Download as pdf or txt
Download as pdf or txt
You are on page 1of 50

REVIEWS OF MODERN PHYSICS, VOLUME 73, APRIL 2001

Bose-Einstein condensation in the alkali gases:


Some fundamental concepts
Anthony J. Leggett
Department of Physics, University of Illinois at Urbana-Champaign, Urbana,
Illinois 61801-3080

Whatever can be said, can be said clearly; and whereof one cannot speak,
thereof one should keep silent.
L. Wittgenstein
(Published 24 April 2001)

The author presents a tutorial review of some ideas that are basic to our current understanding of the
phenomenon of Bose-Einstein condensation (BEC) in the dilute atomic alkali gases, with special
emphasis on the case of two or more coexisting hyperfine species. Topics covered include the
definition of and conditions for BEC in an interacting system, the replacement of the true interatomic
potential by a zero-range pseudopotential, the time-independent and time-dependent Gross-Pitaevskii
equations, superfluidity and rotational properties, the Josephson effect and related phenomena, and
the Bogoliubov approximation.

CONTENTS C. Equilibrium of a BEC system in a rotating


container 331
D. Metastability of superflow 333
I. Introduction. Scope of the Review 307 1. General considerations 333
II. The Systems 309 2. A toy model 334
A. General 309 3. Further remarks 335
B. Trapping potentials 309 E. Real-life BEC alkali gases in harmonic traps 335
1. Laser traps 310 F. The experimental situation 336
2. Magnetic traps 310 VII. BEC in a Two-State System: Josephson-Type
Effects, Phase Diffusion 336
3. Gravity 311
A. General formulation: choice of basis 337
C. The hyperfine and Zeeman interactions 311
B. Realizations in the BEC alkali gases 338
D. Imaging 312 C. Kinematics of the Josephson effect: the Rabi,
E. Orders of magnitude 312 Josephson, and Fock regimes 339
III. The Definition, Origin, and Occurrence of BEC: D. The Josephson regime: Josephson resonance and
The Order Parameter 313 macroscopic quantum self-trapping 341
A. Definition of BEC 313 E. The Fock regime: phase diffusion 343
B. Why BEC? 314 VIII. The Bogoliubov Approximation 347
C. Rigorous results 315 A. Inconsistency of the Gross-Pitaevskii
D. The order parameter and the superfluid velocity 316 approximation 347
B. The Bogoliubov ground state in the translation-
1. Possible definitions of the order parameter 316
invariant case 347
2. The superfluid velocity 317
C. Properties of the Bogoliubov ground state:
IV. The Effective Interaction in a Cold Dilute Gas 317 elementary excitations 349
A. Statement of the problem: neglect of l⫽0 partial D. The inhomogeneous case 350
waves 317 E. Time-dependent Bogoliubov–de Gennes
B. The s-wave scattering length 318 equations: connection with the time-
C. The effective interaction 319 dependent Gross-Pitaevskii equation 351
D. Effects of indistinguishability 320 F. The multicomponent case 352
E. Effect of the hyperfine degree of freedom 321 IX. Further Topics 352
F. Time-dependent situations: the MIT hydrogen A. Attractive interactions 352
experiments 322 B. Optical properties 352
V. The Gross-Pitaevskii Approximation 307 C. Coexistence of three hyperfine species 353
A. The Gross-Pitaevskii ground state of a spinless D. The ‘‘atom laser’’ 353
system 323 E. Kinetics, damping, relaxation, etc. 353
B. The spinless gas: finite-temperature equilibrium 324 F. Late-breaking developments 354
C. The spinless gas: time-dependent Gross- Acknowledgments 354
Pitaevskii theory 324 References 354
D. Effects of the hyperfine degree of freedom 326
E. Applications 329 I. INTRODUCTION. SCOPE OF THE REVIEW
VI. Rotational Properties: Superfluidity 330
A. Phenomenology of superfluidity in liquid 4He 330 The phenomenon known as Bose-Einstein condensa-
B. Rotating frames of reference 331 tion (hereafter abbreviated BEC) was predicted by Ein-

0034-6861/2001/73(2)/307(50)/$30.00 307 ©2001 The American Physical Society


308 Anthony J. Leggett: Bose-Einstein condensation in the alkali gases

stein in 1924 on the basis of ideas of Bose concerning ter. Generally speaking, I have spent a fair amount of
photons: In a system of particles obeying Bose statistics time on the derivation of fundamental equations such as
and whose total number is conserved, there should be a the Gross-Pitaevskii equation and the associated con-
temperature below which a finite fraction of all the par- ceptual issues, but much less on their applications to
ticles ‘‘condense’’ into the same one-particle state. Ein- specific experimental systems. This is not, of course, be-
stein’s original prediction was for a noninteracting gas, a cause I feel that such applications are any less impor-
system felt by some of his contemporaries to be perhaps tant, but because it is this part of the program that tends
pathological, but shortly after the observation of super- to be discussed explicitly in the regular journal litera-
fluidity in liquid 4He below the ␭ temperature (2.17 K), ture, while the logically prior steps are often taken for
Fritz London suggested that despite the strong inter- granted. Second, I have taken advantage of the fact that
atomic interactions BEC was indeed occurring in this there already exist several good reviews of the field (e.g.,
system and was responsible for the superfluid properties; Parkin and Walls, 1998; Pethick and Smith, 2000; and
this suggestion has stood the test of time and is the basis several of the articles in Inguscio et al., 1999). In particu-
for our modern understanding of the properties of the lar, the May 1999 issue of this journal contains an excel-
superfluid phase. In 1995 BEC was realized in a system lent review by Dalfovo et al. (1999) of much of the ex-
that is about as different as possible from 4He, namely, perimental work and relevant theory of the first three
dilute atomic alkali gases trapped by magnetic fields, years of the subject, and I have therefore been able sim-
and over the last few years these systems have been the ply to refer the reader to that for many of the standard
subject of an explosion of research, both experimental applications of the ideas I shall discuss. However, one
and theoretical, which in addition to drawing on estab- aspect that is barely touched on by Dalfovo et al. and
lished lore in the areas of atomic collisions, quantum that has taken center stage over the last 18 months or so
optics, and condensed-matter physics has generated is the special class of phenomena associated with the
problems and ideas specific to these novel systems. hyperfine degree of freedom, so I spend a fair amount of
Perhaps the single aspect of BEC systems that makes time on this. For reasons explained in Sec. IX, I have not
them most fascinating is best illustrated by the cover of attempted to cover the important subject of nonlinear
Science magazine of December 22, 1995, in which the kinetics. Finally, purely for reasons of space, I have been
Bose condensate is declared ‘‘molecule of the year’’ and unable to mention at all the highly nontrivial experimen-
pictured as a platoon of soldiers marching in lockstep: tal techniques required to cool the alkali gases into the
every atom in the condensate must behave in exactly the BEC regime [on this, see, for example, the Nobel lec-
same way, and this has the consequence, inter alia, that tures of Chu, Cohen-Tannoudji, and Phillips (1998) and
effects which are so small as to be essentially invisible at Ketterle et al. (1999)] and have mentioned trapping and
the level of a single atom may be spectacularly ampli- imaging techniques (in Sec. II) only to the extent that it
fied. (An example is the phenomenon of superfluidity, is necessary to motivate the subsequent theoretical dis-
discussed in Sec. VI.) In addition, as we shall see by cussion. Some further material that I would have in-
implication, this property tests, in rather subtle ways, cluded here had space allowed may be found in Leggett
our understanding of the meaning of the formalism of (2000a).
quantum mechanics, the nature of ‘‘randomness,’’ and The plan of the review is as follows. In Sec. II I give a
much else. brief introduction to the experimental systems, with em-
This paper does not attempt to be a comprehensive phasis on orders of magnitude and on the role of the
review of the field of BEC in the alkali gases, even as hyperfine degree of freedom. Section III is devoted to
regards its theoretical aspects. Rather, it is intended to the origin, definition, and occurrence of BEC and some
be tutorial in nature, and the reader I have had specifi- related concepts such as the order parameter and super-
cally in mind is a graduate student about to embark on fluid velocity. In Sec. IV I discuss in some detail the
research, either experimental or theoretical, in this area. effective interatomic interaction (a topic that tends to be
I believe that a major difficulty such a student is likely to passed over in most papers in a couple of lines), with
face is that there are many ideas which are taken for some attention to the infamous ‘‘factor of 2,’’ which is
granted by workers in the field (or at least by a subset of liable to confuse newcomers to the field, particularly in
them) but for which it is difficult to give an explicit and the multispecies case. Section V derives and discusses
useful reference; this is particularly true of ideas that the time-independent and time-dependent Gross-
were originally developed, in some cases as long ago as Pitaevskii equations, including their generalizations to
the 1940s, in the context and language of liquid helium finite temperature and to multispecies systems. Section
and have not always been rephrased in the terminology VI is devoted to the rotational properties of a Bose con-
that has become standard in the BEC field. My goal densate and the associated notion of superfluidity; here,
therefore has been to set out as clearly as I know how a in contrast to much of the existing alkali-gas literature, I
set of concepts which I believe are basic to our under- tend to play down ‘‘vortices’’ and concentrate on intrin-
standing of BEC in the alkali gases, and to discuss in at sically multiply connected topologies, as I believe that
least a schematic way how they relate to existing or con- this permits a much clearer conceptual picture. In Sec.
templated experiments. VII, which is the longest in the paper, I discuss in some
Even given this limited goal, space and other consid- detail the ‘‘toy model’’ that results when the N bosons in
erations dictate a severe restriction of the subject mat- question are restricted to a single two-dimensional Hil-

Rev. Mod. Phys., Vol. 73, No. 2, April 2001


Anthony J. Leggett: Bose-Einstein condensation in the alkali gases 309

bert space; this permits a discussion of various phenom- requires as a first step the recombination of two atoms
ena (Josephson effect, phase diffusion, etc.) that have to form a diatomic molecule, and while this process is
more complicated, and in my opinion often less well- certainly exothermic (with formation energies typically
understood, analogs in the case of an extended system. ⬃0.4– 1.2 eV) it is very slow in the absence of a third
Section VIII introduces the Bogoliubov approximation, atom to carry off the surplus energy, angular momen-
from a point of view that is not the majority one in the tum, etc. Thus in practice the dominant recombination
literature, and discusses its relation to the time- processes are usually ‘‘three-body’’ ones; the rate per
dependent Gross-Pitaevskii description. Section IX atom is typically ⬃10⫺29 – 10⫺30 cm6 sec⫺1, giving a
briefly lists and comments on some further topics which sample lifetime of the order of a few seconds to a few
for various reasons are not covered in the main text. minutes.
The reader who compares this review with much of An alkali atom in its ground state has a single valence
the existing literature on BEC, whether in 4He or the electron in an ns state outside one or more closed shells
alkali gases, will be struck by the fact that I virtually (except for H); the electronic ground state is therefore a
bend over backwards to avoid introducing the idea of doublet. Except in the case of H, the only electronic
‘‘spontaneously broken gauge symmetry.’’ (I do use it excited state that is of much interest is the np state
once, reluctantly, in Sec. VIII, but only for the purpose (since it is overwhelmingly to this state that radiation in
of making contact with the formalism that has become the optical regime will couple the ground state); the
standard in the literature.) It would be disingenuous of wavelengths ␭ of the ‘‘fundamental’’ (ns→np) transi-
me to attempt to conceal the fact that I believe the util- tion lie in the range 5000–7000 Å and the excited-state
ity of this idea is outweighed by its dangers; see Sec. lifetimes in the range 16–35 nsec (see, for example,
III.D.1 Weiner et al., 1999, Table II). [In the case of H (n⫽1)
Finally, a note on referencing policy. In these days of there is of course no 1p state and a single photon will
automated databases, I do not believe it is particularly preliminarily excite the atom to the 2p state; however, a
useful for a review of this type to attempt to provide a pair of photons can excite the 1s→2s transition, and this
comprehensive list of papers in the literature that deal has played an important role in recent experiments.]
with the various topics described, still less to grade them If we treat the atom for the moment as a single indi-
according to relative significance. Generally speaking, visible entity (see Sec. III.A) and consider the exchange
when I am discussing basic conceptual issues I have tried of two atoms of the same (chemical and isotopic) spe-
to reference at least a selection of what I regard as the cies, this involves exchanging Z⫹A fermions, and thus
most important discussions (which are not necessarily the total wave function should be symmetric or antisym-
the chronologically earliest ones); however, once it metric under this exchange according as Z⫹A is even or
comes to applications of the basic equations I have often odd. Since Z is automatically odd for the alkali ele-
referenced only one or two recent papers from which I ments, this means that a system of identical odd-A iso-
believe most of the rest of the relevant literature can be topes will obey Bose-Einstein statistics (and hence has
traced. Obviously the choice of these is highly subjective the possibility to display BEC), while an even-A system
and indeed arbitrary, and I hope that the many authors such as 6Li or 40K will obey Fermi-Dirac statistics. We
whose papers I have failed to reference will not attribute shall be interested from now on in the former case, al-
this to malice! though the latter is also very interesting (see de Marco
and Jin, 1999). The odd-A isotopes, other than 1H, in
which BEC has been demonstrated at the time of writ-
II. THE SYSTEMS ing, are 87Rb, 23Na, 7Li, and, very recently, 85Rb; as it
happens, each of the first three has nuclear spin 3/2, so
A. General
when discussing the effects of nuclear spin in Sec. II.C I
The experimental systems I shall be addressing in this shall concentrate on this case. A table of the principal
review are collections of individual neutral alkali-gas stable or long-lived alkali Bose isotopes may be found in
atoms,1 with total number N ranging from a few hun- Pethick and Smith (2000), Chap. 3.
dred up to ⬃1010, confined by magnetic and/or optical
means to a relatively small region of space. Their (maxi-
mum) densities range from ⬃1011 cm⫺3 to ⬃5⫻1015 B. Trapping potentials
cm⫺3, and their temperature, in the regime of interest to
us, is typically in the range of a few tens of nK– ⬃50 ␮K. For the purposes of this review it is not necessary to
It is needless to mention that such an atomic gas cannot go into the details of the various ingenious schemes that
in fact be the stable thermodynamic state of the N at- have been developed over the last 15 years or so for the
oms, which would at these temperatures certainly corre- trapping of neutral atoms (or, what is equally important
spond to a solid; however, the formation of the solid for the real-life experiments, for their cooling into the
␮ K– nK regime where BEC can occur); for these topics,
the reader is referred, for example, to Ketterle et al.
1
It is convenient for the present purpose to include H in the (1999) or to the 1998 Nobel lectures. However, it is im-
alkalis. However, some of the values quoted as ‘‘typical’’ be- portant to appreciate the general features of the result-
low do not apply to it. ing effective potentials in which the atoms move, and in

