As men-tioned at the end of Sec. II.D, Pu et al. �1999� studied theBogoliubov equations for a doubly quantized vortex andfound imaginary eigenvalues that imply a dynamical in-stability �see also Castin and Dum �1999��. Shin et al.�2004� investigated the stability of these doubly quan-tized vortices, with convincing images of the relativelylarge initial vortex core decaying into two smaller coresin �75 ms. Möttönen et al. �2003� provided a theoreticalanalysis of the MIT experiments, emphasizing the un-stable eigenvalues of the Bogoliubov equations. Numeri-cal simulations �Huhtamäki et al., 2006� with realistic pa-rameters gave good agreement with the measured timefor decay into two singly quantized vortices.

An alternative picture of the same topological trans-formation �Leanhardt et al., 2002; Ogawa et al., 2002�relies on the idea of the Berry �1984� phase that arisesfrom an adiabatic alteration of the physical system �inthis case, the overall direction of the magnetic field�. Inthe context of quantum mechanics, many people haveprovided a description of the Berry phase; see, for ex-ample, Shankar �1994�, Abers �2004�, and Griffiths�2005�.

E. Berezinskii-Kosterlitz-Thouless transition in dilute BECs

Another remarkable application of an optical lattice isthe study of the equilibrium state of a two-dimensionaldilute Bose-Einstein gas. Although a uniform two-dimensional system cannot have a true condensate withlong-range order �Mermin and Wagner, 1966; Hohen-berg, 1967�, Berezinskii �1971� and Kosterlitz and Thou-less �1973� predicted that it can indeed become super-fluid below a certain critical temperature TBKT. Abovethis temperature, free single vortices can form easily.The motion of such vortices disrupts any organized long-range phase coherence �Leggett, 2001� and hence de-

stroys the associated superfluid flow. Below TBKT, incontrast, free vortices do not occur, although boundpairs of a vortex and an antivortex do exist �in thepresent context of dilute ultracold gases, Pitaevskii andStringari �2003� discussed this behavior in more detail�.As the temperature increases, so does the size of thebound pairs, and they eventually become unbound atTBKT, producing free vortices. Experiments with thin su-perfluid 4He films �Bishop and Reppy, 1978� and withJosephson-coupled superconducting arrays �Resnick etal., 1981� have confirmed these predictions in consider-able detail.

In a recent experiment at ENS �Paris�, Hadzibabic etal. �2006� �see also Hadzibabic et al. �2008�� applied astanding-wave laser field perpendicular to the symmetryaxis of a cigar-shaped condensate, splitting the atomsinto two essentially independent thin pancake strips withlateral dimensions �120 m�10 m; see Fig. 6 for thegeometry �although there is no rotating laser beam inthe present case�: the laser beams propagate in the xdirection and the pancakes are parallel to the y-z plane.After these two isolated Bose-Einstein gases reach ther-mal equilibrium, the confining traps are turned off. Bothpancakes then expand primarily in the x direction per-pendicular to their planes and overlap spatially.

Soon after the first observations of BECs in dilutetrapped gases, the MIT group �Andrews et al., 1997�used a repulsive blue-detuned laser at the middle of acigar-shaped trap to create two separate long three-dimensional condensates. When the blue-detuned laserand the trap are turned off, these two expanding con-densates overlap and produce clear straight interferencefringes. Similarly, the ENS experiment looks for andfinds interference fringes between the two pancakes. Atlow temperature �see Fig. 10�a��, the fringes are indeedstraight, indicating constant relative phase. As the tem-perature increases, however, the fringes become wavy�Fig. 10�b��; furthermore, as seen in Fig. 10�c�, theysometimes contain one or more dislocations, which indi-cates the presence of free vortices in one or both pan-cakes �Stock, Hadzibabic, et al., 2005; Simula and Blakie,2006�. They repeated the experiment at various tem-peratures and used the presence of free vortices to de-termine TBKT. The resulting value agrees well with thetheoretical predictions of Berezinskii �1971� and Koster-litz and Thouless �1973�, although the finite size of the

FIG. 10. Interference fringes between two pancake conden-sates �a� at low temperature and �b� at high temperature show-ing wavy interference fringes without vortices. �c� Fringe dis-location indicates the presence of a vortex in one or bothcondensates. From Krüger, 2007.

667Alexander L. Fetter: Rotating trapped Bose-Einstein condensates

Rev. Mod. Phys., Vol. 81, No. 2, April–June 2009

Page 22: Rotating trapped Bose-Einstein condensatespjh/teaching/phz7429/fetter.pdfII. Physics of Bose-Einstein Condensates in Dilute Trapped Gases 648 A. Ideal Bose gas 649 1. Ideal BEC in

present samples yields a crossover instead of a sharptransition.

Schweikhard et al. �2007� reported a study of vortexformation in a two-dimensional array of Josephson-coupled BECs. They started at T below Tc with a disk-shaped condensate along with a residual normal cloudand then applied a hexagonal optical lattice, transform-ing the system into an array of coupled BECs that actlike Josephson junctions. Each condensate has a singleeffective phase, and the coupling energy between adja-cent condensates has the form J�1−cos �!�, where �! istheir relative phase. Equivalently, the tunneling currentbetween any pair of condensates is proportional toJ sin �! �for a discussion, see Josephson �1962�,Tinkham �1996�, and Leggett �2001��. Varying the ampli-tude of the optical lattice through the laser intensitytunes the coupling parameter J.

The experiment starts by inferring the temperature Tfrom a nondestructive image and then ramps down theoptical lattice, allowing the adjacent BECs to merge. AsScherer et al. �2007� discussed, vortices can form de-pending on the relative phases, appearing as holes in theexpanded gas cloud after the trap is turned off. For agiven temperature T, Schweikhard et al. �2007� mea-sured the vortex density D near the center of the cloudas a function of the Josephson coupling parameter J. Fortwo different temperatures, a plot D as a function of J /Tshows nearly universal behavior: D is negligibly small forJ /T1 �low temperature�, but it rises rapidly and satu-rates for J /T�1 �high temperature�, as expected for theBKT vortex unbinding crossover.

IV. VORTEX ARRAYS IN MEAN-FIELD THOMAS-FERMI(TF) REGIME

A rapidly rotating condensate has a dense array ofvortices, with a uniform areal density �Feynman, 1955�

nv =M�

��, �4.1�

as follows from these arguments.�1� Consider a steady classical incompressible flow

with velocity v�r� in a rigid container that rotates withangular velocity �. If the fluid has number density n�r�,then the relevant part of the energy in the rotatingframe is

E� =� dV�12

Mv2 − M� · r � v n . �4.2�

This result is equivalent to

E� =12

M� dV�v − � � r�2n − Esb, �4.3�

where Esb= 12M�dV���r�2n= 1

2Isb�2 is the rotational en-

ergy for solid-body rotation with the corresponding mo-ment of inertia Isb. The first term of Eq. �4.3� is non-negative, and hence the absolute minimum of E� occurswhen v=��r, namely, solid-body rotation vsb=��r;

in addition, the corresponding minimum value of Emin�has the classical value −Esb. Since ��vsb=2�, the irro-tational superflow cannot assume such a solid-body ro-tation, but it seeks to mimic this flow through the nucle-ation of quantized vortices.

The thermodynamic relation

Lz = − ��E�/���N �4.4�

gives the angular momentum Lz= �r�p�, which also fol-lows from the Hellmann-Feynman theorem �see, for ex-ample, Cohen-Tannoudji et al. �1977�; for a history ofthis theorem, see Brown �2000��. Since the actual E� ex-ceeds the classical value − 1

2Isb�2 by a positive correction

that depends on �, the equilibrium value of the angularmomentum necessarily differs from the classical valueIsb�. Evidently, this system exhibits a nonclassical mo-ment of inertia �see Leggett �2006� for a discussion ofthese quantum-mechanical effects�. In addition, the tri-angular vortex array also has transverse shear modes�the Tkachenko modes, see Sec. IV.B.4�, and it “melts”for sufficiently large angular velocities to form a “liquid”state �see Sec. VI�. Whether this system should be con-sidered a “supersolid” in the conventional sense �see, forexample, Leggett �2006�� merits additional study.

�2� For a superfluid, Eq. �3.3� shows that a singly quan-tized vortex at a position r0 has a localized vorticity ��v= �2�� /M���2��r−r0�z. In a rotating superfluid, thevortices will be uniformly distributed to approximate theuniform classical vorticity. Take a contour C containingNv such vortices in an area A; the total circulationaround C is %C= �2�� /M�Nv. Stokes’s theorem showsthat the corresponding classical value would be %C

cl

=2�A, and comparison immediately yields the mean ar-eal vortex density nv=Nv /A=M� /��. The inverse nv

−1 isthe area per vortex �� /M���l2, which defines the ra-dius

l =� �

M��4.5�

of an equivalent circular cell. Thus the intervortex sepa-ration is �2l. Note that l decreases with increasing � �itis also the usual oscillator length for a frequency ��.

A. Physics of BEC in axisymmetric harmonic traps for rapidrotation

The Feynman relation �4.1� shows that the number ofvortices increases linearly with � assuming that the ge-ometry remains fixed. Although this picture indeed ap-plies to superfluid 4He in a rotating bucket, the situationis quite different for a dilute gas in a harmonic trap,because the centrifugal forces expand the condensate ra-dially �number conservation means that the condensatealso shrinks along the axis of rotation�. As a result, thetotal number of vortices in a given condensate

Nv��� � nv�R�2 ��� = M�R�

2 ���/� = R�2 ���/l2 �4.6�

increases faster than linearly with �.

668 Alexander L. Fetter: Rotating trapped Bose-Einstein condensates

Rev. Mod. Phys., Vol. 81, No. 2, April–June 2009

Page 23: Rotating trapped Bose-Einstein condensatespjh/teaching/phz7429/fetter.pdfII. Physics of Bose-Einstein Condensates in Dilute Trapped Gases 648 A. Ideal Bose gas 649 1. Ideal BEC in

To quantify the dependence of R� on �, it is conve-nient to focus on the TF limit of a large condensate,when the spatial variation of the condensate density isnegligible compared to the other terms in the energyfunctional. Thus p�=−i���= ���S��− i�eiS� ����Mv�, and the GP energy functional in the rotatingframe Eq. �3.6� simplifies to

E���� � � dV��12

Mv2 − M� · r � v + Vtr ���2

+12

g���4� . �4.7�

Like the TF limit ETF��� in Eq. �2.26� for a stationarycondensate, this equation depends only on ���2 and ���4,but it now includes the hydrodynamic flow through v.

In the typical case of rotation about z, simple manipu-lations yield the equivalent expression

E���� � � dV12

M�v − vsb�2���2

+� dV��Vtr −12�2r2 ���2 +

12

g���4��Ev� + ETF� , �4.8�

where Ev� and ETF� denote the contributions in the firstand second lines. For a large condensate with many vor-tices, the actual superfluid velocity v closely approxi-mates the classical solid-body value vsb, so that Ev� �theterm containing �v−vsb�2� vanishes to leading order. Theremaining terms ETF� ��� have exactly the same form asETF in Eq. �2.26�, but the radial part of the confiningpotential is altered to 1

2M���2 −�2�r2. Variation with re-

spect to ���2 gives an equation of the same form as Eq.�2.27�, but with the reduced radial confining potential.Thus the TF density again has the form of Eq. �2.29�, butwith modified TF condensate radii that depend on �.The normalization condition in three dimensions showsthat

TF��� = TF�0��1 − �2/��2 �2/5, �4.9�

so that the chemical potential decreases continuouslyand formally vanishes for �→��, as does the centraldensity TF��� /g �because of the reduced radial confine-ment�. A combination of these results leads to the spe-cific � dependence of the condensate dimensions �Buttsand Rokhsar, 1999; Fetter, 2001; Raman et al., 2001�

R����R��0�

= �1 −�2

��2 −3/10

,Rz���Rz�0�

= �1 −�2

��2 1/5

.

�4.10�

As anticipated, the external rotation expands the con-densate radially and shrinks it axially, approaching atwo-dimensional configuration. The corresponding as-pect ratio

Rz���R����

=���

2 − �2

�z�4.11�

provides a convenient diagnostic tool to infer theactual angular velocity of the rotating condensate�Haljan, Coddington, et al., 2001; Raman et al., 2001;Schweikhard et al., 2004�, as seen in Fig. 11.

The two terms in Eq. �4.8� have quite different mag-nitudes in the mean-field TF regime. The dominant con-tribution ETF� ��� is the TF energy of the rotating con-densate with the modified squared radial trap frequency��

2 →��2 −�2. This quantity follows directly from Eqs.

�2.32� and �4.9�,

ETF� ��� = 57TF���N � �1 − �2/��

2 �2/5. �4.12�

The thermodynamic relation in Eq. �4.4� yields the cor-responding TF angular momentum of the rotating con-densate, and some algebra gives the simple expression

LTF = 27MNR�

2 ���� = Isb���� , �4.13�

where Isb���= 27MNR�

2 ��� is the solid-body moment ofinertia for the deformed rotating condensate �this ex-pected result follows from a separate calculation of�r2�TF�.

The remaining term Ev� in Eq. �4.8� gives a small non-classical contribution to the angular momentum of therotating condensate. If the volume integral is approxi-mated as a sum of integrals over circular cylindrical cellsof radius l centered at each vortex, a straightforwardanalysis yields

Ev���� � N�� ln�l/�� . �4.14�

With logarithmic accuracy �ignoring the � dependenceinside the logarithm�, the corresponding nonclassical an-gular momentum then becomes

Lv� = −�Ev�

��� − N� ln� l

� . �4.15�

Thus the total angular momentum of the rotating con-densate indeed displays nonclassical behavior �see Leg-gett �2006��,

Lz � Isb���� − N� ln�l/�� . �4.16�

Note that the correction reflects the inability of the su-perfluid to rotate like a solid body. It arises directly fromthe quantization of circulation and would vanish in theclassical limit ��→0� when �→0 and nv=2� /�→�.Note that the correction Lv� becomes small in the limit of

FIG. 11. Side view of BECs in a trap. �a� Static BEC, withaspect ratio Rz /R�=1.57. Dramatically reduced aspect ratiois evident for �b� � /��=0.953 and �c� � /��=0.993. FromSchweikhard et al., 2004.

669Alexander L. Fetter: Rotating trapped Bose-Einstein condensates

Rev. Mod. Phys., Vol. 81, No. 2, April–June 2009

Page 24: Rotating trapped Bose-Einstein condensatespjh/teaching/phz7429/fetter.pdfII. Physics of Bose-Einstein Condensates in Dilute Trapped Gases 648 A. Ideal Bose gas 649 1. Ideal BEC in

a dense vortex array �l→��. As discussed in Sec. V.B.1,however, the TF limit also fails in this latter regime, anda different approach is needed.

B. Experimental and theoretical studies of vortex lattices inthe TF regime

The first images of one or a few vortices �Sec. III.D�rapidly led to the creation of large vortex arrays. Ini-tially, the MIT group �Abo-Shaeer et al., 2001� used arotating deformation �the MIT-ENS technique—see On-ofrio et al. �2000� and Fig. 6� to obtain images of trian-gular vortex lattices that are strikingly uniform, evennear the outer edge of the condensate �see Fig. 12�.These arrays closely resemble the classic triangular flux-line lattice in type-II superconductors predicted by Abri-kosov �1957� and the triangular vortex lattice in rotatingneutral incompressible superfluids predicted byTkachenko �1966a�.

The JILA group �Haljan, Coddington, et al., 2001; En-gels et al., 2003� took a quite different approach to im-parting angular momentum to the condensate. Theystarted by spinning up a normal cloud above Tc with anelliptically deformed rotating disk-shaped trap. To coolinto the BEC regime, they adiabatically alter their trapto cigar shaped and then evaporate normal atoms nearthe ends of the axis of symmetry �these atoms have smallangular momentum�. This process yields a rotating con-densate with relatively large angular momentum perparticle and good triangular arrays. These rapidly rotat-ing BECs facilitate a series of detailed experiments onthe properties of the vortex lattice.

The formation and decay of vortex lattices in rotatingBECs have generated great interest. Experiments inferthe vortex density from variable time-delay studies ofvortex visibility �Madison et al., 2001� and from the timedependence of the condensate’s aspect ratio �see Eq.�4.11�� �Abo-Shaeer et al., 2002�. In contrast, theoreticalstudies generally can follow the position of vortices asphase singularities in the GP wave function �Feder et al.,2001; Tsubota et al., 2001; Penckwitt et al., 2002; Bradleyet al., 2007�. Theoretical studies of vortex-lattice forma-tion in a paired Fermi gas �Tonini et al., 2006� find asimilar regular array at long times �compare Sec. IV.B.6for recent experiments�.

1. Collective modes of rotating condensates

The study of collective modes of a static condensateprovides a crucial test of the time-dependent GP equa-tion, as discussed by Dalfovo et al. �1999� and Pitaevskiiand Stringari �2003�. One particularly effective theoreti-cal approach �Stringari, 1996� linearizes the hydrody-namic equations �2.40� and �2.41� in the small perturba-tions �n and �v=���.