Rev. Mod. Phys., Vol. 73, No. 2, April 2001


310 Anthony J. Leggett: Bose-Einstein condensation in the alkali gases

particular their dependence (or not) on the hyperfine- ton, in the (admittedly unlikely) event that its energy
Zeeman degree of freedom to be discussed in Sec. II.C. were totally absorbed in the atomic gas, would be more
than enough to heat it right out of the BEC phase]. As
shown in detail by Cohen-Tannoudji (1992), for ⌫Ⰶ⌬
1. Laser traps the emission probability per atom is of order
For an excellent account of the general subject of 关 I(r)/I o 兴 ប ⫺1 ⌫ 3 /⌬ 2 . In view of the ⌬ ⫺2 falloff by com-
atom-laser interactions, the reader is referred to Cohen- parison with the ⌬ ⫺1 behavior in Eq. (2.3), it is generally
Tannoudji (1992). The effect that has been principally advantageous to tune the laser(s) very far off resonance.
exploited in the laser trapping of atoms in the BEC re- In fact, it is quite common to detune by ⬃10% of the
gime (as distinct from their cooling into this regime) is original optical frequency,2 so that the ratio ⌫/⌬ is
the so-called dipole effect, which relies on the (conser- ⬃10⫺7 – 10⫺6 ; then, even in the red-detuned case with,
vative) interaction of the laser field with the electric di- say, 106 atoms in one’s sample, the spontaneous emis-
pole moment it induces on the atom. For the purposes of sion rate is only a fraction of a sec⫺1 and is unlikely to
an initial discussion, let us ignore the fine and hyperfine affect the experiment. When blue-detuned radiation is
structure and define the detuning of the ns→np transi- used to provide a barrier, the spontaneous emission ef-
tion frequency in the standard way: fect is obviously less important, since the atoms do not
appreciably penetrate the barrier region. The crucial
⌬⬅ប ␻ las ⫺ 共 E np ⫺E ns 兲共 ⬅ប ␻ las ⫺2 ␲ បc/␭ 兲 . (2.1) qualitative point is that it is possible in this way to pro-
It is also convenient to define the saturation intensity vide laser-generated potentials that are of an order
I o , that is, the laser-beam intensity, which, when exactly greater than the thermal energy (or other characteristic
on resonance, will induce a population of order unity of energy scales; see Sec. III) without appreciable heating
the excited (p) state. To within a numerical factor of effects. Note also that the time scales over which such
order unity which depends on polarization, etc., this is potentials can be manipulated are extremely short (be-
given by ing in fact typically limited by the turning-on time of the
laser). For a more extended discussion of laser (optical)
I o ⫽ ⑀ 0 c⌫ 2 /d 2 , (2.2) confinement I refer the reader to Stenger et al. (1998b).
where d is an appropriately defined dipole matrix ele-
ment for the transition in question and ⌫⬅ប/ ␶ ; a typical
value of I o is of order 100 W/m2. Then a convenient 2. Magnetic traps
expresssion for the change in energy of the atom in the The magnetic analog of Earnshaw’s theorem forbids
laser field is, in the limit ⌫Ⰶ⌬, the magnitude of the magnetic field B(r) to have a local

⌬E laser 共 r兲 ⫽ 冉 冊
I 共 r兲 ⌫ 2
Io ⌬
. (2.3)
maximum in free space. However, nothing forbids the
occurrence of a local minimum, and various methods
can be used to provide such a minimum, the most widely
Note that in this formula I/I o can be larger than 1 [pro- used being variants of the ‘‘time-orbiting potential’’ and
vided it is Ⰶ(⌬/⌫) 2 ]. A region of high laser intensity Ioffe-Pritchard traps; for a detailed description of these
thus provides an attractive potential for ⌬⬍0 (‘‘red de- the reader is referred to the papers of Petrich et al.
tuning’’) and a repulsive potential for ⌬⬎0 (‘‘blue de- (1995) and of Pritchard (1983), respectively. Virtually all
tuning’’). It should be borne in mind that by arranging to ‘‘pure’’ (i.e., non-laser-assisted) magnetic traps used in
have two counterpropagating laser beams, the potential BEC experiments to date have had axial symmetry and
can be varied over a scale of as short as half the laser a finite offset field, i.e., with an appropriate choice of
wavelength, i.e., ⬃3000 Å. cylindrical polar coordinate system the magnitude of the
An important question is the degree to which, if at all, field has the form
laser-generated potentials are sensitive to the hyperfine-
1 1
Zeeman index discussed in Sec. II.C. To the extent that 兩 B共 r兲 兩 ⫽B o ⫹ ␣ ␳ 2 ⫹ ␤ z 2 . (2.4)
spin-orbit coupling in the excited state is neglected, it is 2 2
clear that there can be no effect at all, since the orbital One might wonder why I have not specified the direction
ground state (s state) is unique. Moreover, for a linearly of the field as a function of r. The reason is that even
polarized laser beam the two electron-spin states are when this direction varies appreciably from its value at
clearly equivalent by time reversal, so any effect would the origin, the atoms move so slowly that it is an excel-
have to be at most of the order of the ratio of hyperfine- lent approximation to regard their magnetic moments as
Zeeman splitting to the detuning, which in most experi- following the direction of the local field adiabatically
ments is ⱗ10⫺4 (see below). The case of a circularly (see Sec. V.D). Thus, if we consider a given hyperfine-
polarized laser beam is more complicated, and by tuning Zeeman species, its potential energy will be a function
fairly close to the (fine-structure-split) resonance a con-
siderable sensitivity to the hyperfine index can be ob-
tained; see Corwin et al., 1999. 2
However, the detuning is usually still small enough relative
An important consideration in laser trapping is that to the resonance frequency that the counter-rotating terms
one usually wishes as far as possible to avoid spontane- (see Cohen-Tannoudji, op. cit.) can be neglected in the analy-
ous emission processes [note that a single optical pho- sis.

Rev. Mod. Phys., Vol. 73, No. 2, April 2001


Anthony J. Leggett: Bose-Einstein condensation in the alkali gases 311

only of the local field magnitude (2.4); in fact in the


simplest case it will be proportional to it with a constant
of proportionality ␮ (see below).

3. Gravity
For atoms in the nK– ␮K temperature regime, the ef-
fect of the Earth’s gravitational field is by no means neg-
ligible: for a 87Rb atom at 100 nK the thermal energy
corresponds to a vertical displacement in this field of
only about 1.6 ␮, less than the extent of the thermal
cloud in a typical trap in the absence of gravity. The
effect of gravity is, crudely speaking, to shift the mini-
mum of the potential in the vertical direction; in the case
of a pure laser trap, where the original potential is
nearly independent of the hyperfine-Zeeman species, FIG. 1. The energies of the different hyperfine-Zeeman states
this effect is not of great importance, but in the case of as a function of magnetic field.
magnetic trapping the species-independent gravitational
effect competes against a species-dependent magnetic
force, and the effect is in general to displace the mini- mF F E共 B 兲
mum of the potentials for different species relative to 2 2 A 共 1⫹B/B hf 兲

再冎
one another, an effect sometimes called ‘‘sag.’’ How-
ever, in certain cases it is possible to eliminate this effect 2
1 ⫾A 关 1⫹B/B hf ⫹ 共 B/B hf 兲 2 兴 1/2
by a judicious choice of field (see below). 1

C. The hyperfine and Zeeman interactions 0 再冎 2


1
⫾A 关 1⫹ 共 B/B hf 兲 2 兴 1/2

For a general account of this subject, I refer the


reader to Woodgate (1970), Chap. 9. Quite generally, a
⫺1 再冎 2
1
⫾A 关 1⫺B/B hf ⫹ 共 B/B hf 兲 2 兴 1/2)
hyperfine-Zeeman3 sublevel of an atom with given total ⫺2 2 A 共 1⫺B/B hf 兲 , (2.5)
electronic angular momentum J and nuclear spin I may
be labeled by the projection m F of total atomic spin F where the plus sign corresponds to F⫽2 and the minus
⬅I⫹J on the axis of the field B (which is a good quan- sign to F⫽1. A graph of these eigenvalues versus B is
tum number for any 兩 B兩 ) and by the value of total shown in Fig. 1; note in particular (1) the inversion of
F 关 F2 ⬅F(F⫹1) 兴 which characterizes it in the limit 兩 B兩 the order of the energies as a function of m F in the
→0: F takes value from 兩 I⫺J 兩 to I⫹J. In the present lower multiplet relative to the upper one; (2) the non-
context, I shall specialize immediately to the electronic monotonic behavior of the two m F ⫽⫺1 states; and (3)
ground state of the alkalis (J⫽S⫽1/2) and approximate the fact that (within this approximation) the initial slope
the electronic-spin g factor g s by 2; moreover, since the of E(B) is identical for the F⫽2, m F ⫽1 and F⫽1, m F
only experiments to date involving more than one hy- ⫽⫺1 states.
perfine species have been done on 87Rb or 23Na, I shall Most BEC experiments have been done with fields
specialize to the I value characterizing both of these that are much less than B hf in the relevant region of
(and also 7Li), namely, 3/2 (so F⫽1 or 2). Finally, for space, and it is then usually legitimate, to a first approxi-
the moment I shall neglect the small nuclear Zeeman mation, to linearize Eqs. (2.5) in B:

再 冎
energy. Then the energies of the various hyperfine sub-
levels are given as a function of the magnetic-field mag- 1
nitude B by the appropriate special case of the Breit- E 共 B 兲 ⬵⫾ A⫹ 兩 ␮ B 兩 m F B , (2.6)
2
Rabi formula [see Woodgate, 1970, Eq. (9.80)]. It is
convenient to choose the zero of energy to be the mean with the ⫹ (⫺) sign referring to the upper (lower) mul-
of the B⫽0 F⫽1 and F⫽2 energies and to define the tiplet. We see that in a field configuration of the form
zero-field splitting E(F⫽2)⫺E(F⫽1) as 2A. A has the (2.4), the states F⫽2, m F ⫽⫺2,⫺1 and F⫽1, m F ⫽0,
sign of the nuclear g factor and is positive for all the I ⫺1 will be expelled from the trap (or in the case F⫽2,
⫽3/2 alkali isotopes. We also introduce a characteristic m F ⫽⫺1 displaced to the locus B(r)⫽B hf /&); these
hyperfine ‘‘crossover’’ field B hf by B hf ⬅A/ 兩 ␮ B 兩 ( ␮ B states are usually called ‘‘high-field seekers.’’ On the
⫽eប/2m e ). Then the energies of the various levels are other hand, the states F⫽2, m F ⫽ ⫺2, 1 or 0 and F⫽1,
given as follows: m F ⫽⫺1 (in the limit BⰆB hf ) are ‘‘low-field seekers’’
and will be attracted to the origin.
Now, if we start with a gas of pure F⫽2, m F ⫽1 or
3
In the following I use ‘‘hyperfine’’ as a shorthand for F⫽2, m F ⫽0 atoms, it turns out that even in the neglect
‘‘hyperfine-Zeeman’’; thus two different hyperfine species may of dipolar forces (which do not in general conserve the
differ in the values of F and/or m F . total m F of the atoms involved), two-body collisions can

Rev. Mod. Phys., Vol. 73, No. 2, April 2001


312 Anthony J. Leggett: Bose-Einstein condensation in the alkali gases

produce lower-energy, high-field-seeking states, and for The most important single point to appreciate is that
this reason most experiments to date on single hyperfine the energy interval over which both the real and the
species in magnetic traps have been on one of the two imaginary parts of the dielectric constant of an atomic
‘‘maximally stretched’’ states, namely, F⫽2, m F ⫽2 and gas show substantial variation, namely, the linewidth ⌫,
F⫽1, m F ⫽⫺1. However, multispecies experiments have is typically of order 0.5 mK (a few MHz) and hence
involved other states, in particular F⫽2, m F ⫽1. We see small compared to the zero-field hyperfine ground-state
that given a field configuration of the form (2.4) with splitting (100 mK–1 K). Thus there is no difficulty in
BⰆB hf , the potential felt by an atom in a given hyper- measuring by optical means not only the total density of
fine state is in our approximation expressed in the form the gas but the density of atoms in the F⫽2 and F⫽1
multiplets separately. If one wishes to distinguish differ-
1 1
V 共 ␳ ,z 兲 ⫽const⫹ M ␻ r2 ␳ 2 ⫹ M ␻ 2z z 2 , (2.7) ent sublevels within each multiplet, the energy resolu-
2 2 tion alone may or may not be sufficient, but one can
where the value of ␻ r ⬅2 ␲␯ r is identical for the F⫽2, exploit the sensitivity of a transition out of a particular
m F ⫽1 and F⫽1, m F ⫽⫺1 states but a factor of & sublevel to the polarization of the probe laser to make
larger for the F⫽2, m F ⫽2 state (and similarly for ␻ z the distinction. In most, though not all, experiments the
⬅2 ␲␯ z ). Experimental papers reporting work on a quantity that is directly measured is either the total
single hyperfine species commonly specify the trap in atomic density or that of a particular species (usually
terms of the relevant values of ␯ r and ␯ z . In practice, integrated along the line of the probe laser beam) as a
typical values of each lie in the region of a few Hz to a function of space and/or time.
few kHz; there is no generic constraint on the ratio The simplest optical imaging technique relies on
␯ r / ␯ z , which may, depending on the trap, be Ⰷ1, ⬃1, or straight absorption: the logarithm of the intensity trans-
Ⰶ1. mitted through a column of gas is simply proportional to
Although Eq. (2.5) [and hence Eq. (2.6)] is usually an the integrated atomic density (or the density of the par-
adequate approximation for BⰆB hf , it is necessary on ticular species selected). This method is inherently de-
occasion to go beyond it. A case in point is the calcula- structive (since real absorption processes are followed
tion of the relative equilibrium positions, in the presence by spontaneous radiation and the accompanying heat-
of gravity, of the F⫽2, m F ⫽1 and F⫽1, m F ⫽⫺1 states. ing), and in addition is difficult to apply at high densities
The position of the minimum in each case is determined (see Andrews et al., 1996, p. 85, paragraph 2); thus in
by balancing the Earth’s gravitational field against the practice one often switches off the trap and allows the
gradient of the magnetic potential, and if we were to use sample to expand (and thus rarefy) before switching on
Eq. (2.5) it would coincide for the two species. However, the probe laser beam.
at this point it is necessary to take into account both the A second, nondestructive measurement technique is
nuclear Zeeman term and the term of order B 2 in the dispersive (phase-contrast) imaging; this relies on the
expansion of formulas (2.5) (the quadratic Zeeman ef- diffracting effect of the gas and does not involve (much)
fect); it turns out that these two small effects exactly heating of the sample; it need not therefore be destruc-
cancel when B has the special value (4/3) ␮ n A/ ␮ B 2 tive, and it is believed that up to ⬃100 successive imag-
(⬃1G). A second respect in which the quadratic Zee- ings may be obtained by this method (Andrews et al.,
man effect plays a useful role is that by making the fre- 1996). The spatial resolution obtainable is typically of
quencies for (say) the m F ⫽1→m F ⫽0 and m F ⫽0→m F the order of a few microns; as to the time resolution, it is
⫽⫺1 transitions unequal it permits selective population, apparently limited only by the switching-on time of the
by an ordinary rf pulse acting on an originally purely laser. As we shall see, this time is many orders of mag-
m F ⫽⫺1 population, of the m F ⫽0 state without appre- nitude smaller than the typical time scales of the dynam-
ciable population of the m F ⫽1 state—something which ics of the system, so optical measurements of density
is not possible so long as the (hyperfine) Zeeman energy may reasonably be regarded as instantaneous.
is purely linear in m F (on this, see Stenger et al., 1998b,
Sec. V). E. Orders of magnitude

To aid our qualitative understanding, it is useful to


compare typical orders of magnitude of various relevant
D. Imaging energies for a set of ultracold atoms in a single system of
units. In Table I, for definiteness we choose 106 87Rb
The raw data from which we infer the static and dy- atoms in a ‘‘typical’’ magnetic trap; numbers for the
namic behavior of an ultracold trapped alkali gas, and in other heavier alkalis are generally comparable, but for
particular of a Bose-Einstein condensate, is almost with- H may be different by up to two or three orders of mag-
out exception optical, involving the fundamental visible- nitude. In discussing the physical significance of some of
region ns→np transition. I shall treat this subject quite these quantities I anticipate some of the results to be
briefly, simply to indicate broadly what kind of informa- obtained in Secs. III–V.
tion can be obtained and with what order of accuracy. It is worthwhile to note explicitly that the first three
For further details I refer the reader to Ketterle et al. quantities in Table I, and the recoil energy, are charac-
(1999). teristics of the isolated atom and do not depend on the

Rev. Mod. Phys., Vol. 73, No. 2, April 2001


Anthony J. Leggett: Bose-Einstein condensation in the alkali gases 313

TABLE I. Energies for a set of 106 ultracold 87


Rb atoms in a typical magnetic trap.

Temperature units Frequency units

Energy of ns→np transition ⬃2⫻104 K 4⫻1014 Hz


Zero-field hyperfine splitting ⬃0.3 K ⬃7 GHz
Energy width of transition (ប⌫) ⬃0.3 mK ⬃6 MHz
Characteristic two-body energy ប 2 /ma s2 ⬃0.3 mK ⬃6 MHz
Transition temperature k B T c ⬃500 nK ⬃10 kHz
Recoil energy ប 2 k opt
2
/2M ⬃200 nK ⬃5 kHz
Mean-field energy nU o ⬃100 nK ⬃2 kHz
Zero-point energy in harmonic well ⬃5 nK ⬃100 Hz

conditions of confinement. The two-body energy ប 2 /ma s2 particle spacing, which is about 3.5 Å; we see that the
can be tuned by adjusting the magnetic-field strength so dilute-gas condition a s Ⰶr int which is so characteristic of
as to vary the scattering length a s (see Sec. IV.B) but is the BEC alkali gases is very far from satisfied for liquid
otherwise insensitive to the details of the trap geometry. helium. As a consequence, liquid helium is, in an intui-
By contrast, not only the zero-point energy ប ␻ o but the tive sense, a much more strongly interacting system than
transition temperature and the mean-field energy de- the BEC gases, by many orders of magnitude, and this
pend on the shape of the confining potential, as well as leads to a number of profound differences between the
(in the latter two cases) on the total number of particles two systems. In particular, (1) in the alkali gases the
trapped; as we shall see in Sec. V.E, for a given species behavior is extremely sensitive to the details of the trap,
whereas in helium it is dominated by the interparticle
with a s fixed, k B T c is proportional to N 1/3␻ o , while typi-
interactions and very insensitive to the confining poten-
cal values of the mean-field energy in the condensed
tial (which is in any case typically due to a fairly rigid
phase are proportional to N 2/5␻ o6/5 . Thus variation in N, ‘‘box’’ and thus flat over most of the region of interest);
in particular, can result in substantial variations in these (2) in helium, interatomic collisions are so frequent that
two quantities around the typical values quoted. Never- any process that is energetically allowed takes place vir-
theless, under realistic conditions with ‘‘typical’’ values tually instantaneously, whereas in the alkali gases the
of a s (⬃50– 100 Å; see Sec. IV.B) we almost always kinetics of the process may be an important bottleneck
have the set of inequalities (see Sec. VI.E); (3) on the theoretical front, quantitative
calculations based on perturbation theory in the (effec-
ប 2 /ma s2 Ⰷk B T c ⰇnU o Ⰷប ␻ o , (2.8)
tive) interatomic interaction are usually believed to be
although in the case of the second inequality the Ⰷ may highly reliable for the alkali gases, whereas for helium
represent a ratio that is only of order 5 (see Table I). they may fail miserably. There are other important dif-
It is useful also to note some length scales character- ferences between the two systems, in particular, that in
istic of the BEC alkali-gas problem, and their relative the alkali gases both the external and, at least in prin-
orders of magnitude. Apart from the s-wave scattering ciple, the interatomic potentials can be adjusted over a
length a s , we define the (typical) mean interparticle time scale very short compared to the time scales char-
spacing r int , the ‘‘healing length’’ ␰ ⬅(2mnU o ប 2 ) ⫺1/2 acterizing the dynamics of the system: this permits types
(see Sec. V.A), the thermal de Broglie wavelength at of experiment, such as the celebrated interference ex-
T c , ␭ DB , and the oscillator length a ho ⬅(ប/m ␻ o ) 1/2, periment of Andrews et al. (1997; see Sec. VII.E), that,
while as it were conceptually equally viable in superfluid
which is the zero-point spread of the ground-state wave 4
He, would in practice be totally impossible to realize.
function of a free particle in the trap in question. Under
For a more detailed discussion of the analogies and dif-
normal conditions, we then have the inequalities
ferences between the superfluid phases5 of 4He and 3He
a s Ⰶr int ⬃␭ DBⱗ ␰ Ⰶa ho . (2.9) on the one hand, and the BEC phase of the alkali gases
on the other, see Leggett (1999a).
Typical values might be a s ⬃50 Å, r int ⬃2000 Å, ␰⬃4000
Å, a ho ⬃1 ␮ .
III. THE DEFINITION, ORIGIN, AND OCCURRENCE
It is interesting to compare the above numbers with OF BEC: THE ORDER PARAMETER
those characteristic of liquid helium.4 In the case of he-
lium at liquid densities, the concept of an s-wave scat- A. Definition of BEC
tering length is not really meaningful, but a somewhat
similar role is played by the characteristic dimension of For pedagogic convenience let us start with the case in
the hard-core part of the interatomic potential, which is which the hyperfine degree of freedom can be ignored.
around 2.5 Å. This should be compared with the inter-

5
The Fermi system 3 He is believed to become superfluid by
4
The numbers quoted are for 4 He at saturated vapor pres- forming Cooper pairs, which then effectively undergo BEC;
sure; the numbers at pressures up to freezing, and for the light the pairs are somewhat analogous to the alkali atoms in that
isotope 3 He, are similar in order of magnitude. they possess a hyperfine degree of freedom.

Rev. Mod. Phys., Vol. 73, No. 2, April 2001


314 Anthony J. Leggett: Bose-Einstein condensation in the alkali gases

We thus consider a system of N identical spinless bosons if one and only one eigenvalue is of order N, all the rest
characterized by spatial coordinates ri (i ⫽ 1, 2•••N), being of order 1. Systems showing nonsimple BEC (i.e.,
with arbitrary interparticle interactions and subject to having more than one eigenvalue of order N) are some-
some external potential, possibly time dependent, which times said to be fragmented. The phrase ‘‘of order
for technical reasons we shall assume is such as to con- N(1)’’ is somewhat vague in a situation where there is
fine the particles in some finite region of space. We do no simple thermodynamic limit, but in practice this does
not assume that the system is necessarily in thermal not usually lead to difficulty.
equilibrium, nor even in a steady state. The many-body In the case of simple BEC, we shall arbitrarily use the
wave functions ⌿ N (r1 r2 ¯rN :t) must be symmetric with value zero of the index i to refer to the unique state
respect to the interchange ri
rj of any two particle co- which has n i (t)⬃N; we shall call the single-particle state
ordinates. ␹ 0 (r:t) the condensate wave function and the eigenvalue
N 0 (t) (where we use the capital to emphasize the mac-
For any given time t we can define the one-particle
roscopic value) the ‘‘(mean) number of particles in the
reduced density matrix ␳ (r,r⬘ :t) in the standard way.
condensate.’’ A very important quantity associated with
We may as well go directly to the most general case, in
␹ 0 , the superfluid velocity vs (r,t), will be introduced in
which the description of the system is by a statistical Sec. III.D.
mixture of mutually orthogonal many-body states ⌿ (s) N One point concerning the above definition is very im-
with probability p s : the definition is portant: In general, a statistical mixture of many-body

兺s p s 冕 dr2 ¯drN ⌿ N* (s)


states, each of which separately possesses (simple or
␳ 共 r,r⬘ :t 兲 ⬅N general) BEC, need not itself possess BEC. (Consider,
for example, a mixture of p⬃N many-body states, each
of which has a condensate in a different one of p mutu-
N 共 r⬘ r2 ¯rN :t 兲
⫻ 共 rr2 ¯rN :t 兲 ⌿ (s)
ally orthogonal states ␹ 0 .) On the other hand, no mix-
⬅ 具 ␺ˆ † 共 rt兲 ␺ˆ 共 r⬘ t 兲 典 , (3.1) ture of many-body states that do not individually show
BEC can itself show BEC.
where in the last expression the quantity ␺ˆ (r) is the Finally, let us generalize the above definitions to take
standard boson field operator, and the average indicated account of the hyperfine degree of freedom. We now
by the pointed brackets is in general statistical as well as characterize the ith particle not just by its spatial coor-
quantum mechanical. Because of the Bose symmetry, dinate ri but by a discrete hyperfine index ␣ i , so that ⌿ N
the fact that in writing the second expression in Eq. (3.1) is a function of 兵 ri , ␣ i 其 , (i⫽1,2¯N). The appropriate
we have arbitrarily picked out the coordinate r1 as ‘‘spe- generalization of Eqs. (3.1) and (3.2) is then [omitting
cial’’ need not worry us. the intermediate expression in Eq. (3.1) whose generali-
It follows from the definition (3.1) that the quantity zation is obvious, and writing the field operator for hy-
␳ (r,r⬘ :t), when regarded as a matrix function of its indi- perfine state ␣ as ␺ˆ (r, ␣ ) rather than the more conven-
ces r and r⬘ , is Hermitian, and can therefore be diago-
tional notation ␺ˆ ␣ (r)]
nalized with real eigenvalues. That is, it is always pos-
sible to find a complete orthonormal basis, in general ␳ 共 r␣ ,r⬘ ␣ ⬘ :t 兲 ⬅ 具 ␺ˆ † 共 r, ␣ 兲 ␺ˆ 共 r⬘ , ␣ ⬘ 兲 典
time dependent, of single-particle eigenfunctions ␹ i (r;t)
such that we can write
⫽ 兺i n i共 t 兲 ␹ i*共 r, ␣ :t 兲 ␹ i共 r⬘ , ␣ ⬘ :t 兲 . (3.4)
␳ 共 r,r⬘ :t 兲 ⫽ 兺i n i共 t 兲 ␹ *i 共 r:t 兲 ␹ i共 r⬘ :t 兲 . (3.2)
The definition of (simple and general) BEC is then ex-
actly as above, the only difference being that the con-
It is important to note that in the general case (a) not densate wave function ␹ 0 (r, ␣ :t) now has a discrete (hy-
only the eigenfunctions ␹ i appearing in Eq. (3.2) but perfine) argument ␣ as well as a continuous (position)
also the eigenvalues n i may be functions of time; (b) the one r, i.e., it can be regarded as a spinor. We shall see
␹ i need not be eigenfunctions of any particular quantity below (Sec. V) that the BEC occurring in a situation
[other than ␳ (rr⬘ :t) itself] and in particular are not nec- where the hyperfine degree of freedom is important is
essarily eigenfunctions of the single-particle terms in the often not of the simple variety.
Hamiltonian; and (c) if we define operators a i by

a i共 t 兲 ⬅ 冕␺ˆ 共 r兲 ␹ *
i 共 r:t 兲 dr, (3.3) B. Why BEC?

then while the n i (t) are the expectation values of the Whether or not BEC occurs in a given Bose system
operators a †i (t)a i (t), the many-body wave function is depends strongly on the sign of the ‘‘effective’’ inter-
not in general an eigenfunction of the latter operator. atomic interaction (a concept that will be defined pre-
We are now in a position to formulate a definition of cisely for the alkali gases in Sec. IV). The case of attrac-
Bose-Einstein condensation (BEC). We shall say that at tive interaction is rather subtle and is discussed briefly in
any given time t, the system shows BEC if one or more Sec. IX; here we confine ourselves to the case of repul-
of the eigenvalues n i (t) is of the order of the total num- sive (or zero) interaction. For such a case the tendency
ber of particles N; and further that it shows simple BEC of the system to undergo BEC, which is by no means

Rev. Mod. Phys., Vol. 73, No. 2, April 2001


Anthony J. Leggett: Bose-Einstein condensation in the alkali gases 315

confined to states of (or close to) thermal equilibrium, is sion D⬎2, and in a harmonic trap (isotropic or not)8 for
a consequence of two mutually reinforcing effects, sta- D⬎1; see, for example, Dalfovo et al. (1999) or Pethick
tistical and energetic. I discuss these in turn. and Smith (2000). For each of these cases the transition
The statistical consideration is a standard textbook temperature T c can of course be calculated exactly as a
subject (though its generic significance is often some- function of N and the geometric parameters, but a
what obscured because it tends to be discussed in the simple rule of thumb, valid except very close to the bor-
specific context of a derivation of the equilibrium ther- derline dimension, is that BEC occurs when the degen-
mal distribution). Consider the problem of distributing eracy condition N⬃p is satisfied, with p taken to be the
N objects (atoms) among p different boxes (states). If number of thermally accessible states, i.e., those with
the objects are distinguishable (the classical case), the energies ⱗk B T. For example, in an anisitropic 3D trap
number of different ways of distributing them so that the with geometrical mean frequency ␻ ¯ 0 ⬅( ␻ r2 ␻ z ) 1/3, where
ith box contains n i objects is the multinomial coefficient the mean density of single-particle states per unit energy
p
N!/⌸ i⫽1 n i !. If on the other hand the objects in question E is E 2 /(ប ␻
¯ 0 ) 3 , this argument gives for large N
are indistinguishable but there are no other constraints
¯ 0 •N 1/3.
k B T c ⫽const ប ␻ (3.7)
(the boson case), then clearly there is one and only one
way of distributing them, for a given set of 兵 n i 其 . The A quantitative calculation (see the cited references) con-
difference becomes important when Nⲏp (the degen- firms this result, with the constant fixed at 0.94.
eracy condition), and its effect, crudely speaking, is that For a noninteracting gas in any geometry in which
in the bosonic case states in which many particles occupy BEC occurs, the condensate number N 0 (T) increases
the same state have a higher relative weight than in the smoothly from zero at T c to the total number N as T
classical (distinguishable) case.6 This tendency of bosons falls to zero. In particular, for the above anisotropic 3D
to ‘‘cluster’’ is very generic; it is in no way restricted to trap we find
thermal equilibrium, or to the noninteracting case.
N 0 共 T 兲 ⫽N 关 1⫺ 共 T/T c 兲 3 兴 . (3.8)
If, however, we specialize for the moment to the case
of noninteracting bosons in thermal equilibrium, and For a comparison of this (noninteracting-gas) formula
moreover are content to use the standard grand canoni- with experiments in the alkali gases, see Hau et al.
cal ensemble,7 the results take a very simple and stan- (1998).
dard form [see, for example, Huang (1987)]: at tempera- In addition to the above effect of statistics, there is
ture T the formula for the mean number of particles n i also an energetic effect: In a dilute Bose gas with weakly
in the single-particle energy eigenstate i with energy ⑀ i repulsive interactions, the interactions tend to reinforce
has the standard Bose-Einstein form the effect of statistics in forming BEC. To see this, we
need to anticipate a result to be proved in Sec. IV,
n i ⫽ 兵 exp关 ␤ 共 ⑀ i ⫺ ␮ 兲 ⫺1 兴 其 ⫺1 ␤ ⬅1/k B T, (3.5) namely that, at least so long as we stay in the dilute
limit, the Hartree-Fock energy of two identical spinless
where the chemical potential ␮ is fixed by the condition bosons in different orbital states is greater than that of
two such bosons in the same state, by a factor between 1
and 2. This effect not only helps us understand why even
兺i n i共 ␮ , ␤ 兲 ⫽N, (3.6) in nonequilibrium situations fragmentation (i.e., an
order-N occupation of two or more single-particle or-
bital states) is unusual in a spinless system; rather gen-
N being the average number of particles in the grand erally it says that, other things being equal, the more
canonical ensemble. repulsive the interatomic interaction (i.e., the greater
The quantity ␮ ( ␤ :N)⬅ ␮ (T,N) implicitly defined by the s-wave scattering length; see Sec. III), the stronger
Eq. (3.6) is very large and negative for T→⬁; as T falls the tendency to condense. (However, at some point this
it increases monotonically. If at some temperature T c tendency will be balanced by the coherent scattering of
condition (3.6) can be met with ␮ →0 ⫺ , then below T c pairs out of the condensate; see Sec. VIII.)
the occupation of the lowest single-particle state (energy It is plausible that in the less common case of a nega-
⑀ 0 ⫽0 by convention) is of order N, while the other n i ’s tive scattering length (attractive interaction) condensa-
are still generally of order unity or less, i.e., BEC is re- tion is likely to be inhibited, and we return to this point
alized. This condition is met, in free space, for dimen- briefly in Sec. IX.