This simplest hydrodynamic formalism explicitly as-sumes irrotational flow. Thus it fails for a rapidly rotat-ing TF condensate, where the average superfluid veloc-ity has uniform vorticity ��v�2� because of thequantized vortices. To treat the dynamics of a rapidlyrotating superfluid, Sedrakian and Wasserman �2001�and Cozzini and Stringari �2003� proposed a simple gen-eralization, based on the Euler equation of nonviscoushydrodynamics, retaining the full convective term �v /�t+ �v ·��v. A familiar vector identity yields the equivalentform of the Euler equation in the laboratory frame,

M�v�t

+ ��12

Mv2 + Vtr + gn = Mv � �� � v� , �4.17�

which includes both the trap potential Vtr and the Har-tree potential gn, but omits the quantum kinetic energyfrom Eq. �2.41�, as appropriate in the TF limit. For anonrotating irrotational condensate, ��v vanishes, andEq. �4.17� then omits the right-hand side, reproducingthe gradient of Eq. �2.41� in the TF limit. In contrast, arotating TF condensate has many vortices, with ��v�2�. Thus the right-hand side of Eq. �4.17� becomes2Mv��, which Cozzini and Stringari �2003� called thelimit of “diffused vorticity,” although “mean-field” or“coarse-grained” vorticity �Sedrakian and Wasserman,2001� may be more descriptive.8 A detailed analysisgives the frequencies of the lowest m= ±2 quadrupolemodes in the laboratory frame

��m = ± 2� = �2��2 − �2 ± � . �4.18�

In particular, the difference of the two frequencies is 2�,whereas the sum of the squared frequencies is 4��

2 , in-

8A similar “average-field” approximation is central to thestudy of “anyons” and the fractional quantum Hall effect withflux attachment; see, for example, Wilczek �1990�.

FIG. 12. Observation of vortex lattices, with approximately �a� 16, �b� 32, �c� 80, and �d� 130 vortices. The diameter of the cloudin �d� was 1 mm after expansion, which represents a magnification of 20. From Abo-Shaeer et al., 2001.

670 Alexander L. Fetter: Rotating trapped Bose-Einstein condensates

Rev. Mod. Phys., Vol. 81, No. 2, April–June 2009

Page 25: Rotating trapped Bose-Einstein condensatespjh/teaching/phz7429/fetter.pdfII. Physics of Bose-Einstein Condensates in Dilute Trapped Gases 648 A. Ideal Bose gas 649 1. Ideal BEC in

dependent of the rotation rate. Haljan, Coddington, etal. �2001� fully confirmed both these detailed predic-tions.

As an extension of these experiments, Engels et al.�2002� studied the dynamics of the �originally triangular�vortex lattice in a rapidly rotating condensate. They canimage in all three perpendicular directions, which allowsthem to confirm theoretical predictions �Feder andClark, 2001; García-Ripoll and Pérez-García, 2001b�that the vortices are nearly straight in such dense arrays.Furthermore, Eq. �4.18� shows that the m= ±2 quadru-pole modes have several intriguing features in the limit�→��. In particular, Engels et al. �2002� noted that thefrequency of the m=−2 mode is small and hence nearlystationary in the laboratory frame, but the correspond-ing m=2 mode is nearly stationary in the rotating framebecause ��m=2��=��m=2�−2� then becomes small�see Eq. �3.18��.

In a particularly interesting experiment, Engels et al.�2002� formed a rotating vortex lattice with an angularvelocity � /���0.95. They then applied a relativelysmall trap distortion with ��x

2−�y2� / ��x

2+�y2��0.036 that

is stationary in the laboratory. Such a fixed perturbationis nearly resonant with the slow m=−2 quadrupolemode. As a result, the rotating condensate acquires astationary quadrupole distortion that grows to �0.4 overa time of �300 ms. Hence the rapidly rotating triangularvortex lattice experiences an oscillating shear distortionwith a period that is 1

6 times the overall rotation period2� /� because of the sixfold symmetry. Every time a ba-sis vector of the rotating lattice lines up with the minoraxis of the fixed elliptical confining potential, the vortexlattice deforms significantly, forming what are effectivelyparallel stripes of closely spaced vortices.

To understand this behavior in detail, recall that thestationary elliptical distortion of the rotating condensateinduces a quadrupolar irrotational flow �see the discus-sion at the beginning of Sec. III.D.2� that combines withthe overall rotation vsb=��r to produce the total flowvelocity v. In the laboratory frame, the vortices followstreamlines of this total local velocity. Cozzini and Strin-gari �2003� integrated the relevant dynamical equationsof motion for the vortices. They indeed found stripelikepatterns whenever one of the lattice vectors lies alongthe minor axis of the distorted condensate.

In a related approach, Mueller and Ho �2003� used avariational two-dimensional wave function. Althoughtheir formalism strictly applies only in the extreme limitof �→�� when the interaction energy is small com-pared to the single-particle energy �the “lowest Landau-level” regime, discussed in Sec. V�, their trial wave func-tion can describe a two-dimensional lattice of vorticeswith general symmetry, using Jacobian elliptic functionsthat are closely related to those in the earlier work ofTkachenko �1966a�. In their simulation, a small ellipticaldistortion of the trap potential at t=0 induces an increas-ing distortion of the condensate’s aspect ratio. The initialtriangular vortex array remains a well-defined lattice butdeforms following the combined velocity field of the ro-

tational and irrotational flow. As time evolves, stripesappear periodically, producing images that closely re-semble those found experimentally �Engels et al., 2002�.

2. Uniformity of the vortex array

Nearly 30 years ago, Campbell and Ziff �1979� studiednumerically the various equilibrium arrangements oftwo-dimensional vortex arrays in uniform incompress-ible ideal fluids inside a fixed circular boundary �as amodel for superfluid 4He in a rotating bucket�. Theyfound predominantly distorted triangular arrangementswith concentric circles of vortices and a depleted regionnear the outer edge. Yarmchuk et al. �1979� see similarpatterns in rotating superfluid 4He �see Fig. 8�, althoughthe small number of vortices precludes a definitive com-parison. In contrast, images of vortex lattices in rotatingBECs are strikingly triangular and regular, even out tothe edge of the visible condensate, as seen in Fig. 12.

The treatment of the density profile in the mean-fieldThomas-Fermi regime omits the term �v−vsb�2 entirely�see Eq. �4.8��, which assumes a uniform vortex distribu-tion and ignores the fine-grain structure of individualvortices. Sheehy and Radzihovsky �2004a, 2004b� inves-tigated these effects by considering a two-dimensionalTF condensate with number density n�r�=n�0��1−r2 /R�

2 �, where R� grows with increasing �. This rotat-ing condensate contains an array of vortices at regulartwo-dimensional lattice sites rj with mean density nv=M� /�� �the Feynman value in Eq. �4.1��. They thensubjected each vortex to a small displacement field u�r�,so that the original site moves to rj+u�rj�, and the vortexdensity becomes

nv�r� = nv�1 − � · u�r�� . �4.19�

This distorted vortex density in turn induces a change inthe mean background fluid flow based on the followingintegral:

v�r� ��

M� d2r�nv�r��

z � �r − r���r − r��2

= � � r − 2� � u�r� , �4.20�

where �=�z and the second term is a correction to thesolid-body value.

A detailed analysis with Eqs. �4.8�, �4.19�, and �4.20�expresses the deformation part of the energy E� in termsof u, and the corresponding Euler-Lagrange equationleads to the general expression

u�r� � −l2

8ln� l2

�2 � ln n�r� , �4.21�

where n�r� is the condensate number density, l isthe mean circular cell radius from Eq. �4.5�, and � is

671Alexander L. Fetter: Rotating trapped Bose-Einstein condensates

Rev. Mod. Phys., Vol. 81, No. 2, April–June 2009

Page 26: Rotating trapped Bose-Einstein condensatespjh/teaching/phz7429/fetter.pdfII. Physics of Bose-Einstein Condensates in Dilute Trapped Gases 648 A. Ideal Bose gas 649 1. Ideal BEC in

evaluated with the central density.9 In the particular caseof the two-dimensional TF density profile with n�r��1−r2 /R�

2 , this equation yields

u�r� �l2

4R�2 ln� l2

�2 r

1 − r2/R�2 . �4.22�

Note that the deformation of the regular vortex lattice ispurely radial, as might be expected from the circularsymmetry, and it increases with increasing distance fromthe center. Furthermore, the quantity R�

2 / l2 is the num-ber Nv of vortices in the rotating TF condensate �see Eq.�4.6��, so that the nonuniform distortion is of order Nv

−1

�at most a few percent�, even though the TF density de-creases significantly near the edge of the condensate.

Finally, Eq. �4.19� gives the corresponding nonuniformaxisymmetric vortex density

nv�r� = nv +1

8�ln� l2

�2 �2 ln n�r� , �4.23�

which becomes

nv�r� � nv −1

2�R�2 ln� l2

�2 1

�1 − r2/R�2 �2 �4.24�

for the two-dimensional TF condensate. The correctionis again small, of order Nv

−1. Coddington et al. �2004�studied this behavior in detail, as shown in Fig. 13. Com-parison of images in Figs. 13�c� and 13�d� show that thedistortion decreases with increasing number of vortices�namely, with increasing ��, and Fig. 13�e� restores thesuppressed zero in Fig. 13�c�, indicating that the overalleffect is indeed small.

3. Vortex core size for large rotation speeds

The principal effect of rapid rotation is the alteredaspect ratio seen in Fig. 11, but the rotation also affectsthe vortex core radius �. In the TF limit of a large con-densate, the basic definition in Eq. �2.33� shows that �2

increases with increasing � because the radial expansionreduces the central density n�0� and thus the chemicalpotential =gn�0�. To quantify this behavior, it is help-ful to consider the dimensionless ratio �2 / l2=�� / �2�,which is roughly the fraction of each vortex cell that thecore occupies. A combination of Eqs. �2.30�, �2.31�,�2.33�, and �4.9� yields the TF expression �Fetter, 2001�

�2

l2 =��

2=

�1 − �2�2/5���

�z

d�

15Na 2/5

, �4.25�

where �=� /�� is a scaled dimensionless rotationspeed.

Since d� /Na is small for a typical TF condensate, the

ratio �2 / l2 also remains small until � approaches 1, andit then grows rapidly because of the small denominator.As Fischer and Baym �2003� emphasized, this limit ofrapid rotation requires a more careful treatment. Spe-cifically, they allowed the squared vortex-core radius�rcore

2 to differ from �2 defined in Eq. �2.33�. Instead,they treated rcore as a variational parameter and found

that rcore�� for small and moderate �. As �→1, how-ever, the ratio rcore

2 / l2 saturates at a value � 12 , when the

vortex core occupies a significant fraction of each unitcell in the vortex lattice. Strictly, the TF limit no longer

applies for this large �"1 because the density variationnow makes a significant contribution to the total kineticenergy. Baym and Pethick �2004� constructed a moregeneral variational theory that includes both the TF

9As a simple application, assume a uniform particle density,so that u vanishes. Equation �4.19� then indicates that the vor-tex density is also uniform. Note that this formalism relies onthe TF energy functional and has little connection to the nu-merical studies of vortex arrays in uniform incompressible flu-ids �Campbell and Ziff, 1979�, where a vortex-free region oc-curs near the outer wall.

FIG. 13. Lattice spacing as a function of the scaled radial po-sition r /R�. The solid curve is the prediction �4.24� fromSheehy and Radzihovsky �2004a, 2004b�, with the rigid-bodyvalue as a dashed line. Plots �c� and �d� show that the fractionalamplitude decreases with increasing �, and �e� restores thesuppressed zero in �c�, indicating that the overall effect is smallbut measurable. The intervortex spacing increases by less than2% over a region where the atomic density changes by �35%.From Coddington et al., 2004.

672 Alexander L. Fetter: Rotating trapped Bose-Einstein condensates

Rev. Mod. Phys., Vol. 81, No. 2, April–June 2009

Page 27: Rotating trapped Bose-Einstein condensatespjh/teaching/phz7429/fetter.pdfII. Physics of Bose-Einstein Condensates in Dilute Trapped Gases 648 A. Ideal Bose gas 649 1. Ideal BEC in

limit and the lowest-Landau-level regime described inSec. V.

Coddington et al. �2004� studied the size of the vortexcore from both experimental and numerical perspec-tives. Because of different fits to the vortex-core densityprofiles and different definitions of the fractional corearea, their numerical values differ somewhat from thosein Eq. �4.25� and in Fischer and Baym �2003� and Baymand Pethick �2004�. Figure 14 shows typical experimen-tal and numerical values of the fractional core area com-pared with the ratio 2�� /. For small values of �, Eq.�4.25� indicates that �2 / l2 should vary linearly with � /in the TF regime. As � increases toward the criticalvalue ��, however, the fractional core area saturates at aconstant value. Detailed numerical studies �Cozzini etal., 2006� of a rapidly rotating condensate yield a verysimilar picture of this behavior.

4. Tkachenko oscillations of the vortex lattice

In 1966, Tkachenko �1966a, 1966b� published two re-markable papers on the behavior of arrays of straightvortices in an unbounded rotating incompressible irrota-tional fluid �as a model for superfluid 4He�. The firstpaper shows that a triangular lattice has the lowest en-ergy of all simple lattices �those with one vortex per unitcell�. The second paper studies small perturbations of ageneral lattice, showing that the square lattice is un-stable for waves in certain directions, but that the trian-gular lattice is stable for all normal modes. For the latterstructure, he also determines the dispersion relation ofthe small-amplitude normal modes of a vortex lattice,along with the corresponding eigenvectors for a givenwave vector k lying in the x-y plane perpendicular to �.The calculation is a tour de force of analytic functiontheory, involving several elliptic-type functions, espe-cially the Weierstrass � function that has a simple pole ateach lattice site �Chandrasekharan, 1985; Wolfram,2007�. Ultimately, however, the final result is simple. Atlong wavelengths kl�1, where l=�� /M� is the vortex-cell radius, the wave is predominantly transverse with alinear dispersion relation

��k� � cTk , �4.26�

where

cT =12���

M=

12

l� �4.27�

is the speed of propagation. This long-wavelength mo-tion is effectively a transverse phonon in the vortexlattice.10

The radial expansion of rapidly rotating BECs meansthat the condensate is essentially two dimensional. Thusbending modes of the vortices are irrelevant, but thenonzero compressibility of these atomic gases requires asignificant modification to Tkachenko’s �1966b� analysis.Sonin �1987� and Baym �2003� generalized Tkachenko’sresult in Eq. �4.26� to find the new long-wavelength ex-pression

��k�2 � cT2 k2 s2k2

4�2 + s2k2 , �4.28�

where s is the speed of sound. This expression assumesan infinite uniform system, but Anglin �2002�; Baym�2003�; Baksmaty et al. �2004�; and Sonin �2005� also in-cluded the nonuniformity of the trapped BEC. If sk� �the short-wavelength or incompressible limit�, thenthis expression reduces to Tkachenko’s ��k��cTk, but ifsk�� �the long-wavelength or compressible limit�, thenthe mode becomes soft with ��k��k2. Similar softeningof the collective-mode spectrum has important conse-quences for stability of the vortex lattice at large � �Si-nova et al., 2002; Baym, 2005�, as discussed in Sec. VI.

Coddington et al. �2003� observed these Tkachenkooscillations in considerable detail. They started with anessentially uniform vortex lattice and then applied aweak perturbation �along with other approaches, theyused a “blasting pulse” that removes a small fraction ofthe atoms near the center of the trap�. At a scaled rota-

tion frequency �=0.95, they waited a variable time andthen turned off the trap, yielding an image of the result-ing distorted vortex lattice. Figure 15 shows the vortexlattice at one-quarter and three-quarters of the oscilla-tion period, with the lines as a sine fit to the distortion ofthe vortex lattice. In practice, the normal mode has thecorrect shape, but the measured frequency is signifi-cantly smaller than that predicted in Eq. �4.28�. Anglinand Crescimanno �2002� suggested that this discrepancyarises from reliance on a continuum theory that ignoresthe increased vortex core radius for large �.

Tkachenko �1969� also evaluated the shear modulus ofthe triangular vortex lattice in an incompressible fluid,which is directly related to the long-wavelength oscilla-tion spectrum. Subsequently, Cozzini et al. �2006� usednumerical methods to study the more general situation

10Unlike Newtonian particles in a two-dimensional lattice,however, vortices obey first-order dynamical differential equa-tions. Hence there is only one normal mode for each wavevector k.

FIG. 14. Fractional vortex core area as a function of 2�� /.The dashed line is the TF theory and the dotted line is thelowest-Landau-level theory discussed in Sec. V. From Cod-dington et al., 2004.

673Alexander L. Fetter: Rotating trapped Bose-Einstein condensates

Rev. Mod. Phys., Vol. 81, No. 2, April–June 2009

Page 28: Rotating trapped Bose-Einstein condensatespjh/teaching/phz7429/fetter.pdfII. Physics of Bose-Einstein Condensates in Dilute Trapped Gases 648 A. Ideal Bose gas 649 1. Ideal BEC in

of a vortex lattice in a rotating BEC, which also includesboth the incompressible �TF� limit and the compressible�LLL� limit �Sinova et al., 2002; Sonin, 2005�.

5. Rotating two-component BECs

The experimental creation of relatively large vortexarrays in single-components BECs in 23Na �Abo-Shaeeret al., 2001� and 87Rb �Engels et al., 2002� rapidly led toproposals for corresponding two-component vortex ar-rays �Mueller and Ho, 2002; Kasamatsu et al., 2003�. Asdiscussed in Sec. III.D.1, the interspecies interaction pa-rameter g12 significantly affects the behavior, especiallyfor trapped condensates �the intraspecies parameters g1and g2 are assumed positive�.

Mueller and Ho �2002� initiated the theory of thesetwo-component condensates, assuming rapid rotation��"���. This limit allows a considerable simplification�see Sec. V� because the condensate wave functions �1and �2 can then be constructed from the lowest-Landau-level �LLL� single-particle states �the same asthose used by Butts and Rokhsar �1999� in Sec. III.C�.Apart from an overall Gaussian factor, the resulting trialsolutions �j are analytic functions of the complex vari-able �=x+ iy; they involve quasiperiodic Jacobi thetafunctions that are closely related to other elliptic func-tions appearing in the treatment of normal modes in arotating vortex lattice �Chandrasekharan 1985;Tkachenko, 1966a, 1966b�, as mentioned in Sec. IV.B.4.This analytical approach allows a relatively simple treat-ment of the change in the vortex lattice as the parameter��g12/�g1g2 varies from negative �when the interspe-cies interaction is attractive� to positive. Specifically, fornegative �, the two components prefer to overlap andform a single triangular lattice. As � becomes increas-ingly positive, however, the overall repulsive interactionenergy leads the two components to separate spatially,with interlaced triangular vortex arrays to minimize theoverall density variation. Eventually, the triangular lat-tices should distort, forming square or rectangular arrayswhen ��1.

Subsequently, Kasamatsu et al. �2003� studied the be-havior at lower angular velocity in the mean-field TFregime. Their extensive numerical analysis of the equi-librium vortex-lattice configuration as a function of the

scaled dimensionless rotation rate �=� /�� and therelative interaction strength � gives a qualitatively simi-

lar picture: For any reasonable �, the interpenetratingvortex arrays assume fourfold square symmetry as � ap-proaches 1− from below.

Schweikhard et al. �2004� performed a series of experi-

ments on these rotating two-component systems ���0.75�. They started with a single-component triangularvortex array in the state �1�= �F=1,mF=−1� �compareSec. III.D.1�. A short pulse with the two-photon electro-magnetic coupling fields transfers �80% of the popula-tion to the state �2�= �F=2,mF=1�. The initial triangularvortex lattice is now unstable, and after a turbulent stage��2 s� it transforms to a square structure that persistsfor another �4 s. Ultimately, however, the state �2� de-cays through a spin-flip transition. As a result, the re-maining single-component condensate in state �1� againtransforms back to a triangular array, similar to the ini-tial state.

6. Rotating gas of paired fermions in a tightly bound (BEC)limit

One of the many remarkable new results from thestudy of ultracold gases is the superfluidity of boundpairs of fermionic atoms such as 6Li and 40K. Although adetailed discussion is inappropriate here, Leggett �2006�,Bloch et al. �2007�, and Giorgini et al. �2008� providetreatments of the behavior near a Feshbach resonance.Here, I review the basic elements of scattering theory toexplain how the s-wave scattering length can exhibit aresonance associated with the appearance of a newbound state.

The usual description of a dilute atomic BEC assumesa short-range interatomic potential with V�r−r���g��3��r−r��, where g is the effective interaction con-stant with dimension of energy�volume. Two-bodyscattering theory shows that g is proportional to thes-wave scattering length a �see the discussion below Eq.�2.21��. This picture holds for the bosonic atoms 87Rband 23Na, where both g and a are positive, and suchsystems are stable both in bulk and in a trap, with avariety of associated collective modes. In contrast, thebosonic atom 7Li has an attractive effective interactiong 0 with a negative a 0. In bulk, the negative interac-tion would yield an imaginary speed of sound, implyinga long-wavelength instability. In a trap, however, thepositive kinetic and potential energies can stabilize thecondensate for sufficiently small N, but the system willcollapse �Ruprecht et al., 1995; Bradley et al., 1997�above a critical number of atoms Nc. As an alternativepicture of this stabilization, note that the quantized wavenumbers in a trap provide a lower bound kmin�� /R,where R depends on N. If the discrete analog of theBogoliubov energy spectrum is real for all k#kmin, thenthe trapped condensate is stable even for negative a.

The preceding simple direct relation between the signof the effective two-body interaction g and the sign of

FIG. 15. Lowest Tkachenko mode of the vortex lattice excitedby atom removal taken �a� 500 ms after the end of the blastingpulse and �b� 1650 ms after the blasting pulse. Lines are sinefits to the distortion of the vortex lattice. From Coddington etal., 2003.

674 Alexander L. Fetter: Rotating trapped Bose-Einstein condensates

Rev. Mod. Phys., Vol. 81, No. 2, April–June 2009

Page 29: Rotating trapped Bose-Einstein condensatespjh/teaching/phz7429/fetter.pdfII. Physics of Bose-Einstein Condensates in Dilute Trapped Gases 648 A. Ideal Bose gas 649 1. Ideal BEC in

the scattering length a holds only for weak two-bodypotentials that have no bound states. The situation be-comes more interesting when the two-body potential cansupport one �or more� s-wave bound states. In this case,the s-wave scattering length exhibits divergent resonantbehavior. To be very specific, consider a sphericalsquare-well potential V�r�=V0��b−r� with height V0 andrange b, where � denotes the unit positive step function.Here the volume integral of V�r� yields g��4� /3�V0b3,so that the sign of V0 is the same as that of g.

�1� Bound states occur only for negative V0=−�V0�. Astandard analysis �see, for example, Landau and Lif-shitz �1965� and Shankar �1994�� shows that thereare no s-wave bound states if �V0� �2�2 /8Mb2; as�V0� increases, the nth s-wave bound state appearswhen

�V0� =�2n − 1�2�2�2

8Mb2 . �4.29�

�2� Consider now the scattering of a particle with posi-tive energy E=�2k2 / �2M�. The radial l=0 wavefunction ��r�=u�r� /r has the asymptotic form u�r�=sin�kr+��, where � is the s-wave phase shift �notethat � depends on k and hence on energy�. A de-tailed analysis for the same square-well potentialV�r� �see, for example, Landau and Lifshitz �1965�and Shankar �1994�� yields the explicit expression

� = − kb + arctan� k

k�tan k�b , �4.30�

where k�2=2M�E−V0� /�2 determines the wavenumber k� inside the square well.11 It is evidentphysically that � is negative for positive V0 becausethe repulsive potential pushes out the wave func-tion, whereas � is positive for negative V0 becausethe attractive potential pulls in the wave function.This phase shift completely characterizes the s-wavescattering by this potential.

It is conventional and convenient to introduce thes-wave scattering length a�−limk→0 ��k� /k; for thesquare-well potential, this result gives

a = b −tan k�b

k�, �4.31�

where k�2=−2MV0 /�2 is now evaluated for k=0 �andhence E=0�.12 If V0 is positive and large, then a→b, butthe dependence can be very different in other situations.For example, a vanishes if V0=0 �as expected�. Moresignificant is the behavior for negative V0, especiallynear the appearance of a bound state. Equation �4.29�

shows that k�b increases through �n− 12 �� when the nth

s-wave bound state first forms. Consequently, Eq. �4.31�diverges to negative infinity when the state is just un-bound and then decreases from positive infinity whenthe trapped state is weakly bound. Such resonant depen-dence is a typical and generic feature of scattering by anattractive potential that can support one or more boundstates.

In the context of dilute ultracold gases, the attractivelong-range van der Waals potential acts to bind the twofermionic atoms to form a molecule. In general, thereare several molecular states, each with different mag-netic moments. Thus an applied magnetic field can shifta bound state in a closed channel across the asymptoticcontinuum in an open channel, leading to a “Feshbachresonance” �see, for example, Moerdijk et al. �1995�,Castin �2006�, Köhler et al. �2006�, and Bloch et al. �2008�for more detailed discussions�. In essence, sweeping anexternal magnetic field through a Feshbach resonancetransforms a tightly bound bosonic molecule of the twofermions into a weakly attracting fermionic pair. Thebosonic molecules can form a BEC that displays essen-tially all the properties of an atomic BEC, whereas theweakly attracting fermionic pairs are analogous to theCooper pairs in the BCS theory of conventional low-temperature superconductors �de Gennes, 1966;Tinkham, 1996�. As a result, this transition through aFeshbach resonance is called the “BEC-BCS crossover.”

Particularly relevant in the present context is the ex-periment of Zwierlein et al. �2005� who rotated a gas ofultracold 6Li with a deformed optical trap �the magneti-cally induced Feshbach resonance precludes the use ofthe usual magnetic trap�. Depending on the appliedmagnetic field, the rotating gas can be in the bosoniclimit, when the tightly bound molecules form a BEC, orin the fermionic �BCS� limit of two weakly bound Fermiparticles, or in the intermediate crossover regime. Figure16 shows the vortex lattice at three different values ofthe magnetic field. Note that the scattering length a di-verges in the resonance region. Consequently, it is con-ventional to use �kFa�−1 as the independent variable,

11If V0�E, then k� is written as i� and tan k�b /k� becomestanh �b /�.

12If V0 is positive, then this relation is replaced by �2

=2MV0 /�2; correspondingly, Eq. �4.31� becomes a=b−tanh �b /�.

FIG. 16. Vortex lattice in a rotating gas of 6Li fermions. Theleft side is on the BEC side, the central is on resonance, andthe right side is on the BCS side �note that the lower horizontalscale reverses positive and negative �kFa�−1, which is effec-tively �n1/3a�−1 apart from a numerical factor�. From Zwierlein,2007.

675Alexander L. Fetter: Rotating trapped Bose-Einstein condensates

Rev. Mod. Phys., Vol. 81, No. 2, April–June 2009

Page 30: Rotating trapped Bose-Einstein condensatespjh/teaching/phz7429/fetter.pdfII. Physics of Bose-Einstein Condensates in Dilute Trapped Gases 648 A. Ideal Bose gas 649 1. Ideal BEC in

which varies smoothly from positive in the BEC limitthrough zero near the crossover region to negative in theBCS limit. Here kF is the Fermi wave number, but, apartfrom a numerical factor of order unity, it is the same asn1/3, with n the boson density. The observation of quan-tized vortices on both sides of the BEC-BCS crossoverprovides convincing evidence for superfluidity on thefermionic side, although the experiment did not reachthe weakly bound BCS limit �Giorgini et al. �2008� gavea detailed discussion of rotating ultracold Fermi gases�.

Sensarma et al. �2006� used the fermionicBogoliubov–de Gennes equations �see, for example, deGennes �1966� and Tinkham �1996�� to study the struc-ture of a single quantized vortex in a uniform three-dimensional atomic Fermi gas at zero temperature. Nearthe BEC-BCS crossover, the interparticle spacing �kF

−1

provides a second �shorter� length scale in addition tothe BCS coherence length. Here I consider only theBEC side, since a full summary of ultracold Fermi gasesis beyond the scope of this review.

We focus on the vortex core radius �healing length� �defined in Eq. �2.33� for a trapped condensate. As notedin Sec. IV.B.3, � grows with increasing rotation speed �because of the reduced central density �and hence re-duced chemical potential ����, but Eq. �4.25� also ex-hibits the explicit dependence on the scattering length a.Here it is preferable to consider the slightly differentdimensionless ratio

d0=

1

�1 − �2�1/5� d0

15Na 1/5

, �4.32�

which involves only the scaled rotation speed �=� /��,the s-wave scattering length a, and the trap geometrythrough d0 and the radial trap frequency ��.

Sensarma et al. �2006� remarked that the healinglength in a uniform cold BEC Eq. �2.23� is proportionalto �na�−1/2 and decreases with increasing positive a forfixed n. Consequently, the vortex core size should shrinkon approaching the resonance regime from the BECside. Note, however, that inclusion of the trap in Eq.�4.32� alters the fractional power from a−1/2 to a−1/5. Thusthe dependence on a may be difficult to measure in cur-rent experiments. It is also clear from Eq. �2.24� that theGP picture fails as � approaches the interparticle spac-ing, which is expected to be the minimum value for �.On the BCS side, the qualitative relation ���vF /�shows that the vortex core again grows on moving awayfrom the resonance regime because the BCS energy gap� then becomes small �Sensarma et al., 2006; see alsoLeggett, 2006�.

On the BEC side, recall that the condensate densityn�r� around a vortex decreases rapidly from its bulkvalue n for r"� inside the vortex core. Consequently,the circulating mass current j�r�=Mn�r�v�r�=n�r�� /rreaches a maximum “critical” value jc�n� /� for r���compare the discussion below Eq. �3.4��. As a increaseson the BEC side toward the Feshbach resonance, thedecreased � implies that this critical mass current jc

grows and saturates when ��n−1/3. A more detailedanalysis by Sensarma et al. �2006� showed that jc thenagain decreases on moving from the resonance regimeto the BCS side. Thus both � and jc display nonmono-tonic dependence in traversing the BEC-BCS crossoverregion.

C. Effect of additional quartic confinement

The discussion of rotating condensates in the mean-field Thomas-Fermi regime �Sec. IV.A� makes it clearthat a purely harmonic trap with radial frequency ��

will cease to confine the system if the rotation speed �reaches or exceeds ��. Indeed, the limit �→�� is sin-gular because the condensate radius and the total angu-lar momentum both diverge. In the regime ��−����,the system is expected to undergo a series of crossoversand complicated phase transitions �Secs. V and VI�whose exact nature remains under debate.

Thus it is interesting to consider the effect of addingan additional strongly confining potential that rises morerapidly than r2. Here I focus on the most common �quar-tic� model �r4 �Fetter, 2001; Kasamatsu et al., 2002;Lundh, 2002; and Kavoulakis and Baym, 2003�. Such ananharmonic trap will confine the condensate even for����, thus allowing a more controlled study of pos-sible new states �they are not generally the same asthose predicted in a harmonic trap in the limit �→���.In terms of the usual dimensional quantities, the anhar-monic radial trap potential has the form

Vtr�r� =���

2� r2

d�2 + �

r4

d�4 , �4.33�

where d�=�� /M�� is the radial oscillator length and �is a dimensionless parameter that characterizes the ad-mixture of quartic contribution.

For simplicity, I focus on the limit of rapid rotationand consider only a two-dimensional condensate that isuniform in the z direction over a length Z �Kavoulakisand Baym, 2003; Fetter et al., 2005�, with ��r ,z�=��r� /�Z. It is also convenient to measure energies inunits of ��� and lengths in units of d�, and to normalizethe dimensionless condensate wave function � with�d2r���2=N. In the mean-field TF regime, the dimen-sionless energy in the rotating frame �compare Eq. �4.7��becomes

E���� =� d2r�12

�v2 + r2 + �r4����2 − � · r � v���2

+12

g���4� , �4.34�

where g=4�a /Z is a dimensionless two-dimensional in-teraction parameter. An alternative model for the axialconfinement uses a tight harmonic trap, in which casethe interaction parameter is g=�8�a /dz, but the subse-quent description is unaffected.

676 Alexander L. Fetter: Rotating trapped Bose-Einstein condensates

Rev. Mod. Phys., Vol. 81, No. 2, April–June 2009

Page 31: Rotating trapped Bose-Einstein condensatespjh/teaching/phz7429/fetter.pdfII. Physics of Bose-Einstein Condensates in Dilute Trapped Gases 648 A. Ideal Bose gas 649 1. Ideal BEC in

When the rotating condensate has many vortices, thetotal velocity v closely approximates the solid-body form��r, in which case Eq. �4.34� reduces to

E���� =� d2r�12

��1 − �2�r2 + �r4����2 +12

g���4� .

�4.35�

As in Eq. �2.27�, variation with respect to ���2 at fixednormalization yields two-dimensional density

g���r��2 = + 12 ���2 − 1�r2 − �r4� . �4.36�

For � 1, the density has a local maximum at the cen-ter, but it changes to a local minimum for ��1. In ad-dition, the central density ���0��2 is proportional to thechemical potential . As in Eq. �4.9�, depends on �through the normalization and decreases continuouslywith increasing �, reflecting the radial expansion of thecondensate.

In contrast to the mean-field TF picture for a har-monic radial trap, however, the chemical potential canvanish at a critical angular velocity �h�1, when the cen-tral density also vanishes. This behavior indicates theformation of a central hole; for ���h, the chemicalpotential becomes negative and the condensate as-sumes an annular form �see Fetter et al. �2005� for adescription of this unusual structure�.

As � continues to increase, the system can, in prin-ciple, make a second transition to a different annularstate with purely irrotational �vortex-free� circulatingflow, known as a “giant vortex.” In this case, the circu-lating velocity v�r� has the quantized form Nv /r, reflect-ing the presence of Nv phantom vortices in the hole�equivalently, the flow simply represents Nv-fold quan-tized circulation in a multiply connected condensate�.Kavoulakis and Baym �2003� suggested the possibility ofsuch irrotational flow in a quadratic plus quartic poten-tial �earlier, Fischer and Baym �2003� proposed a similarstructure in a rotating rigid cylinder�. As noted at thestart of Sec. IV, solid-body rotation �the model used inobtaining Eq. �4.36�� gives the absolute minimum energyfor a rotating system. Consequently, any irrotational gi-ant vortex necessarily depends on the discrete quantiza-tion of circulation that introduces graininess in the su-perflow. In practice, estimates of the critical angularvelocity for formation of a giant vortex require consid-erable numerical work �see Kavoulakis and Baym�2003�; Fetter et al. �2005�; Kim and Fetter �2005�; Fuand Zaremba �2006� for more details�.

In an elegant series of experiments, the ENS �Paris�group �Bretin et al., 2004; Stock et al., 2004� studied the

effect of a weak quartic confinement in addition to theusual quadratic trap potential �see also Stock, Battelier,et al. �2005� for a more general review of rotating BECs�.Specifically, they use a blue-detuned �repulsive� Gauss-ian laser beam propagating along the symmetry axis oftheir elongated condensate, with a resulting dipole po-tential

Vd�r� = V0e−2r2/w2� V0 −

2V0

w2 r2 +2V0

w4 r4 + ¯ .

�4.37�

The positive constant term V0 simply shifts the zero ofenergy, the quadratic term renormalizes �reduces� theoriginal radial trap frequency, and the quartic term pro-vides weak positive confinement. The measured oscilla-tion frequency of the center of mass of the condensategives the effective radial trap frequency �eff /2�=64.8 Hz �Stock et al., 2004�, and the other trap param-eters yield the quartic coupling constant ��10−3, indi-cating that the added confinement in Eq. �4.33� is indeedweak.