6
C. Rigorous results
This tendency may be seen already in the trivial case N⫽p
⫽2: for distinguishable particles, states involving double occu- In the realistic case of an interacting Bose system in
pation are 50% of the whole, while for bosons they are 66%.
7 more than one dimension, rigorous results concerning
Because of the critical role in the theory of BEC played by
conservation of total particle number N, it is not entirely ob-
the existence and degree of BEC as a function of par-
vious a priori that the use of the grand canonical ensemble is ticle number, strength of interaction, etc. are rather few
justified. However, calculations using the microcanonical en-
semble give similar results for large N (see, for example, Gajda
8
and Rzazewski, 1997). The case D⫽1 is rather delicate; see Dalfovo et al., 1999.

Rev. Mod. Phys., Vol. 73, No. 2, April 2001


316 Anthony J. Leggett: Bose-Einstein condensation in the alkali gases

and far between (see Lieb, 1999) and to the best of my D. The order parameter and the superfluid velocity
knowledge are at present confined entirely to extended
systems in the thermodynamic limit. Of course, many
readers will no doubt take the view that in the context of 1. Possible definitions of the order parameter
the real-life alkali gases such results are of rather minor
For the purposes of the present section, let us con-
interest, since both experiment and theoretical prejudice sider a spinless system9 and assume that simple BEC is
suggest that perturbation theory starting from the non- realized in the state ␹ 0 (r:t), with eigenvalue N 0 (t).
interacting gas is likely to be a reliable description. Nev- Then the simplest and most direct definition of the order
ertheless I feel it is worthwhile to summarize without parameter ⌿(r,t) [do not confuse with the many-body
derivation the few rigorous results known to me for the wave function ⌿ N (r1 ¯rN :t)] is
extended case; most of them probably could be (but to
my knowledge have not been) generalized to a realistic
⌿ 共 rt 兲 ⬅ 冑N 0 共 t 兲 ␹ 0 共 r:t 兲 , (3.12)
trap geometry.
(a) For a 3D system in free space, Gavoret and No- that is, apart from its normalization ⌿(rt) is simply the
zières (1964) showed many years ago that if perturba- (Schrödinger) wave function of the single-particle state
tion theory starting from the noninteracting Bose gas into which condensation occurs. Contrary to widespread
converges, then at T⫽0 the system displays BEC. How- popular belief, this definition [which is to all intents and
ever, their argument allows no inference about the con- purposes the one originally given by Penrose and On-
densed fraction N 0 (0)/N, nor about the critical tem- sager (1956)] is perfectly adequate for all the purposes
perature T c . for which the concept of an order parameter has been
(b) If the real-life continuum is replaced by a ‘‘lattice (correctly) used in the literature. One point needs spe-
gas’’ and the interatomic interaction modeled by a hard- cial emphasis: Since ⌿(rt) is in effect simply a Schrö-
core on-site repulsion, then for the case of half filling the dinger wave function, it is clear that while its space and
existence of BEC at T⫽0 has been proved (Kennedy time derivatives (and in the case of a hyperfine degree of
et al., 1988). freedom the phase relations between its hyperfine com-
(c) The best-known theorem concerning the (non)oc- ponents) are meaningful physical quantities, the overall
currence of BEC in extended D-dimensional space is phase of the order parameter has no physical significance.
due to Hohenberg (1967). Denoting by n k the average Various alternative definitions of the order parameter
number of particles per unit volume in the plane-wave are to be found in the literature. Some, such as the defi-
state k(⫽0), he demonstrates rigorously for any dimen- nition explicitly based on ‘‘off-diagonal long-range or-
sion the inequality der’’ (Penrose and Onsager, 1956; Yang, 1962), are con-
ceptually unexceptionable but obviously inapplicable in
a trap geometry. However, there is one definition that is
n k ⭓ 共 n 0 /n 兲共 mk B T/ប 2 k 2 兲 ⫺1/2 共 n⬅N/V 兲 . (3.9) sufficiently common in the alkali-gas BEC literature that
It then follows (Hohenberg, 1967) that for D⭐2, in the it calls for comment, namely, that based on the idea of
thermodynamic limit, BEC cannot occur at any finite so-called ‘‘spontaneously broken gauge symmetry.’’ I re-
temperature. fer the reader to Anderson (1966), Leggett and Sols
(d) While for D⫽3 Hohenberg’s inequality places (1991), and Leggett (1995a) for an extended discussion
no upper limit on T c , it does place a limit on the con- of this concept and its difficulties, and just summarize
densate fraction N 0 (T)/N, which is nontrivial for T the basic idea here: One imagines that the superselec-
ⲏT (0)
c , where T c
(0)
is the transition temperature of the tion rule for the total particle number N is somehow
noninteracting gas. A simple form (not the tightest at- violated, so that one can write the wave function of the
tainable) of this limit is system as a superposition of states corresponding to dif-
ferent N:

N 0 共 T 兲 /N⭐ ␣ 共 T C
(0)
/T 兲 , ␣⫽ 冋 3 ␨ 共 3/2兲 1/2
4
␲ 册 2/3
⬵2•3.
⌿⫽ 兺N a N ⌿ N . (3.13)
(3.10)
This limit is completely independent of the nature and This allows the single-particle destruction operator
even the sign of the interatomic potential. ␺ˆ (rt) to have (possibly) a finite expectation value, and
(e) Finally, for a Bose gas with an interaction that is one then identifies the order parameter with the latter:
everywhere repulsive and has a space integral V 0 , it is
possible to place an upper limit on the condensate frac-
⌿ 共 rt 兲 ⬅ 具 ␺ 共 rt 兲 典 . (3.14)
tion which for small V 0 is stronger than Eq. (3.10) (Leg-
gett, 2000b). For the special case T⫽T (0)
c this reads It should be emphasized that there are no circum-
stances in which Eq. (3.13) is the physically correct de-
N 0 /N⭐const共 nV 0 /k B T (0)
c 兲 ,
1/3
(3.11)
where the current upper bound on the constant is about 9
The generalization to systems with internal degrees of free-
2.5. dom is made in Sec. V.D.

Rev. Mod. Phys., Vol. 73, No. 2, April 2001


Anthony J. Leggett: Bose-Einstein condensation in the alkali gases 317

scription of the system,10 or even of a part of it, and in IV. THE EFFECTIVE INTERACTION IN A COLD
the present author’s opinion the definition (3.14), while DILUTE GAS
possibly streamlining some calculations when judiciously
used, is liable to generate pseudoproblems and is best A. Statement of the problem: neglect of l⫽0 partial waves
avoided. However, the reader should be warned that
this opinion is controversial and that there are even In this section I consider the effect of atom-atom in-
those who feel that Eq. (3.14) is not only a possible, but teractions in a gas at the very low temperatures and den-
the only legitimate, definition of the order parameter. sities characteristic of the alkalis under BEC conditions;
for a much more detailed treatment of this topic, see
2. The superfluid velocity Dalibard (1999) and Weiner et al. (1999). For pedagogi-
The simplest definition of the superfluid velocity cal simplicity I assume, in Secs. I–III, that two colliding
vs (r,t) is directly in terms of the phase gradient of the atoms can be ‘‘tagged’’ (as would be the case if they
condensate wave function ␹ 0 (rt): writing ␹ 0 (rt) were of different chemical and/or isotopic species); this
⬅ 兩 ␹ 0 (rt) 兩 exp关i␸(rt) 兴 , we define will allow us to postpone discussion of the effects of
indistinguishability to Sec. IV.D. However, it will turn
ប out that none of the results of Secs. IV.A–IV.C depend
vs 共 r,t 兲 ⬅ “ ␸ 共 rt 兲 . (3.15)
m on this assumption. Furthermore, we neglect for the mo-
Clearly, since N 0 is by definition not a function of posi- ment the hyperfine degree of freedom, assuming that for
tion, we could equally well have defined vs in terms of each atom its initial and final values in a collision are
the phase of the order parameter ⌿(rt)⬅ 冑N 0 ␹ 0 (rt). equal. The only relevant variable, for a given pair of
We see directly from its definition that vs satisfies two atoms, is then their relative coordinate r, and the out-
important constraints, namely, the condition of irrota- come of a collision will be determined by the value of
tionality the (initial) relative kinetic energy E⬅ប 2 k 2 /2m r (here
and subsequently m r denotes the reduced mass and k
curl vs 共 r,t 兲 ⬅0 (3.16)
the relative wave vector).
and the celebrated Onsager-Feynman quantization con- Consider the true interatomic potential V at (r) as a
dition function of the separation r of the two centers of mass

冖 vs 共 rt 兲 •dl⫽nh/m 共 n integral兲 . (3.17)


(c.m.). At short distances, of the order of molecular di-
mensions, this potential may not even be definable
(since the standard Born-Oppenheimer separation of
Obviously, in a simply connected geometry with ␹ 0 (rt) the c.m. and internal degrees of freedom may break
finite everywhere Eq. (3.16) implies Eq. (3.17) with n down, see Weiner et al., 1999), but at distances ⲏ5 Å
⫽0, but there are more general cases, e.g., involving vor- (beyond which the Born-Oppenheimer approximation
tices, in which n can be nonzero; see Sec. VI below. should certainly be good), V at (r) should be well defined
The reason the quantity defined by Eq. (3.15) is im- and well approximated11 in the limit r→⬁ by the lowest-
portant in the theory of superfluidity, and more gener- order van der Waals interaction ⫺C 6 /r 6 ; if we express
ally of BEC, is that on the one hand it reflects the prop- energies and lengths in the standard atomic units (bohrs
erties of a single quantum state [unlike the and hartrees, respectively), C 6 ranges from ⬃1400 for Li
hydrodynamic velocity v(rt)⬅j(rt)/ ␳ (rt), which is an to ⬃6300 for Cs (see, for example, Marinescu et al.,
average over many different states and thus fails in gen- 1994). This form of potential defines a characteristic van
eral to satisfy Eq. (3.16) or Eq. (3.17)], while on the der Waals length, namely, r 0 ⬅(2m r C 6 /ប 2 ) 1/4, the physi-
other hand, unlike the analogous quantity in single- cal significance of which is that it is the typical extent of
particle Schrödinger quantum mechanics, which while the last bound state in the potential; this length is of the
perfectly definable is subject to huge quantum fluctua- order of 50 Å, much larger than a typical molecular di-
tions, vs characterizes the behavior of a macroscopic mension, and the associated characteristic energy E c
number of particles (the condensate) and can thus in ⬃ប 2 /m r r 20 is of order 0.1– 1 mK. 12
effect be regarded as a classical quantity. (For further An important simplification of our problem results
discussion, see Leggett, 2000a.) from the fact that for all the alkalis (including H), the
Finally, I want to reemphasize that the definitions of values of thermal energy k B T characterizing BEC con-
the order parameter and of the superfluid velocity given ditions are small compared to the above energy. Since
in this subsection rely essentially on the assumption of
simple Bose condensation, i.e., that there is one and only
one eigenvalue of the single-particle density matrix of 11
Strictly speaking, in the limit r→⬁ the leading term in
order N. Generalizations to the case of multiple conden- V at (r) is the electromagnetic interaction between the electron
sates, while possible, need to be carefully defined. spins, which falls off only as r ⫺3 . However, for all the alkalis,
including H, the mean field due to this term may be verified to
be small compared with the ‘‘standard’’ mean field calculated
10
An interesting attempt to justify Eqs. (3.14) and (3.13) via below.
the concept of a ‘‘phase standard’’ has been made by Dunning- 12
This is one respect in which H is very different: C 6 ⬃6.5,
ham and Burnett (1998) and is discussed in Leggett (2000a). r o ⬃5 Å, E c ⬃3 K.

Rev. Mod. Phys., Vol. 73, No. 2, April 2001


318 Anthony J. Leggett: Bose-Einstein condensation in the alkali gases

for l⫽0 the probability of finding the two atoms at a of H is an exception in this respect: a s is positive even
distance r 0 from one another in a collision process falls though there are no bound states.
off as (kr 0 ) 2l , where l is the relative orbital angular mo- The value of the s-wave scattering length, which in
mentum, this means that for l⫽0 the effective scattering general is a function not only of the chemical and isoto-
amplitude, and hence the contribution to the effective pic species involved but of the hyperfine indices of the
interaction (see below) is smaller than that arising from two atoms (see Sec. IV.E) and even of the magnetic
s-wave scattering by a factor of order (k B T/E c ) l and and/or laser field, is a basic datum for the alkali-gas
thus is normally negligible. It follows that in our subse- problem. It may be obtained experimentally by the tech-
quent discussion we may legitimately restrict ourselves nique of photoassociative spectroscopy (see Tiesinga
to l⫽0 (s-wave) scattering. et al., 1996) or from a knowledge of the atomic mean
free path in the gas (see Sec. IV.C). Alternatively, for
the lightest alkalis one can attempt to calculate it from
first principles. A partial list of values found in the lit-
B. The s-wave scattering length erature is given (in atomic units) in Table III of Weiner
et al. (1999). Inspection of this table reveals two obvious
The theory of low-energy s-wave scattering of two dis-
features: Except for the case of H, a s is generally much
tinguishable particles interacting via a central potential
greater than a typical (vibrational-ground-state) molecu-
V(r) is a standard textbook subject; see, for example,
lar radius and is in fact typically of the order of the van
Landau and Lifshitz (1959), Sec. 108. The principal re-
der Waals length r 0 , and positive values are appreciably
sult we need in the present context is the following. Pro-
more common than negative ones. Neither of these fea-
vided V(r) falls off, as r→⬁, faster than r ⫺3 , then in
tures is an accident; see Gribakin and Flambaum (1993),
this limit and the limit k→0 (but for kr possibly ⬎1),
or for a concise version of the argument Pethick and
the s-wave scattering solution to the time-independent Smith (2000), Sec. 5.3.13
Schrödinger equation has, quite generically, the simple
Although the magnitude of a s is thus often large on an
form
atomic scale, it is very important for the theory of the
sin关 k 共 r⫺a s 兲兴 BEC alkali gases that under normal BEC conditions it is
␺ 共 r 兲 ⫽const , (4.1) still small compared to all the other characteristic
r
lengths L c (thermal de Broglie wavelength, interparticle
where the quantity a s is known as the (zero-energy) spacing, and zero-point length of the trap). In particular,
s-wave scattering length and, depending on the details of if we denote the density by n, the so-called gas param-
the potential, may have either sign. In the case of posi- eter na s3 , whose significance will become clear subse-
tive a s one can say that by comparison with the nonin- quently, is under normal BEC conditions at most of or-
teracting case the relative wave function is repelled from der 10⫺4 and often much smaller. Although in principle
the origin, whereas with negative a s the wave function is it is possible to make 兩 a s 兩 tend to infinity (so that na s3
attracted; in the repulsive case a s may be visualized as ⲏ1) by appropriate tuning of the magnetic field (Fesh-
the radius of the hard-sphere potential, which would bach resonance), in practice very large values of a s tend
give rise to the same relative wave function. However, it to lead to rapid three-body recombination (see Stenger
should be emphasized that in the real problem ␺ (r) et al., 1999). Very recent work, however, has indicated
does not vanish for r⬍a s . that such recombination may not be inevitable (Cornish
The general behavior of the scattering length is re- et al., 2000).
lated to the s-wave bound states occurring in the poten- In view of some considerations that will emerge in the
tial. If we imagine starting from the noninteracting state context of the Bogoliubov approximation (Sec. VIII), it
(a s ⬅0) and gradually increasing the strength of the po- is worthwhile to close this brief discussion with a note on
tential (whose shape is taken to be typical of a real the question of depletion. It is clear that even if the
atomic one), then a s will initially (or at any rate after a asymptotic value of the relative wave vector k is zero (as
little) take increasingly negative values. As the point at it will be if atoms are part of a spatially homogeneous
which the potential is just enough to sustain a bound Bose condensate), the short-range interatomic interac-
state is approached, a s will approach ⫺⬁, and when the tion will induce a finite probability p q of a nonzero rela-
state is just bound, will take a large positive value (which tive wave vector q. Taking for definiteness the case a s
approaches ⫹⬁ as the potential is reduced again to the ⬎0, a s Ⰷr 0 (where r 0 is a measure of the effective range
critical value). In this region (a s Ⰷr 0 ) the asymp- of the potential, e.g., the van der Waals length defined
totic form of the bound-state wave function is above), and normalizing in volume V, we find (most eas-
r ⫺1 exp(⫺r/as) and its energy is ⫺ប 2 /2m r a s2 ; the form of ily by requiring the zero-energy scattering state to be
the zero-energy scattering state may be viewed as a con-
sequence of the need to make it orthogonal to the
bound state. This general behavior is expected near the 13
It may be worthwhile to note that the zero-energy Schrö-
points where further bound states appear in the well; it dinger equation can be solved explicitly for any potential of
is worth bearing in mind that for the heavier alkalis the form ⫺ ␣ /r n (and in particular for the van der Waals po-
there are a large number of such l⫽0 bound states tential, n⫽6) in terms of Bessel and Neumann functions; see
(⬃120 for 87Rb; see Mies et al., 1996). The 3 兺 channel Gribakin and Flambaum, 1993, Eq. (13).

Rev. Mod. Phys., Vol. 73, No. 2, April 2001


Anthony J. Leggett: Bose-Einstein condensation in the alkali gases 319

orthogonal to the bound state; see, for example, Leggett, (4.5) follows from Eq. (4.4) if (and only if) in evaluating
1999b) that for qr 0 Ⰶ1 we have 具 E int 典 we work to first order only in a s .

p q⬵ 冉 4 ␲ a s /V
q 2 ⫹a s⫺2 冊 2
. (4.2)
Let us briefly list the conditions necessary for the va-
lidity of Eq. (4.5) in the time-independent case. First, l
⫽0 scattering must be negligible. Second, the ⬃very exis-
If we assume that for N⬅nV atoms the contributions to tence of the limit indicated by the notation → implies
the total depletion from the various pairs are additive, the condition k c a s Ⰶ1, where k c is the characteristic
we find for the total depletion ␦ N⬅ 兺 q⫽0 具 n q 典 the esti- wave-vector scale of the many-body wave function ⌿
mate and is of the order of the inverse of the smallest of the
L c listed in the last subsection (cf., however, below). As
␦ N/N⬃na s3 . (4.3)
we have seen, these conditions are relatively well ful-
This single-pair contribution should be carefully distin- filled for the alkali Bose gases under BEC conditions.
guished from the Bogoliubov-level depletion to be dis- The question of the validity of Eq. (4.5) in a time-
cussed in Sec. VIII, which is a genuinely collective effect dependent situation is a little more subtle, and I return
and, as we shall see, proportional to (na s3 ) 1/2. In fact, it to it in Sec. VI.
is conventional to exclude14 the contribution (4.3) from Equation (4.5) [or Eq. (4.4)] is possibly the single
the depletion. most important result in the whole of the physics of the
dilute ultracold alkali gases (note that it in no way re-
quires Bose statistics). While one can find many deriva-
C. The effective interaction
tions of it in the literature [e.g., Huang (1987), Secs.
I now turn to the effects of the atom-atom scattering 13.1–3 and 5, or Stoof et al. (1996)], few of them are
on the properties of the many-body alkali-gas system. sufficiently general to cover the spatially and, often,
The fundamental result is that under appropriate condi- temporally inhomogeneous situations typical of the
tions, and with appropriate qualifications, the true inter- BEC alkali gases; I therefore now sketch an argument
action potential V at (r) of two atoms of reduced mass m r that, while not particularly rigorous, is at least rather
may be replaced by a delta function of strength general, and indicates the justification for the above con-
2 ␲ ប 2 a s /m r , where a s is the low-energy s-wave scatter- ditions of validity.
ing length. The case of most frequent interest is that of Let us consider a system of N particles and arbitrarily
two similar particles each of mass m (which may or may pick out two of them, say 1 and 2, denoting their relative
not be in the same hyperfine state);15 thus the commonly and center-of-mass coordinates, respectively, by r and R.
quoted form of the effective interaction is The most general (pure) states of the many-body system
will then be described by a wave function ⌿(r:R, 兵 ␰ 其 ,t)
4 ␲ a sប 2 where 兵 ␰ 其 schematically denotes the coordinates of the
U 共 r兲 ⫽ ␦ 共 r兲 . (4.4) remaining N⫺2 particles. In the following, it is the de-
m
pendence on r that is crucial, and the parametric depen-
It is actually rather more physically meaningful to ex- dence of ⌿ on R, ␰, and t will not be written out explic-
press the result (4.4), plus the necessary qualifications, in itly.
the following alternative way: the mean interaction en- The essential question we wish to answer is: How does
ergy of the many-body system is given by the expression the interaction V at (r) between the pair 1 and 2 affect
1 4 ␲ a sប 2 the (mean) total energy of the many-body system? Ac-
兺ij 兩 ⌿ 共 r ij → 0 兲 兩 2 ,

具 E int 典 ⫽ • (4.5) tually, this question is ambiguous. One interpretation
2 m
would be: How do the energies of the exact many-body
where ⌿ is the many-body wave function and the nota- eigenstates change when the interaction V at (r) is

tion r ij →0 means that the separation r ij of the two at- ‘‘switched on?’’ This question is difficult to answer in
oms, while large compared to a s , is small compared to general, and in any case is not of very obvious physical
any other characteristic length (thermal de Broglie relevance, especially in the context of time-dependent
wavelength, interparticle spacing, . . . ). (An equivalent problems. A more physical question would seem to be:
Consider two different states of the system for which
statement is that 兩 ⌿ 兩 2 should be understood as averaged
[for all (R, 兵 ␰ 其 ,t)] the form of ⌿(r) for r greater than or
over a volume Ⰷa s3 .) As we shall see, the very existence
equal to some cutoff r c Ⰷa s ,r o is identical; subject to this
of this limit implies some conditions for the validity of
condition they are constructed to minimize, in the first
Eq. (4.5). With this understanding, it is clear that Eq.
case, the kinetic energy of relative motion, and in the
second, the sum of the latter and the potential energy
14 V at (r). What is the difference in their (mean) total en-
If we attempt an estimate of the kinetic energy on the basis
ergies? Denote the first solution by ␺ 0 (r) and the sec-
of Eq. (4.2), we find it is linearly divergent, so that at this point
we must go back to the true short-range (rⱗr 0 ) behavior of
ond by ␺ (r); in the following it is an essential assump-
␺ (r). We return to this point in Sec. VIII.C. tion that the relevant forms of ␺ 0 (r) can be taken
15
However, to avoid having to consider the effects of indis- constant for r⭓r c (which is crudely speaking equivalent
tinguishability at this stage we continue to assume they can be to the condition k c a s Ⰶ1; see above). We denote the
‘‘tagged.’’ constant value simply by ␺ 0 .

Rev. Mod. Phys., Vol. 73, No. 2, April 2001


320 Anthony J. Leggett: Bose-Einstein condensation in the alkali gases

Define the cutoff radius r c such that a s Ⰶr c Ⰶk ⫺1c


D. Effects of indistinguishability
(where k c is, as above, a characteristic wave vector for
the states described by ⌿) and consider a trial wave In the preceding sections, it has been assumed that the
two atoms whose interactions we are considering are in
function of the form
some way tagged, i.e., are distinguishable. It will turn
␺共 r 兲⫽␺0 , r⭓r c out (see Sec. IV.E) that this is in effect so, even for
atoms that are chemically and isotopically identical, if
共 1⫺a s /r 兲 their initial hyperfine indices are different. However, let
␺共 r 兲⫽ ␺ , r o Ⰶr⭐r c
共 1⫺a s /r c 兲 0 us consider a collision in which the incoming atoms are
identical in all respects including their hyperfine indices,
␺ 共 r 兲 ⫽complicated, rⱗr o . (4.6) and thus indistinguishable (and obey Bose statistics).
This wave function is not, strictly speaking, normalized What difference does this make?
Actually, as regards the results of Secs. IV.A–IV.C,
[if ␺ 0 (r) is] but it may be checked that the effect of
the answer is: none at all. The reason is that we have
correcting for this is to change the energy by a term of
confined the discussion to s-wave states, and for those
order (k 2c r 2c )ប 2 a s ␺ 20 /m, which is smaller than the result the wave function is automatically symmetric with re-
(4.5) by a factor k 2c r 2c Ⰶ1; thus we may consistently spect to interchange of the atomic center-of-mass coor-
neglect this effect. The form (4.6) of ␺ (r) is also not a dinates (r→⫺r). In particular, the validity of Eq. (4.5)
solution of the two-particle zero-energy time- [or Eqs. (4.4), (4.9), or (4.10)] is completely independent
independent Schrödinger equation (⫺ប 2 /m)ⵜ 2 ␺ (r) of the distinguishability or otherwise of the atoms in-
⫽⫺V at (r) ␺ (r), because of the slope discontinuity at r volved.
⫽r c , but it can be made so by introducing a fictitious Nevertheless, there are two important relations that
shell-like potential V ps (r) of the form are affected by the statistics. The first is the relation be-
tween a s and the total elastic-scattering cross section ␴
V ps 共 r 兲 ⫽⫺ 共 ប 2 a s /mr 2c 兲 ␦ 共 r⫺r c 兲 . (4.7) as conventionally defined, that is, the ratio of the num-
If we now evaluate the expectation value of V ps (r) in ber of particles scattered per unit time to the flux inci-
the state (4.6), we find, independently of the value of r c , dent from infinity. For distinguishable particles ␴ can be
calculated in the standard way from the partial-wave ex-
具 V ps 典 ⫽⫺ 共 4 ␲ ប 2 a s /m 兲 兩 ␺ 0 兩 2 . (4.8) pansion of the incoming wave (see, for example, Landau
Since with V ps present state (4.6) is an eigenfunction and Lifshitz, 1959, Sec. 105), and in particular in the
of Schrödinger’s equations with eigenvalue zero, and limit kr 0 Ⰶ1 of relevance to us is given by 4 ␲ a s2 . How-
hence has 具 H ps 典 ⬅ 具 H 典 ⫹ 具 V ps 典 ⫽0, the expectation value ever, for indistinguishable bosons the odd partial waves
具 H 典 ⫽ 具 H r 典 ⫺ 具 V ps 典 of the actual energy of the state (4.6) in the expansion are forbidden, so that for given incident
is ⫺ 具 V ps 典 . Thus, summing over the contributions of all intensity the probability of the particles’ being in a rela-
tive s state is multiplied by 2. Consequently, for indistin-
pairs i,j, we reproduce Eq. (4.5). Note that this result is
guishable bosons one has
independent of the value chosen for r c provided it lies in
the given window, and moreover should be applicable ␴ ⫽8 ␲ a s2 , (4.11)
equally to a time-dependent situation under appropriate
twice the value for distinguishable particles.
conditions (see Sec. IV.F). Incidentally, it is interesting Indistinguishability has a second effect, which will be
that while in the above argument the condition r c Ⰷr o is crucial for arguments about the metastability of super-
essential, the condition r c Ⰷa s is not, and thus it may be flow and related things (Sec. VI): As noted, the validity
possible to reformulate the argument so as to apply it to of Eq. (4.5) is independent of the statistics. However,
the dense conditions obtaining close to a Feshbach reso- the relation between the quantity 兩 ⌿(0) 兩 2 occurring on
nance (nr o3 Ⰶ1 but na s3 Ⰷ1); I shall not attempt to discuss the right-hand side, which is the probability of finding
this question here. two particles at the origin of their relative coordinate,
Assuming that Eq. (4.5) is indeed justified under the and the single-particle wave functions does depend on
conditions we consider, it is sometimes helpful to write it the statistics. As in the last subsection, we focus on a
in second-quantized form: particular pair of atoms and do not write out explicitly
1 4 ␲ ប 2a s
具 E int 典 ⫽ •
2 m
冓冕 冔
␺ † 共 r兲 ␺ † 共 r兲 ␺ 共 r兲 ␺ 共 r兲 dr ,
the dependence on ␰ and t, which enter only parametri-
cally; moreover, we assume for simplicity that (the r 12
→⬁ limit of) the two-particle wave function prima facie
(4.9)
factorizes (the Hartree-Fock approximation), i.e., that
where, however, in evaluating two-particle operators ignoring indistinguishability we would have
like ␺ (r) ␺ (r), we should bear in mind that they are the
⬃ ⌿ 共 r1 ,r2 兲 ⬵ ␸ 共 r1兲 ␹ 共 r2兲 , (4.12)
appropriate → limit as r1 →r2 of ␺ (r1 ) ␺ (r2 ). A third
form, valid to relative order N ⫺1 , which is sometimes where ␸ and ␹ are each normalized over the volume of
helpful to our intuition, is the system.
Suppose first that the states ␸ and ␹ are identical.
1 4 ␲ ប 2a s
具 E int 典 ⫽ •
2 m
冓冕 冔
␳ 2 共 r 兲 dr . (4.10) Then we get from Eq. (4.5) a contribution from this pair
to the total interaction energy of the form