For the dimensionless rotation speed ��1, the rota-tional deformation produces a nearly spherical conden-sate that remains stable up to �"1.05 �Bretin et al.,2004�. Figure 17 shows the expanded cloud for variousrotation speeds both below and above the critical value�eff. Initially, a regular vortex array appears, but it be-comes irregular in Figs. 17�d�–17�f�. In addition, for ���eff, the condensate exhibits a clear local minimum inthe density near the center, as expected from Eq. �4.36�in the mean-field TF picture. Throughout the analysis,they fitted the measured condensate shape to the TFprediction, which provides a direct determination of �.They also used surface-wave spectroscopy as discussedin Sec. III.D.3 �Svidzinsky and Fetter, 1998; Zambelliand Stringari, 1998; Cozzini and Stringari, 2003� to ob-tain an independent measure of � �even though thenumber of visible vortices appears to be too small for�#�eff�.

The absence of visible vortices in Fig. 17�g� is puz-zling. One possible reason for the discrepancy is thethree-dimensional character of the condensate, which al-lows the vortices to bend. Also, the low temperaturemay make it difficult for the system to equilibrate. Nu-merical studies �Aftalion and Danaila, 2004� suggestedthat much longer times are needed to attain a well-ordered vortex lattice in these regimes of large �. In anycase, Fig. 17�h� implies that the condensate simply col-lapses, although the reason remains unknown.

FIG. 17. Pictures of the rotating gas taken along the rotation axis. The number below each picture is the rotation frequency in Hz�note that �eff /2�=64.8 Hz is the effective radial trap frequency�. From Bretin et al., 2004.

677Alexander L. Fetter: Rotating trapped Bose-Einstein condensates

Rev. Mod. Phys., Vol. 81, No. 2, April–June 2009

Page 32: Rotating trapped Bose-Einstein condensatespjh/teaching/phz7429/fetter.pdfII. Physics of Bose-Einstein Condensates in Dilute Trapped Gases 648 A. Ideal Bose gas 649 1. Ideal BEC in

V. VORTEX ARRAYS IN MEAN-FIELD LLL REGIME

In the mean-field Thomas-Fermi description of a ro-tating BEC, the basic approximation is the neglect of the“kinetic energy” associated with the density variations.Specifically, Eq. �4.7� omits the term �2���� � �2 / �2M�whereas the usual kinetic energy from the phase varia-tion 1

2Mv2���2 remains. As seen in Eq. �4.25�, the vortexcores are small for moderate values of the dimensionless

rotation speed �=� /��, so that this approximation is

valid. As � increases toward 1, however, it fails when

1−��d� /15Na�1 because the vortex cores then be-come comparable with the intervortex separation �2l.In this case, it becomes essential to return to the full GPenergy functional E���� in Eq. �3.6�.

In thinking about this new regime, Ho �2001� ob-served that the interaction energy per particle �of order���=gn�0�� is small because the condensate expandsradially �see Eq. �4.9��. In essence, the condensate be-comes two dimensional, and it is convenient to study atwo-dimensional condensate that is uniform in the z di-rection over a length Z. Thus the original condensatewave function ��r ,z� can be written as ��r� /�Z, where��r� is a two-dimensional wave function with normaliza-tion �d2r���2=N. The GP energy functional then be-comes

E���� =� d2r�*� p2

2M+

12

M��2 r2 − �Lz +

12

g2D���2 � ,

�5.1�

where g2D=g /Z. As in Sec. IV.C, one can instead as-sume a tight Gaussian axial confinement, in which caseg2D=g /�2�dz, with dz=�� /M�z.

A. Physics of LLL one-body states

The one-body Hamiltonian in Eq. �5.1� has the crucialadvantage that it is exactly soluble with simple energyeigenvalues �see, for example, Messiah �1961� andCohen-Tannoudji et al. �1977��. Consider the �nonrotat-ing� two-dimensional isotropic harmonic oscillator withHamiltonian

H0 =p2

2M+

12

M��2 r2. �5.2�

Classically, the motion is that of two independent oscil-lators �along x and y� that can combine to give, for ex-ample, circular motion with frequency �� in the positive�counterclockwise� or negative �clockwise� sense, de-pending on the relative phases. When viewed from aframe that rotates with angular velocity � in the positivesense, the two circular orbits with originally degeneratefrequencies now have different frequencies

�± = �� �� , �5.3�

since the positive �negative� mode has a reduced �in-creased� frequency in the rotating frame.

The quantum-mechanical description relies on the fa-miliar creation and annihilation operators

ax =1�2

� x

d�

+ ipxd�

� , ax

† =1�2

� x

d�

− ipxd�

� , �5.4�

along with similar operators ay and ay†. In the present

case, it is more appropriate to use circularly polarizedstates with operators

a± =ax � iay

�2, a±

† =ax

† ± iay†

�2�5.5�

that obey the usual commutation relations �a+ ,a+†�

= �a− ,a−†�=1, with all other commutators vanishing. A

straightforward analysis �Messiah �1961� and Cohen-Tannoudji et al. �1977�� expressed H0 in terms of thesenew operators

H0 = ����a+†a+ + a−

†a− + 1� . �5.6�

Similarly the angular momentum Lz=xpy−ypx becomes

Lz = ��a+†a+ − a−

†a−� . �5.7�

The number operators a±†a± have non-negative integer

eigenvalues n±; the corresponding operators a+† and a+

�a−† and a−� create and destroy one quantum with posi-

tive �negative� circular polarization and one unit of posi-tive �negative� angular momentum. For the nonrotatingsystem with Hamiltonian H0, each quantum simply hasan energy ���.

The principal advantage of this particular circularlypolarized basis appears for the rotating system, withHamiltonian H0�=H0−�Lz, because Eqs. �5.6� and �5.7�together imply

H0� = ��� + ��+a+†a+ + ��−a−

†a−, �5.8�

where �±=���� are the frequencies in Eq. �5.3� forthe positive and negative classical orbits as seen in therotating frame. Correspondingly, the energy eigenvaluesare labeled with two non-negative integers n±,

��n+,n−� = n+���� − �� + n−���� + �� , �5.9�

where I omit the zero-point energy ��� for simplicity. Inthe limit of rapid rotations ��→���, these eigenvaluesare essentially independent of n+, which implies a largedegeneracy. The other integer n− then becomes theLandau-level index, with different Landau levels sepa-rated by an energy gap of �2���.

For small angular velocity, it is often convenient tointroduce a different set of quantum numbers �Cohen-Tannoudji et al. �1977�� n=n++n− for the energy and m=n+−n− for the angular momentum, so that the energyeigenvalues have the equivalent form

�nm � �� 12 �n + m�, 1

2 �n − m�� = n��� − m�� . �5.10�

Figure 18 uses this new basis to illustrate the energyspectrum �nm for different values of �.

�1� On the left side of the figure ��=0�, the excitationenergy is simply n���, independent of m, forming

678 Alexander L. Fetter: Rotating trapped Bose-Einstein condensates

Rev. Mod. Phys., Vol. 81, No. 2, April–June 2009

Page 33: Rotating trapped Bose-Einstein condensatespjh/teaching/phz7429/fetter.pdfII. Physics of Bose-Einstein Condensates in Dilute Trapped Gases 648 A. Ideal Bose gas 649 1. Ideal BEC in

an inverted pyramid of states: for each non-negativeinteger n, there are n+1 degenerate angular-momentum states ranging from −n to n in steps of 2.

For small ��1, the energy quantum number n de-termines the large splitting and the angular-momentum quantum number m lifts the remainingdegeneracy �physically, this Coriolis-induced split-ting is like the weak-field Zeeman splitting of mag-netic sublevels�.

�2� The central figure shows the situation for moderate�, where the �now signficant� Coriolis effect elimi-nates the degeneracy of the nonrotating spectrum.States with m=0 are unshifted, states with negativem shift up by �m ���, whereas those with positive mshift down by the same amount.

�3� The right side of the figure shows the extreme situ-ation for �→��, when the states again becomenearly degenerate, forming essentially horizontalrows �the Landau levels�. The lowest Landau levelhas n=m, which means n−=0 and n+=m; the firstexcited Landau level has n−m=1 �namely, n−=1

and n+=m+1�, etc. In this limit ��"1�, the quantumnumbers n+ and n− of the rotating basis are mostconvenient.

For rapid rotation ��"1�, it is appropriate to focus onstates in the lowest Landau level �those with n−=0, de-noted LLL�, when the single quantum number n+=msuffices to specify the one-body states. The relevant en-ergies are the gap 2��� between adjacent Landau levelsand the mean interaction energy gn�0�=, which is as-sumed small in the present limit. Thus the basic expan-sion parameter of the theory is / �2����, although cur-rent experiments have only achieved values as small as0.6 �Schweikhard et al., 2004�. The corresponding two-dimensional eigenfunctions �m�r� of both H0� and Lzhave a simple form in plane-polar coordinates �note that

these are precisely the trial states from Eq. �3.19� used tostudy vortex arrays in a weakly interacting condensate�Butts and Rokhsar, 1999��

�m�r� � rmeim!e−r2/�2d�2 �, �5.11�

where the normalization is unimportant for the presentpurpose.

In particular, the ground state �0 is just a two-dimensional isotropic Gaussian. It represents thevacuum for both circularly polarized modes

a±�0 = 0. �5.12�

The higher states �m for positive m are proportional to�a+

†�m�0; they can be written in terms of �= �x+ iy� /d�

= �r /d��ei! as �m��me−r2/�2d�

2 �, where the first factor is anon-negative integral power of the dimensionless com-plex variable �. For a given m1, the density ��m�r��2 islarge only in a circular strip of radius ��md� and width�d� /�m. In addition, it is easy to see that the mean-square radius for any non-negative m is given by

�r2�m = �m + 1�d�2 . �5.13�

Since the same LLL wave functions appear in the frac-tional quantum Hall effect for two-dimensional elec-trons in a strong magnetic field, this mean-field LLL re-gime is sometimes called the “mean-field quantum Hall”regime �Watanabe et al., 2004�. In fact, related many-body states from the fractional quantum Hall effect playan important role for a rotating dilute Bose gas at stillhigher angular velocity, where the physics will turn outto be quite different �Sec. VI�. Thus it seems preferableto reserve the name “quantum Hall regime” for thismore unusual �nonsuperfluid� limit.

B. Direct treatment of LLL states

Before using the LLL wave functions to construct aGP condensate wave function, I briefly discuss an alter-native approach to the LLL states. It is natural to re-write the single-particle Hamiltonian in the rotatingframe H0�=p2 / �2M�+ 1

2M��2 r2−� ·r�p by completing

the square to obtain two equivalent forms:

H0� =�p − M� � r�2

2M+

12

M���2 − �2�r2, �5.14�

H0� =�p − M�� � r�2

2M+ ��� − �� · L , �5.15�

where ��=��z. Either form is reminiscent of the non-relativistic Hamiltonian of a particle with charge q in amagnetic field B, where the momentum p and vectorpotential A appear in the gauge-invariant combinationp−qA, with B=��A.

Landau solved the quantum problem of an electron ina uniform magnetic field B �see, for example, Landauand Lifshitz �1965�, but the vector potential has a differ-ent �asymmetric� form A=−Byx�. In contrast, Cohen-Tannoudji et al. �1977� solved the same problem in the

FIG. 18. Excitation energy levels �nm=n���−m� from Eq.�5.10� for various angular velocity � expressed in terms ofquantum numbers n=n++n− for energy and m=n+−n− for an-gular momentum. Left side is for �=0, showing an invertedpyramid of states: for each non-negative integer n, there aren+1 degenerate angular-momentum states ranging from −n ton in steps of 2. Center shows the situation for moderate �.States with m=0 are unshifted, states with negative m shift upwhereas those with positive m shift down. Right side shows theextreme situation for �→��, when the states form essentiallyhorizontal rows �the Landau levels�. The lowest Landau levelhas n=m, which means n−=0, the first excited Landau levelhas n−m=1 �namely, n−=1�, etc. From Dalibard, 2007.

679Alexander L. Fetter: Rotating trapped Bose-Einstein condensates

Rev. Mod. Phys., Vol. 81, No. 2, April–June 2009

Page 34: Rotating trapped Bose-Einstein condensatespjh/teaching/phz7429/fetter.pdfII. Physics of Bose-Einstein Condensates in Dilute Trapped Gases 648 A. Ideal Bose gas 649 1. Ideal BEC in

symmetric gauge A= 12B�r. It is of course possible sim-

ply to quote those results to obtain the LLL wave func-tions �m, but such an approach requires a charge q,whose sign �positive or negative� determines some cru-cial conventions in the choice of quantum-mechanicalstates. As written, Eq. �5.15� suggests a unit positivecharge with fictitious vector potential A=M���r; theresulting uniform magnetic field is B=2M��. Unfortu-nately, the usual LLL picture treats electrons with nega-tive charge, and the corresponding one-body states in-volve reinterpretations of some of the quantumnumbers, as discussed by Cohen-Tannoudji et al. �1977�.For this reason, I choose to work directly from the origi-nal Hamiltonian H0�=H0−� ·L, because the role of n+and n− is particularly transparent.

Another interesting feature of the transformation toa quadratic form in p−qA is the possibility of topolo-gical gauge potentials. In the context of optical lattices,Jaksch and Zoller �2003� proposed introducing variousterms in the single-particle Hamiltonian that mimic theeffect of an external magnetic field. Specifically, when atwo-level atom hops around a closed path in the lattice,external lasers can generate a net phase change that isessentially equivalent to the magnetic flux of an appliedfield �see Mueller �2004�; Sørensen et al. �2005�; Satija etal. �2006� for alternative approaches, including the pos-sibility of “non-Abelian� gauge potentials�.

Similar ideas also apply to a simple trapped conden-sate. Juzeliunas et al. �2005� proposed using a three-levelsystem in the & configuration �two nearly degeneratelower levels, each coupled to a higher third state by ex-ternal laser beams �Arimondo, 1996�; for a “dressed-atom approach” to these three-level systems, see, forexample, Cohen-Tannoudji et al. �1998��. Such an inter-acting three-level system has a “dark” state that effec-tively uncouples from the remaining states. If one of theexternal lasers has orbital angular momentum �see, forexample, Gottfried �1966� and Allen et al. �1999, 2003��,then the dark state experiences a nontrivial topologicalgauge potential that in principle can act like an essen-tially arbitrary magnetic field �and hence an arbitraryapplied vorticity for the trapped condensate�. Whethersuch methods can indeed generate angular momentumin the condensate and the associated vortices remainsuncertain. More work �both theory and experiment� isclearly desirable.

C. LLL condensate wave functions

Ho �2001� proposed a linear combination of the LLLsingle-particle states �m in Eq. �5.11� as the GP conden-sate wave function for the rapidly rotating two-dimensional BEC,

�LLL�r� = m#0

cm�m�r� = f���e−r2/�2d�2 �. �5.16�

Here �d2r��LLL�2=N fixes the normalization and f���=m#0cm�

m is an analytic function of the complex vari-able � �compare Sec. III.C; see also Shankar �1994��.

Specifically, for such a truncated basis set, the analyticfunction f��� is a complex polynomial and thus has afactorized form

f��� � �j

�� − �j� , �5.17�

apart from an overall constant factor �I assume eachzero is distinct�. Evidently, f��� vanishes at each of thepoints �j, which are the positions of the nodes of thecondensate wave function �LLL�r�. In addition, the phaseof this wave function increases by 2� whenever � movesin the positive sense around any of these zeros. Thus thepoints �j are precisely the positions of the vortices in thetrial state �LLL, and minimization with respect to theconstants cm is effectively the same as minimization withrespect to the position of the vortices �Butts and Rokh-sar, 1999�.

1. Unrestricted minimization

This trial state has some very interesting and unusualproperties. Apart from the overall Gaussian envelope,the spacing of the vortices completely determines thespatial variation of the number density nLLL�r�= ��LLL�r��2= �f����2e−r2/d�

2. The number density varies

smoothly between adjacent vortices because �f����2 con-sists of harmonic functions that obey Laplace’s equation.Hence the vortex core size is comparable with half theintervortex spacing l=�� /M�, which is essentially d� inthe rapidly rotating limit. In contrast to the Thomas-Fermi limit at lower angular velocity, the present trialfunction �LLL automatically incorporates all the kineticenergy.

It is important to emphasize that the one-body state�LLL arises from the GP energy functional and that thecorresponding many-body state

�GP�r1,r2, . . . ,rN� � �n=1

N

�LLL�rn� �5.18�

is just a Hartree product of these LLL states. For anysuch ground state, the system exhibits “off-diagonal”long-range order �Penrose and Onsager, 1956; Yang,1962�, and �LLL is the corresponding macroscopically oc-cupied condensate wave function. Thus the mean-fieldLLL regime still has a BEC and, in general, exhibitssuperfluidity.

In the limit of rapid rotation, this LLL wave functioncan serve as a possible variational trial solution for theGP energy functional. Let �¯�LLL denote an expectationvalue evaluated with �LLL. Equations �5.7� and �5.12�show that �Lz�LLL/�= �a+

†a+�LLL because �a−†a−�LLL van-

ishes for any LLL state. Furthermore, the quantity r2

=x2+y2 is easily expressed as �x+ iy��x− iy�=d�2 �a−

+a+†��a++a−

†�, whose expectation value gives �r2�LLL/d�2

= �a+†a+�LLL+1. Comparison yields the remarkably simple

LLL relation between the angular momentum and themean-square radius �Ho, 2001; Watanabe et al., 2004;Aftalion et al., 2005�

680 Alexander L. Fetter: Rotating trapped Bose-Einstein condensates

Rev. Mod. Phys., Vol. 81, No. 2, April–June 2009

Page 35: Rotating trapped Bose-Einstein condensatespjh/teaching/phz7429/fetter.pdfII. Physics of Bose-Einstein Condensates in Dilute Trapped Gases 648 A. Ideal Bose gas 649 1. Ideal BEC in

�Lz�LLL = M���r2�LLL − N� �5.19�

for any linear combination of single-particle states in thelowest Landau level �as discussed in Sec. III.C, a slightlydifferent form of this relation is important in the analysisof Butts and Rokhsar �1999��. Note that this relationalso follows directly from Eq. �5.13�.

A combination of Eqs. �5.2� and �5.19� expresses thefull GP energy functional �5.1� in the very intuitive andfamiliar form,

E���LLL� = N�� +� d2r�M��2 �1 − ��r2��LLL�2

+12

g2D��LLL�4� , �5.20�

where the integrand again involves only r2��LLL�2 and��LLL�4 �compare Eq. �2.26��. Unrestricted minimizationwith respect to ��LLL�2 at fixed normalization readilyyields the number density

��min�r��2 = nmin�r� = nmin�0��1 −r2

R02 , �5.21�

where the �two-dimensional� central density is nmin�0�=min/g2D, and the condensate radius is given by

R02 =

min

M��2 �1 − ��

. �5.22�

Remarkably, this expression has the same parabolicform as the usual TF result in Eq. �2.29�, even thoughthe present functional explicitly includes all the kineticenergy. It does, however, rely on the restriction to thelowest Landau level.

The normalization condition �d2r��min�2=N yields N= 1

2�R02nmin�0�, which is equivalent to

R02 =

2Ng2D

�min. �5.23�

The product of Eqs. �5.22� and �5.23� gives an explicitexpression for R0

4,

R04 =

2Ng2D

�M��2 �1 − ��

=8Nad�

4

Z�1 − ��, �5.24�

that explicitly shows the radial expansion for �→1.Similarly, the ratio of the same two equations yields thechemical potential

min =�2Ng2DM��2 �1 − ��

�=�8Na�1 − ��

Z���.

�5.25�

The first form is similar to Eq. �4.9� in the mean-field TFregime �the different fractional powers arise from thedifferent dimensions for the normalization integrals�,whereas the second form gives the condition min"2��� for the validity of the mean-field LLL regime

1 − � "Z

2Na. �5.26�

If Z /2a�100 and N�104, then a dimensionless rotation

speed ��0.99 just barely approaches the mean-fieldLLL regime.

The energy of this approximate minimizing mean-fieldLLL density ��min�2 follows immediately by evaluatingEq. �5.20�,

Emin − N��

N���

=2min

3���

=43�2Na�1 − ��

Z. �5.27�

As emphasized by Aftalion et al. �2005� the approximatedensity in Eq. �5.21� arises from an unrestricted varia-tion and is not within the LLL. Thus the actual energywill be higher because the minimization then includesthe vortices through the zeros of the analytic function inEq. �5.17�. As seen below, the main effect of the vorticesis to increase the numerical coefficient in Eq. �5.27�,keeping the same dependence on the various param-eters.

The derivative of Eq. �5.27� with respect to � yieldsthe angular momentum of the minimizing mean-fieldLLL state

�Lz�min = 13MNR0

2�� − N� , �5.28�

which also follows from Eq. �5.19�. The first term is justthe solid-body value �since direct integration with theparabolic ��min�2 yields �r2�min= 1

3NR02�. Here the LLL re-

duction from the classical value is � per particle, which issmaller than that in the mean-field TF regime �see Eq.�4.16� in Sec. IV.A�.

2. Inclusion of the vortices

The minimizing parabolic density profile in Eq. �5.21�cannot provide a full description of the rotating LLLcondensate, for it omits the fine-grained structure asso-ciated with the vortices.

a. Renormalization of g2D

In the present mean-field LLL regime, the intervortexseparation �d� is small compared to the size of the con-densate. Thus it is possible to write the number densityobtained from Eq. �5.16� in the form ��LLL�r��2

���LLL�r��2�m�r��2 as the product of a slowly varying en-

velope function ��LLL�r��2 and a rapidly varying modula-tion function �m�r��2 that vanishes at the center of eachvortex �Fetter, 2001; Fischer and Baym, 2003; Baym andPethick, 2004; Watanabe et al., 2004; Aftalion et al.,2005�. For a large vortex lattice, this modulating func-tion is essentially periodic, and it is convenient to nor-malize it so that �celld

2r�m�r��2=1 over each unit cell.Substitution of this approximate density into the LLLenergy functional �5.20� yields products of slowly vary-

ing functions involving powers of ��LLL�r��2 and rapidlyvarying periodic modulating functions involving powersof �m�r��2.

681Alexander L. Fetter: Rotating trapped Bose-Einstein condensates

Rev. Mod. Phys., Vol. 81, No. 2, April–June 2009

Page 36: Rotating trapped Bose-Einstein condensatespjh/teaching/phz7429/fetter.pdfII. Physics of Bose-Einstein Condensates in Dilute Trapped Gases 648 A. Ideal Bose gas 649 1. Ideal BEC in

The integral over the condensate separates into a sumof integrals over each unit cell, with the slowly varyingquantities acting as locally constant factors. Thus theonly effect of the vortex lattice is that the interactionterm acquires a numerical factor �=�celld

2r�m�r��4, whicheffectively renormalizes the interaction constant g2D→�g2D �Aftalion et al. �2005� discussed these steps�. Foran unbounded triangular vortex lattice, the resulting ��1.1596 is the numerical value for the correspondingAbrikosov vortex lattice �Abrikosov, 1957; Kleiner et al.,1964�. Consequently, the radius R0 in Eq. �5.24� expandsby the factor �1/4�1.0377.

b. Nonuniformity of the vortex array

To study the effect of the vortices in more detail, Ho�2001� considered the logarithm of the particle densityfor a LLL condensate wave function of the form in Eq.�5.16�, obtaining

ln ��LLL�r��2 = ln nLLL�r� = −r2

d�2 + 2

jln�r − rj� ,

�5.29�

where rj is the position of the jth vortex. The two-dimensional Laplacian of this equation yields

�2 ln nLLL�r� = −4

d�2 + 4�

j��2��r − rj� �5.30�

because �2 ln�r−rj�=2���2��r−rj�. The sum over the deltafunctions is precisely the two-dimensional vortex densitynv�r�, which gives a striking relation between the particledensity nLLL�r� in the mean-field LLL regime and thevortex density �Ho, 2001; Watanabe et al., 2004; Aftalionet al., 2005�

nv�r� =1

�d�2 +

1

4��2 ln nLLL�r� . �5.31�

The first term is an effective vortex density neff=M�� /�� for �=��, and the second is a �small� correc-tion.

To appreciate the implications of this elegant equa-tion, first assume that the vortex density is uniformwith the actual Feynman value �nv=m� /�. Equation�5.31� then has the Gaussian solution nLLL�r�=nLLL�0�exp�−r2 /'2�, where '−2=��neff−nv�; equiva-lently, the effective squared condensate radius is

'2 =�

M��� − ��=

d�2

1 − �. �5.32�

As expected from the radial expansion, the length ' di-verges as �→��.

In fact, Cooper et al. �2004� used numerical methodsto show that a small distortion of the vortex lattice canboth lower the energy and significantly change the meandensity profile from Gaussian to one closely resemblingthe parabolic shape in Eq. �5.21�. Thus it is better to

substitute this profile into Eq. �5.31� to find the approxi-mate and slightly nonuniform vortex density �Watanabeet al., 2004; Aftalion et al., 2005�

nv�r� �1

�d�2 −

1

�R02

1

�1 − r2/R02�2 . �5.33�

This expression is similar to the result in the mean-fieldThomas-Fermi limit in Eq. �4.24�, and the correction isagain small, of order 1/Nv.

3. Rapidly rotating anisotropic traps

In practice, most traps are not exactly axisymmetric�namely, �x��y, and I take �x �y for definiteness�.The rotational properties of such realistic traps differ incertain significant ways from those of symmetric traps.Similar problems arise in the study of deformed nuclei�Valatin, 1956�. If the anistropic trap is stationary, thenoninteracting single-particle ground state is simply aproduct of two one-dimensional Gaussians in x and ywith oscillator lengths dx and dy. With increasing rota-tion speed �, however, the coupling term −�Lz mixesthe x and y components. As a result, the ground-statedensity elongates along the direction of weak confine-ment �here along x� and diverges as �→�x �Linn et al.,2001�.

In complete analogy with the operators a± and a±† for

the symmetric trap �see Eq. �5.5��, a rotating anisotropictrap has a similar set of creation and destruction opera-tors, but their form is rather more intricate �Valatin,1956; Oktel, 2004; Fetter, 2007�. Nevertheless, in thelimit �→�x, the lowest single-particle states constitute anearly degenerate set that acts like the lowest Landaulevel. The corresponding LLL condensate wave functionagain involves a polynomial in a scaled complex variable�a linear combination of x and iy�. Consequently, manyof the results derived for the mean-field LLL regime inan isotropic trap continue to apply, at least qualitatively,for the rapidly rotating anisotropic trap.

One important difference is the significant anisotropyof the condensate as �→�x �Linn et al., 2001; Oktel,2004; Fetter, 2007�. Indeed, Sánchez-Lotero and Palacios�2005� and Sinha and Shlyapnikov �2005� both modeledthis situation with a one-dimensional effectively infinitestrip of finite width, containing one or more rows of vor-tices. Using the more detailed model of a rapidly rotat-ing anisotropic harmonic trap, Aftalion et al. �2009� pre-dicted a “fast-rotating” regime when the elongatedquasi-one-dimensional condensate has no visible vorti-ces. As discussed in Sec. VI, however, this behavior maywell be unobservable. Before reaching such a one-dimensional limit, there may be a quantum phase tran-sition to a correlated many-body state �sometimes calleda “vortex liquid”� that is not superfluid and has no BEC.Such questions currently remain unanswered and meritmore study.

682 Alexander L. Fetter: Rotating trapped Bose-Einstein condensates

Rev. Mod. Phys., Vol. 81, No. 2, April–June 2009

Page 37: Rotating trapped Bose-Einstein condensatespjh/teaching/phz7429/fetter.pdfII. Physics of Bose-Einstein Condensates in Dilute Trapped Gases 648 A. Ideal Bose gas 649 1. Ideal BEC in

VI. QUANTUM PHASE TRANSITION TO HIGHLYCORRELATED STATES

The mean-field LLL regime relies on the assumptionthat the interaction energy per particle � is small com-

pared to the gap 2���, leading to the lower bound for ��compare Eq. �5.26��,

1 −Z

2Na" � , �6.1�

where a=�a is the renormalized scattering length thatincludes the vortex-induced density variation �note that��1.1596 is simply a numerical factor, see Sec. V.C.2 formore details�. In practice, typical experiments require

��0.99 to reach this regime. For the present purpose, itis important to emphasize that the mean-field LLL re-gime has a BEC with macroscopic occupation and is in-deed a superfluid. Specifically, the associated many-bodyGP ground state is simply the Hartree product in Eq.�5.18�.

For still larger values of �, however, the mean-fieldLLL regime should eventually disappear through aquantum phase transition, leading to a wholly differenthighly correlated many-body ground state �corr. Unlikethe mean-field LLL ground state, the new ground state�corr does not have a BEC and is not a superfluid. Asevidence for such a transition, the exact two-dimensional ground state for bosons with small N andrelatively large fixed angular momentum Lz indeed hasa correlated form �Wilkin et al., 1998; Wilkin and Gunn,2000�. Furthermore, detailed numerical studies �Cooperet al., 2001� explicitly exhibited such a phase transition�for a recent review, see Viefers �2008��.

To understand the physics of this quantum phase tran-sition, it is helpful to focus on the number of vortices Nv.In the mean-field LLL regime, the two-dimensional con-densate has a radius R0, and Eq. �4.6� yields the numberof vortices Nv=R0

2 / l2�R02 /d�

2 . Use of Eq. �5.24� and thediscussion in Sec. V.C.2 give

Nv �R0

2

d�2 =� 8Na

Z�1 − ��. �6.2�

As an alternative derivation of this result �Bloch et al.,2008�, note that the mth LLL state �m has a mean-square radius �md�

2 for m1 �Eq. �5.13��. If the sum inEq. �5.16� extends to mmax, then the condensate has aneffective squared radius �r2�mmax

�mmaxd�2 . The number

of vortices �the degree of the polynomial� is Nv=mmax,which is the same as Eq. �6.2� found directly from theFeynman relation for the vortex density. The parametersused below Eq. �5.26� yield Nv�200.

It is valuable to consider the ratio (�N /Nv �Wilkinand Gunn, 2000; Cooper et al., 2001� known as the “fill-ing factor” or “filling fraction” because of a similar

quantity in the fractional quantum Hall effect �see, forexample, Prange and Girvin �1987��. In the mean-fieldLLL regime, it has the simple form

( �N

Nv=�Z�1 − ��N

8a, �6.3�

but it also remains well defined for general many-bodyground states of rotating bosons �and for charged fermi-ons in a magnetic field, which is its original role �Prangeand Girvin, 1987��.

In the mean-field LLL approach for weakly interact-ing bosons in a rotating trap, typified by the work ofButts and Rokhsar �1999� and discussed in Secs. III.Cand V.C, the equilibrium state is a vortex array thatbreaks rotational symmetry and is not an eigenstate ofLz. As an alternative, Cooper et al. �2001� used exactdiagonalization to study the ground states for Nv=4, 6,and 8 and for many values of N �see also Cooper andWilkin �1999�; Viefers et al. �2000�; Wilkin and Gunn�2000� for less extensive studies�. This numerical analysisused a toroidal geometry with periodic boundary condi-tions in both directions, allowing a study of bosonic sys-tems with relatively large integral values of both N andNv. For a particular aspect ratio of the torus that canaccommodate a triangular array with Nv=8, they inves-tigated filling fractions ( ranging through several ratio-nal values from 1

2 up to 9. They also studied the mean-field GP ground state containing a vortex lattice withNv=8 for the same values of (. Comparison of these twoclasses of states shows that the GP vortex lattice is theground state for (c�6, where the specific numericalvalue depends on the aspect ratio of the torus �andhence the number of vortices that can fit into the peri-odic structure�. Other choices of this aspect ratio yieldcritical values in the range (c�6–10.

A combination of Eq. �6.1� for the validity of themean-field LLL regime and the lower bound (�1 forthe mean-field GP ground state yields the inequalitiesthat define the allowed range of the mean-field LLLstate �Bloch et al., 2008�

1 −Z

2Na" � " 1 −

8a

ZN. �6.4�

The right-hand inequality for � illustrates the experi-mental difficulty of studying the quantum phase transi-tion. Since Z /a typically exceeds 100, Eq. �6.3� makes itclear that reaching even (�10 will require significant

reduction of N �note that the high rotation speed ��0.999 presents a daunting experimental challenge�.

In contrast to the vortex lattices for (�(c, the groundstates for smaller ( (c are rotationally symmetric in-compressible vortex liquids that are eigenstates of Lz.They have close similarities to the bosonic analogs of theJain �1989� sequence of fractional quantum Hall statesthat describe a two-dimensional electron gas in a strongmagnetic field. The simplest of these many-body groundstates is the bosonic Laughlin state �Laughlin, 1983�

683Alexander L. Fetter: Rotating trapped Bose-Einstein condensates

Rev. Mod. Phys., Vol. 81, No. 2, April–June 2009

Page 38: Rotating trapped Bose-Einstein condensatespjh/teaching/phz7429/fetter.pdfII. Physics of Bose-Einstein Condensates in Dilute Trapped Gases 648 A. Ideal Bose gas 649 1. Ideal BEC in

�Laughlin�r1,r2, . . . ,rN� � �n n�

N

�zn − zn��2

�exp�−12

n=1

N

�zn�2 , �6.5�

where zn= �xn+ iyn� /d� refers to the nth particle. Herethe power 2 ensures that this state is symmetric underthe interchange of any two particles �in contrast to thefermionic case with power 3�. Wilkin et al. �1998�; Coo-per and Wilkin �1999�; and Wilkin and Gunn �2000� notethat this state occurs for (= 1

2 .It is worth emphasizing the difference between this

Laughlin ground state and the GP Hartree ground statein Eq. �5.18�. For the Laughlin state in Eq. �6.5�, the first�product� factor means that there is no off-diagonallong-range order �Penrose and Onsager, 1956; Yang,1962� and hence no BEC. In addition, the Laughlin statevanishes whenever two particles come together, enforc-ing the many-body correlations. Indeed, the usual short-range two-body potential V�r−r��=g��r−r�� has zero ex-pectation value in this correlated state �Trugman andKivelson, 1985�. In contrast, the GP ground state hasevery particle in the same single-particle Hartree state�LLL; thus particles in the GP condensate tend to over-lap wherever ��LLL�r�� is large.

Sinova et al. �2002� provided an appealing physicalpicture of the onset of this quantum phase transition.Using various approximate descriptions, they found thatthe long-wavelength spectrum of collective modes in thevortex lattice becomes quadratic in the limit of rapidrotations, in contrast to the familiar linear Bogoliubovspectrum in Eq. �2.46�. This softened spectrum impliesthat the vortex lattice melts for a critical filling fraction(c�8, in reasonable agreement with the estimates ofCooper et al. �2001�. An alternative but essentiallyequivalent description of the vortex-lattice melting�Baym, 2005� relies on the softening of the Tkachenkovortex-lattice modes, as seen in Eq. �4.28�.