Rev. Mod. Phys., Vol. 73, No. 2, April 2001


Anthony J. Leggett: Bose-Einstein condensation in the alkali gases 321

具 E int 典 12⫽
4 ␲ ប 2a s
m
冕␹ 兩 共 r兲 兩 4 dr. (4.13)
in the collision of interest, the pair of outgoing levels is
the same as the pair of incoming ones (‘‘elastic’’ colli-
sion). This excludes the exothermic collision processes
Now suppose that ␸ (r) and ␹ (r) represent different, that are often responsible for depopulation in magnetic
mutually orthogonal states. Then we must symmetrize traps; it also excludes a rather delicate effect that may
⌿(r1 ,r2) with respect to the interchange of (1) and (2), occur when different pairs of levels are very nearly de-
i.e., replace Eq. (4.13) by the normalized expression generate (see Sec. IX).
⌿ 共 r1 ,r2兲 ⫽2⫺1/2关 ␸ 共 r1兲 ␹ 共 r2兲 ⫹ ␹ 共 r1兲 ␸ 共 r2兲兴 . (4.14) If the two atoms in question are in the same orbital
state, the situation is straightforward; they must then be
The result is that the probability density at the relative in a symmetric state as regards their hyperfine (internal)
origin is multiplied16 by a factor of 2:
degrees of freedom, and this state then just factors out
⬃ of the calculation. The interaction energy is just given by
兩 ⌿ 共 r 12 → 0 兲 兩 2 ⫽2 兩 ␸ 共 r兲 兩 2兩 ␹ 共 r兲 兩 2 共 r⫽r1⫽r2兲 , (4.15)
the expression (4.14) with a s replaced by the generalized
(s)
scattering length a ␣␤ defined below.
and as a result the contribution to the interaction energy
Consider a pair of otherwise identical Bose atoms
is now
characterized by hyperfine labels ␣, ␤ ( ␣ ⫽ ␤ ) and occu-
具 E int 典 12⫽
8 ␲ ប 2a s
m
冕␸ 兩 共 r兲 兩 2兩 ␹ 共 r兲 兩 2 dr. (4.16)
pying mutually orthogonal orbital states ␸, ␹. The sim-
plest description of their collisions is obtained in terms
of the odd and even wave functions,
Crudely speaking, Eq. (4.16), when compared with Eq.
(4.13), says that two bosons in different (but mutually ⌿ ⫹⬅ ␩ ⫹␨ ⫹ , ⌿ ⫺⬅ ␩ ⫺␨ ⫺ , (4.19)
overlapping) states, e.g., two different momentum eigen-
where ␩ ⫾ , ␨ ⫾ are, respectively, normalized spin and or-
states, interact twice as strongly as when they are in the
bital wave functions of the form
same state (or as they would if distinguishable).
It is clear that this result can be generalized to an ␩ ⫾ ⬅2 ⫺1/2共 ␣ 1 ␤ 2 ⫾ ␣ 2 ␤ 1 兲 , ␨ ⫾ ⬅2 ⫺1/2共 ␸ 1 ␹ 2 ⫾ ␹ 1 ␸ 2 兲 ,
arbitrary many-body state. Expanding the field opera- (4.20)
tors ␺ (r), ␺ † (r) in Eq. (4.9) in an orthogonal one-
where ␸ 1 ⬅ ␸ (r1 ), etc. Because the antisymmetric orbital
particle basis ␹ i (r), we rewrite that equation as
function ␨ ⫺ has zero probability density at the relative
4 ␲ ប 2a s 1 origin, it undergoes no scattering and contributes noth-
E int ⫽
m
• • 兺
具 a †a †a a 典
2 ijkl i j k l ing to the interaction. We now define a generalized
(s)
s-wave scattering amplitude a ␣␤ just as in Sec. IV.B, in
⫻ 冕 ␹* i 共 r兲 ␹ *
j 共 r 兲 ␹ k 共 r 兲 ␹ l 共 r 兲 dr, (4.17)
terms of the asymptotic behavior of the s-wave part of
␨ ⫹ ; note that this definition refers only to the phase of
and a standard Hartree-Fock decoupling of the expecta- ␨ ⫹ as r 12→⬁ and is completely independent of its mag-
tion value then gives nitude. We then see by the arguments of Sec. IV.D that
the energy shift associated with the (normalized) state
4 ␲ ប 2a s 1 ␩ ⫹ ␨ ⫹ is 8 ␲ ប 2 a ␣␤(s)
E int ⫽
m
• •
2 兺ij n i n j共 2⫺ ␦ ij 兲 /m (i.e., including the factor of 2).
Consider now what this means for a collision in which, in
the incoming state, the atom with hyperfine index ␣ is in
⫻ 冕␹ 兩 i 共 r兲 兩 2• 兩 ␹ j共 r兲 兩 2 dr. (4.18) orbital state i(⬅ ␸ ) and that with hyperfine index ␤
is in orbital state j(⬅ ␹ ), and label it for brevity 兩 ␣ i , ␤ j 典 ;
The salient consequence of Eq. (4.18) is that a repul- note that this state is physically distinguishable from
sive interaction (a s ⬎0) favors multiple occupation of a 兩 ␣ j , ␤ i 典 . If now we write the (correctly symmetrized and
single one-particle state—an effect that is in addition to normalized) wave function of the state 兩 ␣ i ␤ j 典 in terms of
the statistical one of Sec. III and is sometimes called an the ⌿ ⫹ and ⌿ ⫺ defined in Eq. (4.19), and take into
‘‘attraction in momentum space’’ (Huang, 1987; No- account that ⌿ ⫺ drops out, we find that the effective
zières, 1995). interaction Hamiltonian is a 2⫻2 matrix in the space
spanned by 兩 ␣ i , ␤ j 典 and 兩 ␣ j , ␤ i 典 of the form

冉 冊
E. Effect of the hyperfine degree of freedom
1 1 4 ␲ ប 2 a ␣␤
(s)
Ĥ⫽ ⫻ ␦ 共 r 12兲 . (4.21)
I shall assume here for simplicity that while the hyper- 1 1 m
fine index may be different for the two atoms involved
Following Oktel and Levitov (1999), it is convenient
to call the process 兩 ␣ i , ␤ j 典 → 兩 ␣ i , ␤ j 典 a forward collision
16 and the process 兩 ␣ i , ␤ j 典 → 兩 ␣ j , ␤ i 典 a momentum-exchange
By an obvious extension of the argument, the probability of
collision; we see from Eq. (4.21) that the matrix ele-
finding three particles at the same point is six times larger
when they are all in different states than when they are all in ments for those two processes are identical in sign and
the same one-particle state. As pointed out by Kagan et al. magnitude.
(1985), this leads to a significant reduction of three-body re- A similar analysis in terms of ⌿ ⫹ and ⌿ ⫺ demon-
combination processes in the BEC state, a prediction verified strates that the total cross section ␴ ␣␤ for (incoherent)
experimentally by Burt et al. (1997). mutual scattering of two different hyperfine species is

Rev. Mod. Phys., Vol. 73, No. 2, April 2001


322 Anthony J. Leggett: Bose-Einstein condensation in the alkali gases

4 ␲ (a ␣␤
(s) 2
) (no factor of 2)—a result that could have been the 2s state as a function of laser frequency was inferred
obtained intuitively by the argument that the hyperfine by monitoring the prompt fluorescence produced when
index is as good a ‘‘label’’ of a particular atom as a the 2s state (which is highly metastable) was mixed with
chemical or isotopic one. the rapidly decaying 2p state by application of an elec-
Finally, a note on the relation of the scattering lengths tric field. Although in general the interpretation of the
(s)
a ␣␤ for different pairs of hyperfine indices ␣, ␤: In some data would be complicated by the Doppler effect, the
cases an exact relation can be obtained by using rigorous authors were able to identify the ‘‘Doppler-free’’ part of
invariance principles such as rotational invariance (see the spectrum, which is due to absorption of two photons
Sec. V.D). It is usually not possible to go beyond this. In traveling in opposite directions and hence not subject to
particular, the tempting idea of trying to obtain relations this complication. Thus one can regard the raw data as
between (e.g.) the three different amplitudes involving being the probability P(⌬ ␯ ) of absorption at a given
atoms in F⫽2, m F ⫽ 1 and F⫽1, m F ⫽ ⫺1 states by value of the laser frequency, which differs by ⌬ ␯ from
taking purely electronic singlet and triplet amplitudes the free-space value; ⌬ ␯ is interpreted as the effect of
and weighting them with the probability of finding, in
the interatomic interactions (the cold collision frequency
the initial (asymptotic) states, the two electrons to be in
shift) as follows.
a relative singlet or triplet usually does not work very
The crucial point is that the interatomic potential
well (see Tiesinga et al., 1996); the near equality of these
three amplitudes for 87Rb (see Julienne et al., 1997) is V at (r) is different for two atoms both in the 1s ground
believed to be a coincidence resulting from the near van- state and for a pair in which one atom is in the 2s ex-
ishing of the electronic triplet amplitude for two 87Rb cited state (the probability of both being in the 2s state
atoms. is negligible). Since the rate of increase of the 2s ampli-
tude, which is given by the two-photon Rabi frequency17
⍀, is almost certainly very tiny compared to ប/ma s2 , or
F. Time-dependent situations: the MIT hydrogen even ប/mr 20 , the above considerations justify us in re-
experiments placing V at (r) in both cases by the appropriate s-wave
scattering length a 1s⫺1s or a 1s⫺2s (Oktel and Levitov,
How fast do the physical conditions have to change in 1999). So, at a point r where the atomic density is n(r),
time before the replacement of the true interatomic po- the expression for the interaction correction to the
tential V at (r) by the scattering-length approximation 1s-2s energy difference (i.e., for the quantity 2h⌬ ␯ )
(4.4) becomes invalid or dubious? It is, at least, ex- should be of the form ␩ •4 ␲ ប 2 (a 1s ⫺a 2s )n(r)/m. But is
tremely plausible that the answer should be the slower ␩ equal to 1, 2, or neither? The simplest argument is that
of the rates ប/m r r o2 and ប/m r a s2 (where r 0 was defined in the action of the laser cannot suddenly change the value
Sec. IV.A); this is the maximum time the system needs of 兩 ⌿(r 12⯝0) 兩 2 ; thus for a state with 100% BEC the
to adjust the relative wave function out to distances of quantity ␩ should be 1, and in the normal state (where
the order of the r c defined in Sec. IV.C. Now, it follows to order N ⫺1 all atoms should be in different orbital
from the results to be obtained in Sec. V [see Eq. (5.13)] states) ␩ should be 2. For a state with a fraction f of the
that if the time dependence is due to a change in the atoms in the condensate, a naive extension of this argu-
trap potential, then even if this is itself very fast the ment would give the result ␩ ⫽2⫺f. However, as
typical inverse time taken by the mean-field wave func- pointed out by Oktel and Levitov (1999), a given atom
tion to respond is at most of the order of nU 0 /ប may commute between the condensate and the normal
⬅4 ␲ nបa s /m; under BEC conditions (na s3 Ⰶ1) this is component on a time scale Ⰶ⍀ ⫺1 , and under those con-
small compared to the above rates, and thus it is self- ditions it may be necessary to take careful account of the
consistent to assume that Eq. (4.4) is valid at all times. momentum-exchange terms (Sec. IV.E); they conclude
A different situation may, however, occur when it is that in this case there are two different resonance fre-
V at (r) itself that is changed in time, for example by quencies for any given f. The situation is further compli-
sweeping the system through a Feshbach resonance. In cated by the fact that, as noted by Fried et al. (1999,
that case it is possible in principle for the rate of change p. 3813, col. 2, paragraph 3), an analysis of their data
to be large compared to ប/ma s2 , and in particular in the according to the above considerations, i.e., assuming
immediate vicinity of the resonance this will always be that a pure condensate occurs and results in a value 1 of
true. Under these conditions the use of Eq. (4.4) may be ␩, yields a total condensate that is implausibly large
dubious. from the point of view of the cryogenics; in fact, the
An interesting real-life case in which not only the authors analyze their data under the assumption that ␩
above considerations but those of Secs. IV and V be- ⫽2 everywhere. (The change in the absorption spectrum
come relevant is the experiment of Fried et al. (1998), on the onset of BEC then comes entirely from
who reported the first observation of BEC in atomic the change in the density distribution.) At the time
hydrogen. The authors tuned a laser to approximately of writing, this whole complex of questions remains
half the 1s-2s transition frequency and directed it on controversial.
trapped atomic hydrogen, arranging to have the beam
retroreflected so that it passed through the sample twice
17
in opposite directions; the probability of excitation of See, for example, Mandel and Wolf (1995), p. 752.

Rev. Mod. Phys., Vol. 73, No. 2, April 2001


Anthony J. Leggett: Bose-Einstein condensation in the alkali gases 323

V. THE GROSS-PITAEVSKII APPROXIMATION ប2 2


⫺ ⵜ ␹ 0 共 r兲 ⫹V ext 共 r兲 ␹ 0 共 r兲 ⫹NU 0 兩 ␹ 0 共 r兲 兩 2 ␹ 0 共 r兲
2m
The simplest possible approximation for the wave
function of a many-body system is a (correctly symme- ⫽ ␮ ␹ 0 共 r兲 , (5.4)
trized) product of single-particle states, i.e., the Hartree- ⫺1
Fock ansatz. In the case of a BEC system at T⫽0, this where ␮ enters as N times the Lagrange multiplier
approximation is usually known as the Gross-Pitaevskii associated with the normalization of ␹ 0 . Because of the
or mean-field approximation. It results in a very simple nonlinearity of Eq. (5.4), ␮ is not in general the energy
equation or set of equations, which are very convenient per particle 具 H 典 N /N; rather, by multiplying Eq. (5.4) by
for numerical calculations and, in the case of the BEC ␹*0 (r), integrating over r, and using the fact that 具 H 典 N is
alkali gases, appear to give a rather good quantitative stationary against small variations of ␹ 0 (r), we see that
description of the behavior in a large variety of experi- ␮ is equal to ␦ 具 H 典 N / ␦ N, i.e., the chemical potential
ments. In the following I first treat, using this ansatz, a (whence the notation). Note that the form of the solu-
gas of a single hyperfine species (‘‘spinless’’ gas) in equi- tion ␹ 0 (r) of Eq. (5.4) is itself implicitly a function of N.
librium at zero temperature, and then generalize the In the literature it is conventional to rewrite the non-
treatment to finite temperature (Sec. V.B), to the time- linear Schrödinger equation (5.4) in terms of the order
dependent case (Sec. V.C), and finally to the case of parameter ⌿(r) defined by Eq. (3.12) (with N 0 ⫽N at
more than one hyperfine species (Sec. V.D). Section zero temperature); it then reads
V.E briefly reviews some applications. Unless explicitly ប2 2
stated otherwise, I shall assume the effective interac- ⫺ ⵜ ⌿ 共 r兲 ⫹V ext 共 r兲 ⌿ 共 r兲 ⫹U 0 兩 ⌿ 共 r兲 兩 2 ⌿ 共 r兲
(s)
tion(s) to be all repulsive (all a ␣␤ ⬎0); for the case of a 2m
single species with attractive interaction, see Dalfovo ⫽ ␮ ⌿ 共 r兲 , (5.5)
et al. (1999), Sec. II.C.
with ⌿(r) normalized so that the space integral of its
A. The Gross-Pitaevskii ground state of a spinless system squared modulus is N. Equation (5.5) is the celebrated
Gross-Pitaevskii equation;19 note that in the present
In this subsection I consider a gas of N atoms all of (time-independent) case this is strictly equivalent to the
the same hyperfine species (and will drop the associated one-particle equation (5.4).20
label), in equilibrium in some trapping potential V ext (r) Associated with Eq. (5.4) or Eq. (5.5) is the important
at zero temperature. The Hartree-Fock ansatz for the concept of a healing length. Consider a situation in
ground state is simply18 which in a given region of space the particle density
N
␳ (r)⫽ 兩 ⌿(r) 兩 2 is nearly constant at a value n. Then the
⌿ N 共 r1 ¯rN 兲 ⫽ 兿 ␹ 0 共 ri 兲 , (5.1) (local) healing length ␰ is defined by

冉 冊
i⫽1
⫺1/2
2 mnU 0
where ␹ 0 (r) is some normalized single-particle wave ␰⬅ ⫽ 共 8 ␲ na s 兲 ⫺1/2. (5.6)
function, to be determined. With this ansatz and the ប2
considerations of Sec. III, the expectation value of the To see the physical significance of ␰, consider for a
energy takes the form moment the case (of course unrealistic for the alkali

具 H 典 N ⫽N 冕 再
dr
ប2
2m
兩 “ ␹ 0 共 r兲 兩 2 ⫹V ext 共 r兲 兩 ␹ 0 共 r兲 兩 2 冎 gases) of a gas of N⬅nV atoms confined to a cubic box
of volume V⬅L 3 ; V ext ⬅0 within the box, but the wave
function ␹ 0 (r) [hence ⌿(r)] must of course vanish on
1
⫹ N 共 N⫺1 兲 U 0
2
冕 dr兩 ␹ 0 共 r兲 兩 4 , (5.2)
the walls. For a noninteracting gas, ⌿(r) would be a
product of sine waves with wavelength 2L. However, it
is intuitively clear that, in the presence of repulsive in-
where the effective interaction constant U 0 is given, ac-
cording to Sec. III, by teractions, it is energetically favorable to make ␳ (r)
nearly constant in the bulk of the box, and in the fact
U 0 ⬅4 ␲ ប 2 a s /m. (5.3) explicit solution of the Gross-Pitaevskii equation (5.4)
Since typically in BEC experiments N is at least 105 , in shows this to be the case. As we approach a wall (z
the following I shall neglect the difference between N →0) ⌿(z) falls off to zero as tanh z/(&␰) (in fact, for a
⫺1 and N. Minimization of the right-hand side of Eq. half-infinite space this is the exact form of the solution).
(5.2) subject to the constraint of normalization of ␹ 0 , Thus ␰ is indeed the length over which the perturbing
followed by division by N, then yields the Hartree equa- effect of the wall is ‘‘healed.’’ Note that under normal
tion for condensed bosons, BEC conditions (na s3 Ⰶ1) ␰ is large compared to a s (but
generally small compared to typical trap dimensions; see

18
Strictly speaking this is an oversimplification: even at this
19
level, the many-body wave function must build in the short- Strictly speaking, the T⫽0 time-independent Gross-
range (r ij ⱗa s ) correlations discussed in Sec. IV.B. In the fol- Pitaevskii equation.
20
lowing I implicitly assume that this has been done. Similar remarks apply to Eq. (5.9) below.