The periodic toroidal geometry of Cooper et al. �2001�has the strong advantage of avoiding boundary effects;Regnault and Jolicoeur �2003, 2004� used a spherical ge-ometry to perform a similar numerical analysis. Bothgroups proposed specific sequences of known quantumHall states for the rotating dilute Bose gas when 1

2 ( (c, although there is currently no experimental evi-dence for such sequences. Nevertheless, Cooper et al.�2001� found strong overlap between their results fromexact diagonalization and the Laughlin state �(= 1

2 � andother similar but more complicated quantum Hall statesfor the particular filling fractions (=1, 3

2 ,2 , 52 ,3. In addi-

tion, Cooper and Rezayi �2007� found stripe phases at afilling factor (=2. For a more detailed discussion ofthese various quantum Hall states, see, for example,Cooper et al. �2001�, Regnault and Jolicoeur �2003,2004�, Bloch et al. �2008�, and Viefers �2008�.

What is the physics of this ground-state transitionfrom a vortex lattice to a vortex liquid? Why is the criti-

cal value of the filling fraction (c�5–10, as opposed to(c�1 or �0.1? To think about this question, it is helpfulto consider N bosonic particles in a plane, with 2N de-grees of freedom. Vortices appear as the system rotates,and the corresponding vortex coordinates provide Nvcollective degrees of freedom.13 This relation betweenparticle and collective degrees of freedom is familiar invarious areas of many-body theory: For example, theDebye theory of specific heat uses the phonon normalmodes instead of the particle positions in the crystal, andthe original number of particle degrees of freedom �3N�also determines the cutoff frequency for the phononspectrum �see, for example, Ashcroft and Mermin�1976��. Similarly, the theory of a charged electron gasoften relies on collective density modes �plasma oscilla-tions� instead of the particle coordinates �Pines, 1961�.

For a slowly rotating Bose-Einstein gas in a plane, the2N particle coordinates provide a convenient descrip-tion. In principle, the Nv collective vortex degrees offreedom should reduce the original total 2N degrees offreedom to 2N−Nv, but the effect is unimportant as longas Nv�N. When Nv becomes comparable with N, how-ever, the depletion of the particle degrees of freedombecomes crucial, ultimately producing a phase transitionto a wholly different ground state. With this picture, thecritical value (c�5–10 is not surprising, since Nv /N=1/( is still relatively small.

If experiments can indeed create these correlatedstates, how might they be detected? Bloch et al. �2008�proposed several techniques, including �1� a reducedthree-body recombination rate �because of the anticor-relation between the bosons inherent in the first factorin Eq. �6.5�� and �2� the characteristic density profile ofthese various quantum Hall states. Another possibilitywould be to study the ensemble-averaged density-density noise correlations �Brown and Twiss, 1956; Readand Cooper, 2003; Altman et al., 2004�, which shouldexhibit a different structure in the correlated state fromthat in the mean-field LLL regime.

VII. OUTLOOK

As this review demonstrates, experimental and theo-retical studies of rotating Bose-Einstein condensates

generally agree well for scaled angular velocities ��� /�� up to �0.995.

A. Successes

These accomplishments include several different re-gimes.

�1� For slow rotation ��� /���1 �Sec. III�, the con-densate includes only a few vortices, and centrifugal

13Note that vortices obey first-order dynamical equations, sothat the associated x and y coordinates are not independent; inparticular, they can serve as a pair of Hamiltonian canonicalvariables �Fetter, 1967; Haldane and Wu, 1985�.

684 Alexander L. Fetter: Rotating trapped Bose-Einstein condensates

Rev. Mod. Phys., Vol. 81, No. 2, April–June 2009

Page 39: Rotating trapped Bose-Einstein condensatespjh/teaching/phz7429/fetter.pdfII. Physics of Bose-Einstein Condensates in Dilute Trapped Gases 648 A. Ideal Bose gas 649 1. Ideal BEC in

effects are negligible. Thus the overall density pro-file is essentially that of the nonrotating condensate,apart from the small holes associated with the vor-tex cores. Experimental observations of vortex pre-cession in one- and two-component superfluids �seeSec. III.D.1� provide the only direct measurementsof vortex dynamics in dilute BECs.

�2� As the rotation increases, the centrifugal effects be-come important, leading to significant changes in thecondensate’s aspect ratio �Sec. IV�. For not too large

�, however, the intervortex spacing �l= �� /M��1/2

remains large compared to the vortex core size �. Inthis mean-field Thomas-Fermi regime, the gradientenergy associated with density variation remainsnegligible, and the hydrodynamic flow velocity v=��S /M predominates in the energy balance. Typi-

cally this case holds for 0.75"�"0.99.

�3� Finally, for very large rotation speed 0.99"�"0.999, the vortex cores expand, filling much of theavailable space. The energy associated with spatialvariation of the density now becomes comparable tothe hydrodynamic flow energy. In this mean-fieldlowest-Landau-level regime �Sec. V�, the single-particle part of the energy in the rotating frame isexactly soluble, leading to the rotational analog ofthe Landau levels for an electron in a uniform mag-netic field. In particular, there is an energy gap�2��� between the nearly degenerate lowest Lan-dau level and the first excited Landau level. In ad-dition, the condensate expands radially, decreasingboth the central particle density and the mean inter-action energy gn�0�, which thus becomes smallerthan the energy gap 2���. The set of lowest-Landau-level states then provides a convenient andflexible description of the equilibrium of the rapidlyrotating condensate. Although current experimentswith / �2�����0.6 only just reach this mean-fieldLLL regime �see Sec. V.A�, it is reasonably wellunderstood.

B. Challenges

Despite the many remarkable achievements of thepast decade, several theoretical predictions of consider-able interest remain unverified experimentally.

1. Rapid rotation of both Bose and Fermi gases in the quantumHall regime at small filling fraction

For dilute ultracold trapped single-component Bosegases of the sort considered in this review, one obviouschallenge for future experiments is the quantum-phasetransition from a superfluid with a Bose-Einstein con-densate to a nonsuperfluid correlated state �see, for ex-ample, Viefers �2008��. As discussed in Sec. VI, forma-tion of this new state requires large angular momentumand smaller particle number N than is typical in mostcurrent experiments. Whether such states can indeed be

realized is uncertain. In principle, one might start withan anharmonic trap of the sort considered in Sec. IV.C.

Once the condensate rotates with ��1, it might be fea-sible adiabatically to turn off the external laser beam�and hence the anharmonic component of the trap po-tential�, raising the effective �� and leaving the system

with high angular velocity �"1. Alternatively, numeri-cal studies �Morris and Feder, 2007� for small N suggestthat a repulsive Gaussian beam itself may facilitate for-mation of correlated quantum Hall states.

Another possibility is to rely on topological gauge po-tentials �Sec. V.B�, in which applied laser fields with or-bital angular momentum can produce effective nonuni-form topological “magnetic” fields that are more generalthan the usual constant B. Essentially all studies of rap-idly rotating trapped bosons rely on the detailed form ofthe operator −� ·L=−A ·p, with A=��r as the equiva-lent vector potential �but note recent work �Murray etal., 2007� that introduces more general topological vec-tor potentials�. The existence of a nonsuperfluid corre-lated many-body state reflects the high degeneracy ofthe lowest Landau level, which in turn depends on thespecific form of the effective vector potential �A=��r,apart from a possible gauge transformation�. It is notobvious that such nonsuperfluid correlated many-bodystates occur for more general topological gauge poten-tials. This question certainly merits additional detailedstudy.

A related area for future activity is the study of rap-idly rotating Fermi gases. At present, this theoreticaltopic is only partially explored �see, for example, Zhaiand Ho �2006�; Bhongale et al. �2007�; Möller and Coo-per �2007��, and much work remains. The only relevantexperiments are those of Zwierlein et al. �2005� on thevortex lattice in rotating fermionic 6Li �see Sec. IV.B.6�.The experiment studies both a BEC of tightly boundbosonic molecules and the more weakly bound fermi-onic regime, including the BEC-BCS crossover regimenear the Feshbach resonance. Unfortunately, any movefar to the BCS side of the transition at fixed low tem-perature necessarily encounters the severely reducedcritical temperature of the weak-coupling BCS limit.

As a pure theoretical argument, consider the weak-

coupling fermionic ground state as a function of �

=� /��. For intermediate �, a vortex lattice forms, as in

Zwierlein et al. �2005�. What happens for larger �? Isthere a transition to a normal state similar to that at Hc2in a type-II superconductor �see, for example, Tinkham�1996��? Or does a transition to a correlated quantumHall state intervene? If so, for reasonable experimental

parameters, how does the filling fraction ( relate to �?These intriguing questions deserve careful study.

2. Rotation of ultracold dilute gases with long-range dipole-dipole interactions

Dilute quantum gases with long-range anisotropicdipole-dipole interactions constitute a fascinating gener-

685Alexander L. Fetter: Rotating trapped Bose-Einstein condensates

Rev. Mod. Phys., Vol. 81, No. 2, April–June 2009

Page 40: Rotating trapped Bose-Einstein condensatespjh/teaching/phz7429/fetter.pdfII. Physics of Bose-Einstein Condensates in Dilute Trapped Gases 648 A. Ideal Bose gas 649 1. Ideal BEC in

alization of the usual case with short-range isotropic in-teractions characterized by a scattering length a. Al-though the general dipole-dipole potential is rathercomplicated, experiments usually use a strong uniformfield to orient the dipoles �either magnetic or electric�along the z direction. Such fields lead to the simpler in-teraction potential �see, for example, the review byMenotti and Lewenstein �2008� and the experimentalstudies of 52Cr �Griesmaier et al., 2005; Lahaye et al.,2007; Koch et al., 2008��

Vdd�r − r�� =d2

�r − r��3�1 − 3�z · n�2� , �7.1�

where d is the relevant dipole moment and n is a unitvector along the direction r−r� joining the two atoms.

The dipole-dipole potential in Eq. �7.1� is repulsive ifz · n vanishes �namely, if the oriented dipoles lie in thex-y plane�, whereas it is attractive if �z · n�2�

13 �in par-

ticular, for oriented dipoles located along the z axis�.Since a BEC with attractive interactions tends to col-lapse, a long cigar-shaped condensate with dipole-dipoleinteractions should be unstable, but a flat disk-shapedcondensate should be stable. Koch et al. �2008� used aFeshbach resonance to decrease the s-wave scatteringlength, thus enhancing the relative role of Vdd. Theseexperiments confirmed the qualitative expectations ingreat detail.

Recent theoretical studies have predicted many prop-erties of the ground states of rotating BECs with dipolarinteractions. In the weak-interaction limit, Cooper et al.�2005� �see also Zhang and Zhai �2005�� proposed thatan increased dipolar coupling strength induces a seriesof transitions in the mean-field �large filling fraction (�condensate involving vortex lattices of varying symme-try �triangular, square, stripe, and “bubble” phases�.They also studied the behavior for small (, when thesystem is no longer superfluid. Subsequently, Komineasand Cooper �2007� analyzed the more general situationas a function of both the dipolar interaction and thechemical potential relative to the s-wave contact interac-tion.

Another system of great interest is the rapidly rotat-ing dipolar Fermi gas, whose predicted ground state�Baranov et al., 2005� has a simple incompressible formfor filling fraction (= 1

3 . For smaller (, however, theorypredicts that the quantum Hall states ultimately becomeunstable to the formation of crystalline Wigner states�Osterloh et al., 2007�. Experimental studies of such non-superfluid fermionic states involve many technical issuesthat remain for future investigation.

3. Rotating spinor condensates

The development of optical laser traps that confine allthe magnetic substates �Stamper-Kurn et al., 1998� al-lows a direct study of spin-one BECs �as well as higher-spin cases�. These experiments use a focused �red-detuned� infrared laser that attracts the BEC to thewaist of the beam. For a general theoretical introduc-

tion, see Sec. II.B, along with Fetter �2002�, Pethick andSmith �2002�, and the original papers by Ho �1998� andOhmi and Machida �1998�. In contrast to the two-component mixtures of the different hyperfine states F=1 and F=2 �see Secs. III.D.1 and IV.B.5�, the rotationalinvariance of the interaction between two atoms withthe same F=1 now imposes special restrictions on theallowed interaction potential. The macroscopic orderparameter becomes a spin-one object with three compo-nents,

� = � �1

�0

�−1� , �7.2�

labeled by the mF value. In the low-energy limit whereonly s-wave scattering is relevant, the effective short-range interaction has the form

Vint�r1 − r2� = �g0 + g2F1 · F2���3��r1 − r2� , �7.3�

where Fj is the hyperfine spin of the jth atom. The cou-pling constant g0 is generally positive, but g2 is propor-tional to a2−a0 �the difference of the scattering lengthsin the two symmetric channels F=F1+F2=0 or 2�. Thusg2 can be either positive or negative, depending on thedetails of the two-body scattering.

A mean-field description �Ho, 1998� used an effectiveenergy functional and wrote the spinor order parameterin a factored form ���r�=�n�r����r�. Here � runs overthe three allowed values �= �1,0 ,−1�, n�r� is the as-sumed common density for all three spin components,and ���r� is a normalized spinor with �† ·�=1. Apart fromthe gradient term in the effective energy, the ground-state spinor follows by minimizing the spin-dependentpart of the energy density 1

2n2g2�F�2, where �F�=����

*F����.If g2 is positive, then the minimum energy occurs for

��F��=0. Such states are known as “polar” or “antiferro-magnetic,” and they follow by spatial rotations of thehyperfine state �mF=0� �this situation applies for 23Na�Stenger et al., 1998��. If, however, g2 is negative, thenthe minimum energy is for a “ferromagnetic” state with��F��=1, obtained by spatial rotations of the state �mF

=1� �this situation applies for 87Rb �Sadler et al., 2006;Vengalattore et al., 2008��.

The multicomponent structure of the spinor order pa-rameter in Eq. �7.2� is reminiscent of the order param-eter for superfluid 3He, where each atom has a nuclearspin 1/2 and zero net electronic spin. Low-temperatureexperiments show that 3He forms a p-wave paired super-fluid, with the two atoms in a triplet nuclear spin stateand unit relative orbital angular momentum �see, for ex-ample, Vollhardt and Wölfle �1990��. In particular, the3He order parameter has several components, analogousto those for a spinor BEC. Many groups have studiedthe structure of vortices in rotating 3He, with variousproposed candidates, both for a single vortex and forvortex lattices �for detailed reviews, see Fetter �1986�;Salomaa and Volovik �1987��.

686 Alexander L. Fetter: Rotating trapped Bose-Einstein condensates

Rev. Mod. Phys., Vol. 81, No. 2, April–June 2009

Page 41: Rotating trapped Bose-Einstein condensatespjh/teaching/phz7429/fetter.pdfII. Physics of Bose-Einstein Condensates in Dilute Trapped Gases 648 A. Ideal Bose gas 649 1. Ideal BEC in

Unfortunately, the semiphenomenological Ginzburg-Landau theory for superfluid 3He is often inadequate topredict the detailed experimental phase diagram forvarious vortices as a function of pressure and tempera-ture. Specifically, the energy difference between many ofthe vortex states is small, and the relevant phenomeno-logical Ginzburg-Landau parameters have considerableuncertainties in their experimental values �see, for ex-ample, Vollhardt and Wölfle �1990��. Consequently, ex-perimental studies provide essential guidance for theo-retical analyses. Eltsov et al. �2005� and Finne et al.�2006� reviewed recent experimental work on vorticesand other related structures in superfuid 3He.

In an unbounded dilute trapped BEC with a single-component order parameter, the vortex core size andstructure arise from the balance of the kinetic energyand the interaction energy �see Sec. III.A�. For atrapped BEC with a single component, the new featureis the additional energy of the confining trap, and vari-ous regimes occur depending on the dimensionless inter-action parameter Na /d0 �see Secs. II.C and III.B.1�. IfNa /d0"1, then the vortex core is comparable with thecondensate radius �both of order d0, the mean oscillatorlength� or the intervortex separation �see Sec. III.C�. Inthe more typical Thomas-Fermi regime with Na /d01,however, the vortex core radius � �the healing length� ismuch smaller than both d0 and the still larger mean con-densate radius R0.

For a three-component spinor condensate, the spatialvariation of each component provides additional free-dom in minimizing the overall free energy. Since the re-sulting coupled GP field equations are nonlinear, severalsolutions can exist, at least in principle, each corre-sponding to a local minimum of the GP energy �see Voll-hardt and Wölfle �1990� for the similar situation involv-ing the Ginzburg-Landau theory of superfluid 3He�. Inpractice, it turns out that �g2��g0 for both 23Na and 87Rb�Yip, 1999�, which means that gradient energies andother interaction energies play a significant role in deter-mining the condensate wave function. This near degen-eracy is especially important in understanding the coreof a quantized vortex in a spinor condensate becausethese various energies compete with the energy of thecirculating flow. Yip �1999� proposed several differentvortex structures, and many others have amplified theseconsiderations, both for single vortices �see, for ex-ample, Isoshima et al. �2001�; Mizushima et al. �2002�;Bulgakov and Sadreev �2003� and references therein�and for vortex lattices �see, for example, Mizushima etal. �2004�, and references therein�. One experimentalchallenge in studying a rotating spinor BEC is ensuringboth the full axisymmetry of the optical trap and itsalignment with the axis of rotation; otherwise the life-time decreases significantly. These difficulties appear inthe experiments on a rotating paired Fermi gas, wherethe magnetic field associated with a Feshbach resonancealso requires an optical trap �see Sec. IV.B.6 and Zwier-lein et al. �2005��.

ACKNOWLEDGMENTS

I thank A. Aftalion, N. Cooper, E. Cornell, J. Dali-bard, E. Demler, W. Ketterle, P. Krüger, M. Krusius, M.Möttönen, M. Oktel, R. Packard, D. Roberts, D. Rokh-sar, A. Sedrakian, S. Viefers, N. Wilkin, E. Zaremba,and M. Zwierlein for valuable discussions and sugges-tions, and for permission to reproduce various figures.Part of this work was completed during a stay at theInstitut Henri Poincaré-Centre Emile Borel, Paris, andat the Kavli Institute for Theoretical Physics, Santa Bar-bara, where the research was supported in part by theNational Science Foundation under Grant No. PHY05-51164. I thank these institutions for their hospitality andsupport.

REFERENCES

Abers, E. S., 2004, Quantum Mechanics �Pearson PrenticeHall, Upper Saddle River, NJ�.

Abo-Shaeer, J. R., C. Raman, and W. Ketterle, 2002, Phys.Rev. Lett. 88, 070409.

Abo-Shaeer, J. R., C. Raman, J. M. Vogels, and W. Ketterle,2001, Science 292, 476.

Abrikosov, A. A., 1957, Zh. Eksp. Teor. Fiz. 32, 1442 �Sov.Phys. JETP 5, 1174 �1957��.

Aftalion, A., 2006, Vortices in Bose-Einstein Condensates�Birkhäuser, Boston�.

Aftalion, A., X. Blanc, and J. Dalibard, 2005, Phys. Rev. A 71,023611.

Aftalion, A., X. Blanc, and N. Lerner, 2009, Phys. Rev. A 79,011603�R�.

Aftalion, A., and I. Danaila, 2004, Phys. Rev. A 69, 033608.Aftalion, A., and T. Rivière, 2001, Phys. Rev. A 64, 043611.Allen, L., S. M. Barnett, and M. J. Padgett, 2003, Eds., Optical

Angular Momentum �IOP, Bristol�.Allen, L., and J. H. Eberly, 1987, Optical Resonance and Two-

Level Atoms �Dover, New York�.Allen, L., M. J. Padgett, and M. Babikar, 1999, Prog. Opt. 39,

291.Altman, E., E. Demler, and M. D. Lukin, 2004, Phys. Rev. A

70, 013603.Andersen, M. F., C. Ryu, P. Cladé, V. Natarajan, A. Vaziri, K.

Helmerson, and W. D. Phillips, 2006, Phys. Rev. Lett. 97,170406.

Anderson, B. P., P. C. Haljan, C. A. Regal, D. L. Feder, L. A.Collins, C. W. Clark, and E. A. Cornell, 2001, Phys. Rev. Lett.86, 2926.

Anderson, B. P., P. C. Haljan, C. E. Wieman, and E. A. Cor-nell, 2000, Phys. Rev. Lett. 85, 2857.

Anderson, B. P., and M. A. Kasevich, 1998, Science 282, 1686.Anderson, M. H., J. R. Ensher, M. H. Matthews, C. E. Wie-

man, and E. A. Cornell, 1995, Science 269, 198.Andrews, M. R., C. G. Townsend, H.-J. Miesner, D. S. Durfee,

D. M. Kurn, and W. Ketterle, 1997, Science 275, 637.Anglin, J. R., 2002, Phys. Rev. A 65, 063611.Anglin, J. R., and M. Crescimanno, 2002, e-print arXiv:cond-

mat/0210063.Arimondo, E., 1996, Prog. Opt. 35, 257.Ashcroft, N. W., and N. D. Mermin, 1976, Solid State Physics

�Holt, Rinehart and Winston, New York�.Bagnato, V., and D. Kleppner, 1991, Phys. Rev. A 44, 7439.

687Alexander L. Fetter: Rotating trapped Bose-Einstein condensates

Rev. Mod. Phys., Vol. 81, No. 2, April–June 2009

Page 42: Rotating trapped Bose-Einstein condensatespjh/teaching/phz7429/fetter.pdfII. Physics of Bose-Einstein Condensates in Dilute Trapped Gases 648 A. Ideal Bose gas 649 1. Ideal BEC in

Baksmaty, L. O., S. J. Woo, S. Choi, and N. P. Bigelow, 2004,Phys. Rev. Lett. 92, 160405.

Baranov, M. H., K. Osterloh, and M. Lewenstein, 2005, Phys.Rev. Lett. 94, 070404.

Baym, G., 2003, Phys. Rev. Lett. 91, 110402.Baym, G., 2005, J. Low Temp. Phys. 138, 601.Baym, G., and C. J. Pethick, 1996, Phys. Rev. Lett. 76, 6.Baym, G., and C. J. Pethick, 2004, Phys. Rev. A 69, 043619.Berezinskii, V. L., 1971, Zh. Eksp. Teor. Fiz. 61 1144 �Sov.

Phys. JETP 34, 610 �1972��.Berry, M. V., 1984, Proc. R. Soc. London, Ser. A 392, 45.Bhongale, S. G., M. R. Goosen, and S. J. J. M. F. Kokkelmans,

2007, e-print arXiv:0710:5288.Bishop, D. J., and J. D. Reppy, 1978, Phys. Rev. Lett. 40, 1727.Bloch, I., J. Dalibard, and W. Zwerger, 2008, Rev. Mod. Phys.

80, 885.Bogoliubov, N. N., 1947, J. Phys. �USSR� 11, 23.Bradley, A. S., C. W. Gardiner, and M. J. Davis, 2007, e-print

arXiv:0712.3436v1.Bradley, C. C., C. A. Sackett, and R. G. Hulet, 1997, Phys. Rev.

Lett. 78, 985.Bradley, C. C., C. A. Sackett, J. J. Tollett, and R. G. Hulet,

1995, Phys. Rev. Lett. 75, 1687.Bretin, V., P. Rosenbusch, F. Chevy, G. V. Shlyapnikov, and J.

Dalibard, 2003, Phys. Rev. Lett. 90, 100403.Bretin, V., S. Stock, Y. Seurin, and J. Dalibard, 2004, Phys. Rev.

Lett. 92, 050403.Brown, L. M., 2000, Ed., Selected Papers of Richard Feynman

�World Scientific, Singapore�.Brown, R. H., and R. Q. Twiss, 1956, Nature �London� 177, 27.Bulgakov, E. N., and A. F. Sadreev, 2003, Phys. Rev. Lett. 90,

200401.Butts, D. A., and D. S. Rokhsar, 1999, Nature �London� 397,

327.Campbell, L. J., and R. M. Ziff, 1979, Phys. Rev. B 20, 1886.Castin, Y., 2007, in Ultra-Cold Fermi Gases, edited by M. In-

guscio, W. Ketterle, and C. Salomon �IOS, Amsterdam�.Castin, Y., and R. Dum, 1999, Eur. Phys. J. D 7, 399.Chandrasekharan, K., 1985, Elliptic Functions �Springer-

Verlag, Berlin�.Chevy, F., K. W. Madison, V. Bretin, and J. Dalibard, 2002, in

Trapped Particles and Fundamental Physics, edited by S. N.Atutov, R. Calabrese, and L. Moi �Kluwer, Dordrecht�, p.109.

Chevy, F., K. W. Madison, and J. Dalibard, 2000, Phys. Rev.Lett. 85, 2223.

Coddington, I., P. Engels, V. Schweikhard, and E. A. Cornell,2003, Phys. Rev. Lett. 91, 100402.

Coddington, I., P. C. Haljan, P. Engels, V. Schweikhard, S.Tung, and E. A. Cornell, 2004, Phys. Rev. A 70, 063607.

Cohen-Tannoudji, C., B. Diu, and F. Lalöe, 1977, QuantumMechanics �Wiley, New York�.

Cohen-Tannoudji, C., J. Dupont-Roc, and G. Grynberg, 1998,Atom-Photon Interactions: Basic Processes and Applications�Wiley, New York�.

Colson, W. B., and A. L. Fetter, 1978, J. Low Temp. Phys. 33,231.

Cooper, N. R., S. Komineas, and N. Read, 2004, Phys. Rev. A70, 033604.

Cooper, N. R., and E. H. Rezayi, 2007, Phys. Rev. A 75,013627.

Cooper, N. R., E. H. Rezayi, and S. H. Simon, 2005, Phys. Rev.Lett. 95, 200402.

Cooper, N. R., and N. K. Wilkin, 1999, Phys. Rev. B 60,16279�R�.

Cooper, N. R., N. K. Wilkin, and J. M. F. Gunn, 2001, Phys.Rev. Lett. 87, 120405.

Cornell, E. A., D. S. Hall, M. R. Matthews, and C. E. Weiman,1998, J. Low Temp. Phys. 113, 151.

Cozzini, M., and S. Stringari, 2003, Phys. Rev. A 67, 041602�R�.Cozzini, M., S. Stringari, and C. Tozzo, 2006, Phys. Rev. A 73,

023615.Dalfovo, F., S. Giorgini, L. Pitaevskii, and S. Stringari, 1999,

Rev. Mod. Phys. 71, 463.Dalfovo, F., and S. Stringari, 1996, Phys. Rev. A 53, 2477.Dalfovo, F., and S. Stringari, 2000, Phys. Rev. A 63, 011601�R�.Dalibard, J., 2007, private communication.Davis, K. B., M.-O. Mewes, M. R. Andrews, N. J. van Druten,

D. S. Durfee, D. M. Kurn, and W. Ketterle, 1995, Phys. Rev.Lett. 75, 3969.

DeConde, K., and R. E. Packard, 1975, Phys. Rev. Lett. 35,732.

de Gennes, P. G., 1966, Superconductivity of Metals and Alloys�Benjamin, New York�.

Dodd, R. J., K. Burnett, M. Edwards, and C. W. Clark, 1997,Phys. Rev. A 56, 587.

Donnelly, R. J., 1991, Quantized Vortices in Helium II �Cam-bridge University Press, Cambridge, England�.

Dutton, Z., M. Budde, C. Slowe, and L. V. Hau, 2001, Science293, 663.

Edwards, M., P. A. Ruprecht, K. Burnett, R. J. Dodd, and C.W. Clark, 1996, Phys. Rev. Lett. 77, 1671.

Eltsov, V. B., M. Krusius, and G. E. Volovik, 2005, Prog. LowTemp. Phys. 15, 1.

Engels, P., I. Coddington, P. C. Haljan, and E. A. Cornell, 2002,Phys. Rev. Lett. 89, 100403.

Engels, P., I. Coddington, P. C. Haljan, V. Schweikhard, and E.A. Cornell, 2003, Phys. Rev. Lett. 90, 170405.

Esry, B. D., C. H. Greene, J. P. Burke, and J. L. Bohn, 1997,Phys. Rev. Lett. 78, 3594.

Feder, D. L., and C. W. Clark, 2001, Phys. Rev. Lett. 87,190401.

Feder, D. L., M. S. Pindzola, L. A. Collins, B. I. Schneider, andC. W. Clark, 2000, Phys. Rev. A 62, 053606.

Feder, D. L., A. A. Svidzinsky, A. L. Fetter, and C. W. Clark,2001, Phys. Rev. Lett. 86, 564.

Fedichev, P. O., A. E. Muryshev, and G. V. Shlyapnikov, 1999,Phys. Rev. A 60, 3220.

Fetter, A. L., 1967, Phys. Rev. 162, 143.Fetter, A. L., 1969, in Lectures in Theoretical Physics, edited by

K. T. Mahanthappa and W. E. Britten �Gordon and Breach,New York�, Vol. XIB, p. 351.

Fetter, A. L., 1972, Ann. Phys. �N.Y.� 70, 67.Fetter, A. L., 1974, J. Low Temp. Phys. 16, 533.Fetter, A. L., 1986, Prog. Low Temp. Phys. 10, 1.Fetter, A. L., 1996, Phys. Rev. A 53, 4245.Fetter, A. L., 1998, J. Low Temp. Phys. 113, 189.Fetter, A. L., 2001, Phys. Rev. A 64, 063608.Fetter, A. L., 2002, J. Low Temp. Phys. 129, 263.Fetter, A. L., 2007, Phys. Rev. A 75, 013620.Fetter, A. L., and K. Harvey, 1971, Phys. Rev. A 4, 2305.Fetter, A. L., B. Jackson, and S. Stringari, 2005, Phys. Rev. A

71, 013605.Fetter, A. L., and A. A. Svidzinsky, 2001, J. Phys.: Condens.

Matter 13, R135.

688 Alexander L. Fetter: Rotating trapped Bose-Einstein condensates

Rev. Mod. Phys., Vol. 81, No. 2, April–June 2009

Page 43: Rotating trapped Bose-Einstein condensatespjh/teaching/phz7429/fetter.pdfII. Physics of Bose-Einstein Condensates in Dilute Trapped Gases 648 A. Ideal Bose gas 649 1. Ideal BEC in

Fetter, A. L., and J. D. Walecka, 2003, Quantum Theory ofMany-Particle Systems �Dover, New York�.

Fetter, A. L., and J. D. Walecka, 2006, Nonlinear Mechanics: ASupplement to Theoretical Mechanics of Particles and Con-tinua �Dover, New York�.

Feynman, R. P., 1955, Prog. Low Temp. Phys. 1, 17.Feynman, R. P., 1965, The Feynman Lectures on Physics:

Quantum Mechanics �Addison-Wesley, Reading, MA�, Vol.III.

Finne, A. P., V. B. Eltsov, R. Hänninen, N. B. Kopnin, J. Kopu,M. Krusius, M. Tsubota, and G. E. Volovik, 2006, Rep. Prog.Phys. 69, 3157.

Fischer, U. R., and G. Baym, 2003, Phys. Rev. Lett. 90, 140402.Fu, H., and E. Zaremba, 2006, Phys. Rev. A 73, 013614.García-Ripoll, J. J., and V. M. Pérez-García, 2001a, Phys. Rev.

A 63, 041603.García-Ripoll, J. J., and V. M. Pérez-García, 2001b, Phys. Rev.

A 64, 053611.Ginzburg, V. L., and L. P. Pitaevskii, 1958, Zh. Eksp. Teor. Fiz.

34, 1240 �Sov. Phys. JETP 7, 858 �1958��.Giorgini, S., L. P. Pitaevskii, and S. Stringari, 2008, Rev. Mod.

Phys. 80, 1215.Gottfried, K., 1966, Quantum Mechanics �Benjamin, New

York�, Vol. I.Greiner, M., O. Mandel, T. Esslinger, T. W. Hänsch, and I.

Bloch, 2002, Nature �London� 415, 39.Griesmaier, A., J. Werner, S. Hensler, J. Stuhler, and T. Pfau,

2005, Phys. Rev. Lett. 94, 160401.Griffiths, D. J., 2005, Introduction to Quantum Mechanics, 2nd

ed. �Pearson Prentice Hall, Upper Saddle River, NJ�.Gross, E. P., 1961, Nuovo Cimento 20, 454.Hadzibabic, Z., P. Krüger, M. Cheneau, B. Battelier, and J.

Dalibard, 2006, Nature �London� 441, 1118.Hadzibabic, Z., P. Krüger, M. Cheneau, S. P. Rath, and J. Dali-

bard, 2008, New J. Phys. 10, 045006.Haldane, F. D., and Y.-S. Wu, 1985, Phys. Rev. Lett. 55, 2887.Haljan, P. C., B. P. Anderson, I. Coddington, and E. A. Cor-

nell, 2001, Phys. Rev. Lett. 86, 2922.Haljan, P. C., I. Coddington, P. Engels, and E. A. Cornell, 2001,

Phys. Rev. Lett. 87, 210403.Hall, D. S., M. R. Matthews, C. E. Wieman, and E. A. Cornell,

1998, Phys. Rev. Lett. 81, 1543.Ho, T.-L., 1998, Phys. Rev. Lett. 81, 742.Ho, T.-L., 2001, Phys. Rev. Lett. 87, 060403.Hodby, E., G. Hechenblaikner, S. A. Hopkins, O. M. Maragò,

and C. J. Foot, 2002, Phys. Rev. Lett. 88, 010405.Hohenberg, P. C., 1967, Phys. Rev. 158, 383.Hohenberg, P. C., and P. C. Martin, 1965, Ann. Phys. �N.Y.� 34,

291.Huhtamäki, J. A. M., M. Möttönen, T. Isoshima, V. Pietilä, and

S. M. M. Virtanen, 2006, Phys. Rev. Lett. 97, 110406.Inguscio, M., S. Stringari, and C. E. Weiman, 1999, Eds., Bose-

Einstein Condensation in Atomic Gases �IOS, Amsterdam�.Isoshima, T., K. Machida, and T. Ohmi, 2001, J. Phys. Soc. Jpn.

70, 1604.Isoshima, T., M. Nakahara, T. Ohmi, and K. Machida, 2000,

Phys. Rev. A 61, 063610.Jain, J. K., 1989, Phys. Rev. Lett. 63, 199.Jaksch, D., C. Bruder, J. I. Cirac, C. W. Gardiner, and P. Zol-

ler, 1998, Phys. Rev. Lett. 81, 3108.Jaksch, D., and P. Zoller, 2003, New J. Phys. 5, 56.1.Josephson, B. D., 1962, Phys. Lett. 1, 251.Juzeliunas, G., P. Öhberg, J. Ruseckas, and A. Klein, 2005,

Phys. Rev. A 71, 053614.Kasamatsu, K., and M. Tsubota, 2008, Prog. Low Temp. Phys.

16, 349.Kasamatsu, K., M. Tsubota, and M. Ueda, 2002, Phys. Rev. A

66, 053606.Kasamatsu, K., M. Tsubota, and M. Ueda, 2003, Phys. Rev.

Lett. 91, 150406.Kavoulakis, G. M., and G. Baym, 2003, New J. Phys. 5, 51.1.Khawaja, U. A., 2005, Phys. Rev. A 71, 063611.Kim, J.-K., and A. L. Fetter, 2005, Phys. Rev. A 72, 023619.Kleiner, W. H., L. M. Roth, and S. H. Autler, 1964, Phys. Rev.