Rev. Mod. Phys., Vol. 73, No. 2, April 2001


324 Anthony J. Leggett: Bose-Einstein condensation in the alkali gases

Sec. II.E); we shall see in Sec. VIII that ␰ is also in a We note that while the effective potential felt by the
sense the length over which the gas heals from internal ‘‘normal’’ (or ‘‘thermal’’) (i⫽0) particles is a function
collisions. only of the total density ␳ (r) given by

␳ 共 r兲 ⫽N 0 兩 ␹ 0 共 r兲 兩 2 ⫹ 兺 n i兩 ␹ i共 r兲 兩 2 ⬅n c共 r兲 ⫹n T共 r兲 ,
i⫽0
B. The spinless gas: finite-temperature equilibrium (5.11)
the potential felt by the condensate is sensitive to n c and
The extension of the Gross-Pitaevskii (Hartree-Fock,
n T separately. This creates a small technical problem,
mean-field) ansatz to excited states of the many-body
since it is not guaranteed that the solutions of Eq. (5.9)
system, and thus to finite temperatures, is at first sight
are necessarily orthogonal to that of Eq. (5.10). This
straightforward: We assume that the approximate en-
point is discussed in detail by Huse and Siggia (1982),
ergy eigenstates of the many-body system are character-
who conclude that in realistic circumstances its correc-
ized by specifying a complete orthonormal set of single-
tion is unlikely to change much the conclusions drawn
particle wave functions ␹ i (r) and occupying each of
from Eqs. (5.9) and (5.10); however, a more satisfactory
them with n i bosons, in such a way that 兺 i n i ⫽N. The
resolution is to note that, while the finite-temperature
form of the ␹ i (r) must be obtained self-consistently by
Gross-Pitaevskii theory is indeed not strictly internally
generalizing the argument used in Sec. V.A for the
consistent in this respect, the next (Bogoliubov) level of
ground state: see below. For the case of thermal equilib-
approximation (see Sec. VIII) will remove the inconsis-
rium at temperature T⬅( ␤ k B ) ⫺1 , we take the expecta-
tency while still in many cases leaving Eqs. (5.9) and
tion value of n i for i⫽0 to be given by the usual Bose
(5.10) as quantitatively good approximations.
distribution:
具 n i 共 T 兲 典 ⫽ 兵 exp关 ␤ 共 ⑀ i ⫺ ␮ 兲 ⫺1 兴 其 ⫺1 (5.7)
and substitute this expression in the interaction term in C. The spinless gas: time-dependent Gross-Pitaevskii
the energy. The number N 0 of particles in the conden- theory
sate (if any), that is, in the special single-particle state
␹ 0 (r), is obtained from the prescription In attempting to generalize the considerations of
Secs. I and II to the time-dependent case, we encounter
N 0 共 T 兲 ⫽N⫺ 兺 具 n i共 T 兲 典 .
i⫽0
(5.8) a nontrivial complication: The order parameter
⌿(r,t) is conventionally defined, as in Eq. (3.12), as
In the following I shall neglect terms of relative order 冑N 0 (t)• ␹ 0 (rt), where N 0 (t) is the single macroscopic
n i /N, i⫽0. eigenvalue of the single-particle density matrix ␳ (rr⬘ :t)
Bearing in mind the expression (4.18) for the interac- and ␹ 0 (rt) is the associated normalized single-particle
tion energy, and noting that terms with i⫽j⫽0 are of eigenfunction. Now, we can write down the obvious
order N ⫺1 relative to those with i⫽j⫽0, we see that for time-dependent generalizations of Eqs. (5.4) and (5.5),
any particular energy eigenstate the function ␹ 0 and ␹ j respectively (neglecting for the moment terms in n i , i
(j⫽0) are given, respectively, by the solutions of the ⫽0),
equations ⳵␹ 0 共 rt 兲 ប2 2


iប ⫽⫺ ⵜ ␹ 0 共 rt 兲 ⫹V ext 共 rt 兲 ␹ 0 共 rt 兲
ប2 2 ⳵t 2m
⫺ ⵜ ⫹V ext 共 r兲
2m ⫹N 0 共 t 兲 U 0 兩 ␹ 0 共 rt 兲 兩 2 ␹ 0 共 rt 兲 (5.12)


4␲ប as
m
2


N 0 兩 ␹ 共 r兲 兩 2 ⫹2
i⫽0

n i 兩 ␹ i 共 r兲 兩 2 冊冎 ␹ 0 共 r兲
and
⳵ ⌿ 共 rt 兲 ប2 2
iប ⫽⫺ ⵜ ⌿ 共 rt 兲 ⫹V ext 共 rt 兲 ⌿ 共 rt 兲
⫽ ␮ ␹ 0 共 r兲 , (5.9) ⳵t 2m

再 ⫺
ប2 2
2m
ⵜ ⫹V ext 共 r兲
⫹U 0 兩 ⌿ 共 rt 兲 兩 2 ⌿ 共 rt 兲 .
In the literature, one can find the term ‘‘time-dependent
(5.13)

冉 冊冎
Gross-Pitaevskii equation’’ used to refer either to Eq.
4 ␲ ប 2a s

m
2N 0 兩 ␹ 共 r兲 兩 2 ⫹2
i⫽0

n i 兩 ␹ i 共 r兲 兩 2 ␹ j 共 r兲 (5.12) or to Eq. (5.13). However, it is clear that, unlike
their time-independent analogs, Eqs. (5.12) and (5.13)
are not equivalent unless the condensate number N 0 (t)
⫽ ⑀ j ␹ j 共 r兲 共 j⫽0 兲 , (5.10) is conserved in time; since this condition itself follows
where the chemical potential ␮ is defined as ␦ 具 E 典 / ␦ N at from Eq. (5.13), it is clear that Eq. (5.13) implies Eq.
constant 兵 n i 其 , or what comes to the same thing, constant (5.12) but not vice versa. A very systematic analysis of
entropy. In thermal equilibrium Eqs. (5.9) and (5.10) the time-dependent problem, in terms of an expansion
must be solved self-consistently by replacing the quanti- in a parameter that is essentially the typical value of the
ties n i and N 0 by their thermal expectation values (5.7) quantity (na s3 ) 1/4, has been given by Castin and Dum
and (5.8), respectively. (1998); see also Gardiner (1997). In this subsection I

Rev. Mod. Phys., Vol. 73, No. 2, April 2001


Anthony J. Leggett: Bose-Einstein condensation in the alkali gases 325

adopt a less rigorous but hopefully rather intuitive ap- emerges from their calculation, is that if we define the
proach to the question of the validity of Eqs. (5.12) and condensate wave function ␹ 0 (rt) in the more general
(5.13). way described in Sec. III (i.e., as the wave function of
We work at T⫽0 and start from the many-body the unique ‘‘macroscopically occupied’’ single-particle
Hamiltonian consisting of the usual kinetic and external state), then it still satisfies Eq. (5.12) at the order in ⑀
potential-energy terms plus the interaction term where Eq. (5.13) fails. It should be noted that the quan-
2 U 0 兺 ij ␦ (ri ⫺rj ), and insert it in the many-body Schrö- tity ⑀ may itself depend on time.
1

dinger equation. We then make for the solution of the It is often useful to define a ‘‘reference’’ solution to
latter the Hartree-type ansatz which is the obvious gen- Eq. (5.13)—call it ⌿ 0 (rt)—and then study the behavior
eralization of Eq. (5.1): of small deviations ␦ ⌿(rt) from ⌿ 0 . Because of the
N presence of the nonlinear term, we get a pair of coupled
⌿ N 共 r1 r2 ¯rN :t 兲 ⫽ 兿
i⫽1
␹ o 共 ri :t 兲 . (5.14) equations for ␦ ⌿(rt) and its complex conjugate
␦ ⌿ * (rt):
Note that in making this ansatz we already explicitly ⳵ ប2 2
assume that N 0 (t)⫽N⫽const. Substituting this ansatz iប ␦ ⌿ 共 rt 兲 ⫽⫺ ⵜ ␦ ⌿⫹V ext 共 rt 兲 ␦ ⌿ 共 rt 兲
⳵t 2m
into the time-dependent many-body Schrödinger equa-
tion, we obtain ⫹U 0 兩 ⌿ 0 共 rt 兲 兩 2 ␦ ⌿ 共 rt 兲
N
⳵␹ 0 共 ri ,t 兲 ⫹U 0 关 ⌿ 0 共 rt 兲兴 2 ␦ ⌿ * 共 rt 兲 . (5.16)
iប 兺
i⫽1 ⳵t
•⌳ i 共 兵 rj 其 :t 兲
The equation for ␦ ⌿ * is the complex conjugate of Eq.

冋再
(5.16). A case of special interest is when ⌿ o (rt) repre-
ប2 2 1
⫽ 兺i ⫺
2m
ⵜ i ⫹V ext 共 ri 兲 ⫹ U 0
2
sents the Gross-Pitaevskii ground state: in that case, to
get agreement between Eqs. (5.13) and (5.5) we write

⫻ 兺
N

j⫽i
冎 册
␦ 共 ri ⫺rj兲 • ␹ 0 共 ri :t 兲 ⌳ i 共 兵 rj 其 :t 兲 , (5.15)
⌿ o 共 rt 兲 ⫽⌿ 共 r兲 exp⫺i ␮ t/ប,
where ⌿(r) satisfies the time-independent Gross-
Pitaevskii equation (5.5). It is then convenient to write
(5.17)

where ⌳ i (rj :t) is simply a shorthand for the ␦ ⌿(rt) in the form
(ri -independent) quantity 兿 j⫽i N
␹ 0 (rj :t). In view of the
product nature of the ansatz (5.14), it is consistent to ␦ ⌿ 共 rt 兲 ⫽exp⫺i ␮ t/ប 关 u 共 r兲 exp⫺i ␻ t
take the density of particles other than i at ri , 兺 j⫽i N
␦ (ri ⫹ v * 共 r兲 exp⫹i ␻ t 兴 , (5.18)
⫺rj ), to be given by its expectation value over Eq.
and some straightforward algebra then shows that the
(5.14), namely, the total density ␳ (rt)⬅ 兺 i⫽1 N
兩 ␹ 0 (ri :t) 兩 2
functions u(r), v (r) satisfy the Bogoliubov–de Gennes
(where as usual we ignore the difference between N
equations
⫺1 and N). Then it is clear that Eq. (5.15) is satisfied if
Eq. (5.12) is; moreover, since we have already assumed ប ␻ u 共 r兲 ⫽ 兵 Ĥ o ⫺ ␮ ⫹2U o 兩 ⌿ 共 r 兲 兩 2 其 u 共 r兲 ⫹U o ⌿ 2 共 r兲v共 r兲
that N 0 (t)⫽N⫽const, Eq. (5.13) also follows. The argu- (5.19)
ment can clearly be generalized (with some labor) to the
case of finite temperature (i.e., 具 n i 典 ⫽0 for i⫽0), and the ⫺ប ␻ u 共 r兲 ⫽ 兵 Ĥ o ⫺ ␮ ⫹2U o 兩 ⌿ 共 r 兲 兩 2 其 v共 r兲
result is, as we might intuitively expect, simply to replace
⫹U o ⌿ * 2 共 r兲 u 共 r兲 , (5.20)
␮ in Eq. (5.9) and ⑀ i in Eq. (5.10) by the quantities
iប ⳵␹ 0 (rt)/ ⳵ t and iប ⳵␹ i (rt)/ ⳵ t, respectively.21 where Ĥ o is a shorthand for 关 ⫺(ប /2m)ⵜ ⫹V ext (r) 兴 .
2 2

The above simple argument shows that once we have We shall rederive Eqs. (5.19) and (5.20) from a different
made the ansatz (5.14) for the many-body wave func- point of view in Sec. VIII. In the special case of a weakly
tion, then the unique consistent choice of ␹ 0 (rt) is the interacting gas in free space 关 V ext ⬅0,⌿(r)⫽ 冑n o ⬵ 冑n 兴 it
function determined by Eq. (5.12). However, it does not is clear that u(r) and v (r) have the form of plane waves:
of course assure us that the ansatz (5.14) is consistent, u(r)⫽A exp(ik•r), v (r)⫽B exp(ik•r), and moreover
and indeed in the presence of finite interactions it is from Eq. (5.5) ␮ is simply equal to U o n. Explicit solu-
not—even if we start at t⫽0 with a simple product wave tion of Eqs. (5.19) and (5.20) then yields the dispersion
function of the type (5.14), the last term in Eq. (5.15) relation
will in general mix in more complicated (correlated)
many-body states. Thus we should expect Eq. (5.13) to ប ␻ 共 k 兲 ⫽ 关 ⑀ k 共 ⑀ k ⫹2nU 0 兲兴 1/2 共 ⑀ k ⬅ប 2 k 2 /2m 兲 . (5.21)
be valid only to lowest order in ⑀ ⬅(na s3 ) 1/4, and indeed Introducing the hydrodynamic speed of sound c s by
the rigorous calculation of Castin and Dum (1998)
n ⳵ 2 E nU 0
shows this to be so. What is less expected, but also c s2 ⫽ ⫽ , (5.22)
m ⳵n2 m
we can rewrite Eq. (5.21) in the form
21
The quantity 具 n i 典 is then constant in time and determined
by its t⫽0 value. ␻ 共 k 兲 ⫽ 共 c s2 k 2 ⫹ប 2 k 4 /4m 2 兲 1/2. (5.23)