133, A1226.Koch, T., T. Lahaye, J. Metz, B. Fröhlich, A. Griesmaier, and

T. Pfau, 2008, Nat. Phys. 4, 218.Köhler, T., K. Góral, and P. S. Julienne, 2006, Rev. Mod. Phys.

78, 1311.Komineas, S., and N. R. Cooper, 2007, Phys. Rev. A 75,

023623.Kosterlitz, J. M., and D. J. Thouless, 1973, J. Phys. C 6, 1181.Krüger, P., 2007, private communication.Lahaye, T., T. Koch, B. Fröhlich, M. Fattori, J. Metz, A. Gries-

maier, S. Giovanazzi, and T. Pfau, 2007, Nature �London�448, 672.

Lamb, H., 1945, Hydrodynamics, 6th ed. �Dover, New York�.Landau, L. D., 1941, J. Phys. �USSR� 5, 71.Landau, L. D., and E. M. Lifshitz, 1960, Mechanics �Pergamon,

Oxford�.Landau, L. D., and E. M. Lifshitz, 1965, Quantum Mechanics

Non-Relativistic Theory, 2nd ed. �Pergamon, Oxford�.Landau, L. D., and E. M. Lifshitz, 1987, Fluid Mechanics, 2nd

ed. �Pergamon, Oxford�.Laughlin, R. B., 1983, Phys. Rev. Lett. 50, 1395.Leanhardt, A. E., A. Görlitz, A. P. Chikkatur, D. Kielpinski, Y.

Shing, D. E. Pritchard, and W. Ketterle, 2002, Phys. Rev. Lett.89, 190403.

Leggett, A. J., 2001, Rev. Mod. Phys. 73, 307.Leggett, A. J., 2006, Quantum Liquids: Bose Condensation and

Cooper Pairing in Condensed-Matter Systems �Oxford Uni-versity Press, New York�.

Lifshitz, E. M., and L. P. Pitaevskii, 1980a, Statistical Physics,3rd ed. �Pergamon, Oxford�, Vol. 1.

Lifshitz, E. M., and L. P. Pitaevskii, 1980b, Statistical Physics,3rd ed. �Pergamon, Oxford�, Vol. 2.

Linn, M., and A. L. Fetter, 1999, Phys. Rev. A 60, 4910.Linn, M., M. Niemeyer, and A. L. Fetter, 2001, Phys. Rev. A

64, 023602.London, F., 1954, Superfluids: Microscopic Theory of Super-

fluid Helium �Dover, New York�, Vol. II.Lundh, E., 2002, Phys. Rev. A 65, 043604.Lundh, E., and P. Ao, 2000, Phys. Rev. A 61, 063612.Lundh, E., C. J. Pethick, and H. Smith, 1997, Phys. Rev. A 55,

2126.Madison, K. W., F. Chevy, F. Bretin, and J. Dalibard, 2001,

Phys. Rev. Lett. 86, 4443.Madison, K. W., F. Chevy, W. Wohlleben, and J. Dalibard,

2000a, Phys. Rev. Lett. 84, 806.Madison, K. W., F. Chevy, W. Wohlleben, and J. Dalibard,

2000b, J. Mod. Opt. 47, 2715.Martikainen, J.-P., and H. T. C. Stoof, 2003, Phys. Rev. Lett. 91,

240403.Mason, P., N. G. Berloff, and A. L. Fetter, 2006, Phys. Rev. A

74, 043611.Matthews, M. R., B. P. Anderson, P. C. Haljan, D. S. Hall, M.

689Alexander L. Fetter: Rotating trapped Bose-Einstein condensates

Rev. Mod. Phys., Vol. 81, No. 2, April–June 2009

Page 44: Rotating trapped Bose-Einstein condensatespjh/teaching/phz7429/fetter.pdfII. Physics of Bose-Einstein Condensates in Dilute Trapped Gases 648 A. Ideal Bose gas 649 1. Ideal BEC in

J. Holland, J. E. Williams, C. E. Wieman, and E. A. Cornell,1999a, Phys. Rev. Lett. 83, 3358.

Matthews, M. R., B. P. Anderson, P. C. Haljan, D. S. Hall, C.E. Wieman, and E. A. Cornell, 1999b, Phys. Rev. Lett. 83,2498.

McGee, S. A., and M. J. Holland, 2001, Phys. Rev. A 63,043608.

Menotti, C., and M. Lewenstein, 2008, Recent Progress inMany-Body Theories �World Scientific, Singapore�, p. 79.

Mermin, N. D., and H. Wagner, 1966, Phys. Rev. Lett. 17, 1133.Messiah, A., 1961, Quantum Mechanics �Wiley, New York�,

Vol. I.Mizushima, T., M. Ichioka, and K. Machida, 2003, Phys. Rev.

Lett. 90, 180401.Mizushima, T., N. Kobayashi, and K. Machida, 2004, Phys.

Rev. A 70, 043613.Mizushima, T., K. Machida, and T. Kita, 2002, Phys. Rev. Lett.

89, 030401.Modugno, M., L. Pricoupenko, and Y. Castin, 2003, Eur. Phys.

J. D 22, 235.Moerdijk, A. J., B. J. Verhaar, and A. Axelsson, 1995, Phys.

Rev. A 51, 4852.Möller, G., and N. R. Cooper, 2007, Phys. Rev. Lett. 99,

190409.Morris, A. G., and D. L. Feder, 2007, Phys. Rev. Lett. 99,

240401.Möttönen, M., T. Mizushima, T. Isoshima, M. M. Salomaa, and

K. Machida, 2003, Phys. Rev. A 68, 023611.Mueller, E. J., 2004, Phys. Rev. A 70, 041603�R�.Mueller, E. J., and T.-L. Ho, 2002, Phys. Rev. Lett. 88, 180403.Mueller, E. J., and T.-L. Ho, 2003, Phys. Rev. A 67, 063602.Murray, D. R., S. M. Barnett, P. Öberg, and D. Gomila, 2007,

e-print arXiv:0709.0895v1.Muryshev, A. E., H. B. van Linden van den Heuvell, and G. V.

Shlyapnikov, 1999, Phys. Rev. A 60, R2665.Myatt, C. J., E. A. Burt, R. W. Ghrist, E. A. Cornell, and C. E.

Wieman, 1997, Phys. Rev. Lett. 78, 586.Nakahara, M., T. Isoshima, K. Machida, S. Ogawa, and T.

Ohmi, 2000, Physica B 284-288, 17.Nepomnyashchi, Y. A., 1974, Theor. Math. Phys. 20, 904.Nozières, P., and D. Saint James, 1982, J. Phys. �Paris� 43, 1133.Ogawa, S.-I., M. Möttönen, M. Nakahara, T. Ohmi, and H.

Shimada, 2002, Phys. Rev. A 66, 013617.Ohmi, T., and K. Machida, 1998, J. Phys. Soc. Jpn. 67, 1822.Oktel, M. Ö., 2004, Phys. Rev. A 69, 023618.Onofrio, R., D. S. Durfee, C. Raman, M. Köhl, C. E.

Kuklewicz, and W. Ketterle, 2000, Phys. Rev. Lett. 84, 810.Onsager, L., 1949, Nuovo Cimento, Suppl. 6, 249.Osterloh, K., N. Barberán, and M. Lewenstein, 2007, Phys.

Rev. Lett. 99, 160403.Packard, R. E., and T. M. Sanders, 1972, Phys. Rev. A 6, 799.Parker, N. G., B. Jackson, A. M. Martin, and C. S. Adams,

2007, e-print arXiv:0704.0146v1.Pathria, R. K., 1996, Statistical Mechanics, 2nd ed.

�Butterworth-Heineman, Oxford�.Penckwitt, A. A., R. J. Ballagh, and C. W. Gardiner, 2002,

Phys. Rev. Lett. 89, 260402.Penrose, O., and L. Onsager, 1956, Phys. Rev. 104, 576.Pérez-García, V. M., H. Michinel, J. I. Cirac, M. Lewenstein,

and P. Zoller, 1996, Phys. Rev. Lett. 77, 5320.Pethick, C. J., and H. Smith, 2002, Bose-Einstein Condensation

in Dilute Gases �Cambridge University Press, Cambridge, En-gland�.

Pines, D., 1961, The Many-Body Problem �Benjamin, NewYork�.

Pismen, L. M., 1999, Vortices in Nonlinear Fields �Clarendon,Oxford�.

Pismen, L. M., and B. Y. Rubinstein, 1991, Physica D 47, 353.Pitaevskii, L. P., 1961, Zh. Eksp. Teor. Fiz. 40, 646 �Sov. Phys.

JETP 13, 451 �1961��.Pitaevskii, L., and S. Stringari, 2003, Bose-Einstein Condensa-

tion �Clarendon, Oxford�.Pollock, F., 1967, Phys. Fluids 10, 473.Prange, R. E., and S. M. Girvin, 1987, Eds., The Quantum Hall

Effect �Springer-Verlag, New York�.Pu, H., C. K. Law, J. H. Eberly, and N. P. Bigelow, 1999, Phys.

Rev. A 59, 1533.Raman, C., J. R. Abo-Shaeer, J. M. Vogels, K. Xu, and W.

Ketterle, 2001, Phys. Rev. Lett. 87, 210402.Read, N., and N. R. Cooper, 2003, Phys. Rev. A 68, 035601.Regnault, N., and T. Jolicoeur, 2003, Phys. Rev. Lett. 91,

030402.Regnault, N., and T. Jolicoeur, 2004, Phys. Rev. B 69, 235309.Resnick, D. J., J. C. Garland, J. T. Boyd, S. Shoemaker, and R.

S. Newrock, 1981, Phys. Rev. Lett. 47, 1542.Rokhsar, D. S., 1997a, e-print arXiv:cond-mat/9709212.Rokhsar, D. S., 1997b, Phys. Rev. Lett. 79, 2164.Rosenbusch, P., V. Bretin, and J. Dalibard, 2002, Phys. Rev.

Lett. 89, 200403.Rubinstein, B. Y., and L. M. Pismen, 1994, Physica D 78, 1.Ruprecht, P. A., M. J. Holland, K. Burnett, and M. Edwards,

1995, Phys. Rev. A 51, 4704.Ryu, C., M. F. Andersen, P. Cladé, V. Natarajan, K. Helmerson,

and W. D. Phillips, 2007, Phys. Rev. Lett. 99, 260401.Sadler, L. E., J. M. Higbie, S. R. Leslie, M. Vengalattore, and

D. M. Stamper-Kurn, 2006, Nature �London� 443, 312.Salomaa, M. M., and G. E. Volovik, 1987, Rev. Mod. Phys. 59,

533.Sánchez-Lotero, P., and J. J. Palacios, 2005, Phys. Rev. A 72,

043613.Satija, I. I., D. C. Dakin, and C. W. Clark, 2006, Phys. Rev.

Lett. 97, 216401.Scherer, D. R., C. N. Weiler, T. N. Neely, and B. P. Anderson,

2007, Phys. Rev. Lett. 98, 110402.Schiff, L. I., 1968, Quantum Mechanics, 3rd ed. �McGraw-Hill,

New York�.Schweikhard, V., I. Coddington, P. Engels, V. P. Mogendorff,

and E. A. Cornell, 2004, Phys. Rev. Lett. 92, 040404.Schweikhard, V., S. Tung, and E. A. Cornell, 2007, Phys. Rev.

Lett. 99, 030401.Sedrakian, A., and I. Wasserman, 2001, Phys. Rev. A 63,

063605.Sensarma, R., M. Randeria, and T.-L. Ho, 2006, Phys. Rev.

Lett. 96, 090403.Shankar, R., 1994, Principles of Quantum Mechanics, 2nd ed.

�Plenum, New York�.Sheehy, D. E., and L. Radzihovsky, 2004a, Phys. Rev. A 70,

051602�R�.Sheehy, D. E., and L. Radzihovsky, 2004b, Phys. Rev. A 70,

063620.Shin, Y., M. Saba, M. Vengalattore, T. A. Pasquini, C. Sanner,

A. E. Leanhardt, M. Prentiss, D. E. Pritchard, and W. Ket-terle, 2004, Phys. Rev. Lett. 93, 160406.

Simula, T. P., and P. B. Blakie, 2006, Phys. Rev. Lett. 96,020404.

Sinha, S., 1997, Phys. Rev. A 55, 4325.

690 Alexander L. Fetter: Rotating trapped Bose-Einstein condensates

Rev. Mod. Phys., Vol. 81, No. 2, April–June 2009

Page 45: Rotating trapped Bose-Einstein condensatespjh/teaching/phz7429/fetter.pdfII. Physics of Bose-Einstein Condensates in Dilute Trapped Gases 648 A. Ideal Bose gas 649 1. Ideal BEC in

Sinha, S., and Y. Castin, 2001, Phys. Rev. Lett. 87, 190402.Sinha, S., and G. V. Shlyapnikov, 2005, Phys. Rev. Lett. 94,

150401.Sinova, J., C. B. Hanna, and A. H. MacDonald, 2002, Phys.

Rev. Lett. 89, 030403.Sonin, E. B., 1987, Rev. Mod. Phys. 59, 87.Sonin, E. B., 2005, Phys. Rev. A 71, 011603�R�.Sørensen, A., E. Demler, and M. Lukin, 2005, Phys. Rev. Lett.

94, 086803.Stamper-Kurn, D. M., M. R. Andrews, A. P. Chikkatur, S. In-

ouye, H.-J. Miesner, J. Stenger, and W. Ketterle, 1998, Phys.Rev. Lett. 80, 2027.

Stenger, J., S. Inouye, D. M. Stamper-Kurn, H. J. Miesner, A.P. Chikkatur, and W. Ketterle, 1998, Nature �London� 396,345.

Stock, S., B. Battelier, V. Bretin, Z. Hadzibabic, and J. Dali-bard, 2005, Laser Phys. Lett. 2, 275.

Stock, S., V. Bretin, F. Chevy, and J. Dalibard, 2004, Europhys.Lett. 65, 594.

Stock, S., Z. Hadzibabic, B. Battelier, M. Cheneau, and J. Dali-bard, 2005, Phys. Rev. Lett. 95, 190403.

Stringari, S., 1996, Phys. Rev. Lett. 77, 2360.Svidzinsky, A. A., and A. L. Fetter, 1998, Phys. Rev. A 58,

3168.Svidzinsky, A. A., and A. L. Fetter, 2000a, Phys. Rev. Lett. 84,

5919.Svidzinsky, A. A., and A. L. Fetter, 2000b, Phys. Rev. A 62,

063617.Thomson, W., 1880, Philos. Mag. 10, 155.Tinkham, M., 1996, Introduction to Superconductivity, 2nd ed.

�McGraw-Hill, New York�.Tkachenko, V. K., 1966a, Zh. Eksp. Teor. Fiz. 49, 1875 �Sov.

Phys. JETP 22, 1282 �1966��.Tkachenko, V. K., 1966b, Zh. Eksp. Teor. Fiz. 50, 1573 �Sov.

Phys. JETP 23, 1049 �1966��.Tkachenko, V. K., 1969, Zh. Eksp. Teor. Fiz. 56, 1763 �Sov.

Phys. JETP 29, 949 �1969��.Tonini, G., F. Werner, and Y. Castin, 2006, Eur. Phys. J. D 39,

283.Trugman, S. A., and S. Kivelson, 1985, Phys. Rev. B 31, 5280.Tsubota, M., K. Kasamatsu, and M. Ueda, 2001, Phys. Rev. A

65, 023603.Valatin, J. G., 1956, Proc. R. Soc. London, Ser. A 238, 132.Vengalattore, M., S. R. Leslie, J. Guzman, and D. M. Stamper-

Kurn, 2008, Phys. Rev. Lett. 100, 170403.Viefers, S., 2008, J. Phys.: Condens. Matter 20, 123202.Viefers, S., T. H. Hansson, and S. M. Reimann, 2000, Phys.

Rev. A 62, 053604.Vollhardt, D., and P. Wölfle, 1990, The Superfluid Phases of

Helium 3 �Taylor and Francis, London�.Watanabe, G., G. Baym, and C. J. Pethick, 2004, Phys. Rev.

Lett. 93, 190401.Wilczek, F., 1990, Fractional Statistics and Anyon Superconduc-

tivity �World Scientific, Singapore�.Wilkin, N. K., and J. M. F. Gunn, 2000, Phys. Rev. Lett. 84, 6.Wilkin, N. K., J. M. F. Gunn, and R. A. Smith, 1998, Phys. Rev.

Lett. 80, 2265.Williams, J. E., and M. J. Holland, 1999, Nature �London� 401,

568.Williams, J., R. Walser, J. Cooper, E. A. Cornell, and M. Hol-

land, 2000, Phys. Rev. A 61, 033612.Wolfram, S., 2007, http://functions.wolfram.com/Wu, B., and Q. Niu, 2001, Phys. Rev. A 64, 061603�R�.Yang, C. N., 1962, Rev. Mod. Phys. 34, 694.Yarmchuk, E. J., M. J. V. Gordon, and R. E. Packard, 1979,

Phys. Rev. Lett. 43, 214.Yarmchuk, E. J., and R. E. Packard, 1982, J. Low Temp. Phys.

46, 479.Yip, S.-K., 1999, Phys. Rev. Lett. 83, 4677.Zambelli, F., and S. Stringari, 1998, Phys. Rev. Lett. 81, 1754.Zhai, H., and T.-L. Ho, 2006, Phys. Rev. Lett. 97, 180414.Zhang, J., and H. Zhai, 2005, Phys. Rev. Lett. 95, 200403.Zwierlein, M. W., 2007, private communication.Zwierlein, M. W., J. R. Abo-Shaeer, A. Schirotzek, C. H.

Schunk, and W. Ketterle, 2005, Nature �London� 435, 1047.

691Alexander L. Fetter: Rotating trapped Bose-Einstein condensates

Rev. Mod. Phys., Vol. 81, No. 2, April–June 2009