Rev. Mod. Phys., Vol. 73, No. 2, April 2001


326 Anthony J. Leggett: Bose-Einstein condensation in the alkali gases

This is the famous Bogoliubov spectrum for a dilute order parameter ⌿(r, ␣ ) which is a spinor in the space of
Bose gas; it reduces to ⑀ k /ប for kⰇប/mc s [i.e., k ␰ Ⰷ1, the hyperfine axes. The kinetic-energy operator then has
since ␰ can be written from Eqs. (5.6) and (5.22), as the standard form ⫺(ប 2 /2m)ⵜ 2 , and the local external
冑2ប/mc s ] but in the opposite limit has the sound-wave potential is a matrix V ␣␤ (r) with respect to the hyper-
form c s k. We return to Eq. (5.21) in Sec. VIII, just not- fine axes. It remains to consider the interaction energy.
ing here that regardless of the value of k the fluctuation By an obvious generalization of the considerations of
of the atomic density ␳ (rt)⫽ 兩 ⌿(rt) 兩 2 around its equilib- Sec. IV, it is intuitively plausible that the general expres-
rium value n is given by the standard sound-wave form sion for the atomic interaction energy is
␦ ␳ 共 rt 兲 ⫽const cos关 k•r⫺ ␻ 共 k 兲 t 兴 . (5.24) 1
E int ⫽ 兺 U 具 ␺ † 共 r兲 ␺ ␤† 共 r兲 ␺ ␥ 共 r兲 ␺ ␦ 共 r兲 典 ,
2 ␣␤ , ␥ ␦ ␣␤␥ ␦ ␣
(5.26)
D. Effects of the hyperfine degree of freedom
where ␺ ␣† (r) is the creation operator for an atom at
Introduction of an internal (hyperfine) degree of free- point r and with spin (hyperfine) index ␣, and U ␣␤␥ ␦ is
dom leads immediately to at least three complications in related to the s-wave scattering length matrix a ␣␤␥ ␦ by
the picture developed above. First, the axes in the hy- 4␲ប2
perfine space with respect to which the external poten- U ␣␤␥ ␦ ⫽ a ␣␤␥ ␦ . (5.27)
m
tial is diagonal may vary in (real) space and, possibly, in
time. Second, the s-wave scattering amplitude will in Note that because of the Bose statistics U ␣␤␥ ␦ may be
general have a matrix structure in the hyperfine space. taken to be symmetric under the exchanges ␣ ↔ ␤ and
Third, ‘‘simple’’ BEC is no longer the only natural pos- ␥ ↔ ␦ . In any given physical situation the possible forms
sibility (see Nozières, 1995). of the matrix a ␣␤␥ ␦ may be constrained by symmetry
In dealing with these complications we have a choice considerations. Consider, for example, the case of a
of basis in the hyperfine space: Either we regard the axes system moving in a small magnetic field B(r)
as fixed independently of the spatial coordinates r and ⫽n̂(r) 兩 B(r) 兩 , 兩 B(r) 兩 ⰆB hf , and suppose that all atoms
time t, or we allow them to depend on r and, possibly, t, are known to be in the lower (F⫽1) hyperfine manifold,
for example in such a way that the local external poten- as in the experiments of Stenger et al. (1998a). Under
tial is everywhere diagonal (adiabatic basis). these conditions the local interaction energy can be a
Let us first briefly use the latter choice to discuss the function only of the only two quantities one can form
very simplest case, in which the spinor ⌿(r, ␣ ) describ- that are linear in ␺ ␣† ␺ ␤† ␺ ␥ ␺ ␦ and invariant under spin
ing the condensate is of the form ⌿(r) 兩 ␣ o 典 where 兩 ␣ o 典 is rotation, namely, the squares of the total density ␳ (r)
a fixed spinor in the adiabatic basis (e.g., that corre- and of the spin density S(r). Expressing these quantities
sponding to F⫽1,m F ⫽1); 22 it can be shown (Ho and as linear combinations of ␺ ␮† (r) ␺ ␯ (r) and using Eq.
Shenoy, 1996) that this is a good approximation for the (5.27), we find (see Ho, 1998)
ground state when the axes of the latter basis are suffi-
ciently slowly varying. Then it turns out that if we wish 1
a ␣␤␥ ␦ ⫽ 关 a 1 ␦ ␣␥ ␦ ␤ ␦ ⫹a 2 S␣␥ •S␤ ␦ ⫹ 共 ␣ ↔ ␤ 兲兴 , (5.28)
to define a superfluid velocity vs in the natural way, so 2
that the local particle current is vs 兩 ⌿(r) 兩 2 , the appropri-
where S is the (vector) spin operator for the F⫽1 state.
ate definition is no longer Eq. (3.15) but rather
The quantities a 1 and a 2 can be related to the scattering
ប amplitudes of two atoms in the K⫽2 and K⫽0 states,
vs 共 rt 兲 ⫽ D␸ 共 rt 兲 , ␸ 共 rt 兲 ⬅arg ⌿ 共 rt 兲 , (5.25) where K is the magnitude of the (conserved) total
m
atomic spin K⬅F1 ⫹F2 . By considering, for example, the
where D is the covariant derivative introduced by the K⫽0 state (S␣␥ •S␤ ␦ ⫽⫺2) and the K⫽2,m K ⫽2 state
‘‘bending’’ in space of the hyperfine axes (see Weinberg, (S␣␥ •S␤ ␦ ⫽1) we find
1972, Chap. 4). As a result, vs no longer satisfies the
irrotationality condition (3.16), and even in a simply 1 1
a 1 ⫽ 共 a K⫽0 ⫹2a K⫺2 兲 , a 2 ⫽ 共 a K⫽2 ⫺a K⫽0 兲 .
connected geometry the equilibrium state may sustain a 3 3
finite orbital angular momentum (Ho and Shenoy, 1996; (5.29)
cf. Loss et al., 1990). To obtain a Gross-Pitaevskii formalism in a general
For most purposes, however, and in particular where choice of axes, we assume as above simple BEC, i.e.,
the effects of interactions are important, a choice of hy- make the ansatz
perfine axes that is independent of spatial coordinate
N
(and time) is much more convenient, and I shall use it
from now on. For the moment let us assume simple ⌿ N⫽ 兿
i⫽1
␹ o 共 ri , ␣ i :t 兲 , (5.30)
BEC, so that we can define, as in Sec. III.D, a unique
consider a particular point r (and time t), and use our
freedom to choose the axes arbitrarily to choose them,
22
In words, the condensate atoms are in the lower hyperfine temporarily, so that only a single component (call it ␣ o )
manifold and their spins are everywhere oriented along the of the spinor ␹ o (r, ␣ ), or equivalently of the order pa-
direction of the local magnetic field. rameter ⌿(r, ␣ ), is nonvanishing: ⌿(r, ␣ o )⬅ ␦ ␣␣ o ⌿(r).

Rev. Mod. Phys., Vol. 73, No. 2, April 2001


Anthony J. Leggett: Bose-Einstein condensation in the alkali gases 327

Then the formulas of Sec. IV go through verbatim, and Eq. (5.34) may be regarded as related to a special case of
we find for the local interaction energy density E int (r) the solution ⌿ ␣ (rt) of Eq. (5.35) by
within the Gross-Pitaevskii approximation the simple
⌿ ␣ 共 rt 兲 ⫽exp共 ⫺i ␮ t/ប 兲 ⌿ ␣ 共 r 兲 . (5.36)
expression
It is worth noting explicitly that in Eqs. (5.34) and (5.36)
1 4␲ប2
E int 共 r兲 ⫽ • ā s 兩 ⌿ 共 r兲 兩 4 , (5.31) the chemical potential ␮ is common to all components
2 m ⌿ ␣ (r); see below.
where ā s denotes the component of a ␣␤␥ ␦ corresponding The generalization of even the static Gross-Pitaevskii
to ␣ ⫽ ␤ ⫽ ␥ ⫽ ␦ ⫽ ␣ o . Transforming back to a general theory (let alone the time-dependent one) to nonzero
frame and integrating over r, we obtain for the total temperature in the multicomponent case appears not to
interaction energy the expression23 be entirely trivial, and to the best of my knowledge has
not been much discussed in the literature for the most
E int ⫽
1
兺 U
2 ␣␤␥ ␦ ␣␤␥ ␦
冕 ⌿ ␣* 共 r兲 ⌿ ␤* 共 r兲 ⌿ ␥ 共 r兲 ⌿ ␦ 共 r兲 dr,
general case. The basic difficulty is that while we can
certainly define normal-component eigenfunctions
(5.32) ␹ i␭ (r, ␣ ), we have in general no a priori guarantee that
their spinor structure will be simply related to that of the
4␲ប2 condensate [in particular, the relevant spinors need not
U ␣␤␥ ␦ ⬅ a ␣␤␥ ␦ (5.33)
m be locally parallel or orthogonal to the condensate
spinor ⌿(r, ␣ )]. 24 I shall therefore discuss here only the
[where ⌿ ␣ (r) and hence E int may also be a function of
simplest case, namely, that in which not only does the
time]. The right-hand side of Eq. (5.32) is the general
external potential V ␣␤ (r) have r-independent eigenval-
expression, within the T⫽0 Gross-Pitaevskii approxima-
ues but the condensate has an r-independent spinor
tion, for the interaction energy of a Bose system exhib-
structure, i.e., ⌿(r, ␣ )⫽⌿(r) ␰ o ( ␣ ); it is convenient to
iting simple BEC, and it is routinely used in the litera- choose the basis so that ␰ o ( ␣ )⫽ ␦ ␣␣ o . Then, in this basis,
ture for the analysis of such a system.
the normal eigenfunctions within the Hartree-Fock ap-
We now proceed by straightforward analogy with the
proximation will also be fixed spinors: ␹ i␭ (r, ␣ )
arguments of Secs. V.A–V.C above. Given the ansatz
⫽ ␸ i␭ (r) ␰ ␭ ( ␣ ), where one of the ␰ ␭ ’s is ␰ o ( ␣ ) and the
(5.30), the time-independent T⫽0 Gross-Pitaevskii
rest are orthogonal to the latter. One can then go
equation is obtained by minimizing the total energy and
through the standard Hartree-Fock decoupling of the
takes the form of a set of equations for the j spinor
interaction energies in Sec. IV, with the upshot that the
components ⌿ ␣ (r),
interaction between normal particles in states ␭ and ␭ ⬘
ប2 2 has the factor of 2 only for ␭⫽␭ ⬘ , and the interaction of

2m
ⵜ ⌿ ␣共 r 兲 ⫹ 兺␤ V ␣␤共 r 兲 ⌿ ␤共 r 兲 the condensate with a normal particle of hyperfine index
␭ has it only for ␭⫽0.
I now turn to the question of simple versus multiple
⫹ 兺
␤␥ ␦
U ␣␤␥ ␦ ⌿ ␤* 共 r兲 ⌿ ␥ 共 r兲 ⌿ ␦ 共 r兲 ⫽ ␮ ⌿ ␣ 共 r兲 (general) BEC. A very important consideration here is
that the generic factor of 2, which for a spinless system
共 ␣ ⫽1¯j 兲 . (5.34) with repulsive interactions opposes simultaneous occu-
Similarly, the zero-temperature time-dependent Gross- pation of two different (but spatially overlapping) or-
Pitaevskii equation takes the form bital states, is absent in the case of simultaneous occu-
pation of two mutually orthogonal hyperfine states, and
⳵⌿␣ ប2 2 thus such occupation (fragmentation) is not at all un-
iប
⳵t
共 rt 兲 ⫽
2m
ⵜ ⌿ ␣ 共 rt 兲 ⫹ 兺␤ V ␣␤共 rt 兲 ⌿ ␤共 rt 兲 common. For pedagogical convenience let us consider
the very simplest situation, that of a spatially uniform

system with only two relevant hyperfine states character-
⫹ U ␣␤␥ ␦ ⌿ ␤* 共 rt 兲 ⌿ ␥ 共 rt 兲 ⌿ ␦ 共 rt 兲
␤␥ ␦ ized by r-independent spinors 兩 ␣ 典 , 兩 ␤ 典 . (We choose the
basis to diagonalize the external potential energy.)
共 ␣ ⫽1¯j 兲 . (5.35) Then, normalizing in unit volume and setting the par-
Just as in the case of a simple one-particle Schrö- ticle density equal to n, we see that the ground-state
dinger equation, the time-independent solution ⌿ ␣ (r) of Gross-Pitaevskii order parameter has the form
⌿ ␣ 共 r兲 ⫽n 1/2C ␣ ,
23
Needless to say, if we accept the concept of spon- ⌿ ␤ 共 r兲 ⫽n 1/2C ␤ 共 兩 C ␣ 兩 2 ⫹ 兩 C ␤ 兩 2 ⫽1 兲 , (5.37)
taneously broken U(1) symmetry (see Sec. III.D), the
above derivation may be short-circuited by the re-
placements, in Eq. (5.26), 具 ␺ ␣† (r) ␺ †␤ (r) ␺ ␥ (r) ␺ ␦ (r) 典 24
The origin of this difficulty is essentially the point noted
→ 具 ␺ ␣ (r) 典具 ␺ ␤ (r) 典具 ␺ ␥ (r) 典具 ␺ ␦ (r) 典 , 具 ␺ ␣ (r) 典 →⌿ ␣ (r), etc. I pre-
† †
at the end of Sec. V.B above, namely, that within the Hartree-
fer to avoid this ‘‘automatic’’ derivation both for the reasons Fock (Gross-Pitaevskii) approximation the condensate and
given in Sec. III.D and because it makes the generalization to the normal quasiparticles experience different effective Hamil-
general (nonsimple) BEC less transparent. tonians.

Rev. Mod. Phys., Vol. 73, No. 2, April 2001


328 Anthony J. Leggett: Bose-Einstein condensation in the alkali gases

i.e., the many-body ground state has the form [including ⌿ ␣ 共 r,t 兲 ⫽n 1/2C ␣ exp⫺i ␮ ␣ t/ប,
(part of) the time dependence] (5.46)
⌿ ␤ ⫽n 1/2C ␤ exp⫺i ␮ ␤ t/ប.
⌿ N ⫽⌸ i 兵 共 C ␣ 兩 ␣ i 典 ⫹C ␤ 兩 ␤ i 典 )exp⫺i ␮ t/ប 其 . (5.38)
Substituting this into the time-dependent Gross-
Substituting Eq. (5.37) in the time-independent Gross- Pitaevskii equation (5.35) and using the form of U ␣␤␥ ␦
Pitaevskii equation (5.34), we find that the coefficients given by Eqs. (5.40)–(5.42), we find that Eq. (5.46) is a
C ␬ ( ␬ ⫽ ␣ , ␤ ) satisfy the equation solution of Eq. (5.35) provided the ␮ ␬ satisfy

E ␬C ␬⫹ 兺
␭, ␮ , ␯ ⫽ ␣ , ␤
nU ␬ ␭ ␮ ␯ C ␭* C ␮ C ␯ ⫽ ␮ C ␬ 共 ␬ ⫽ ␣ , ␤ 兲 , ␮ ␬ ⫽E ␬ ⫹ 兺 nU ␬ ␭ 兩 C ␭ 兩 2 , ␬⫽␣,␤, (5.47)
␭⫽ ␣ , ␤
(5.39)
which is equivalent to the statement that (for example)
where E ␬ is the eigenvalue of the single-particle Hamil-
␮ ␣ is the partial derivative of the total energy with re-
tonian (kinetic plus external potential energy) corre-
spect to n ␣ at constant n ␤ . It is not difficult to see that
sponding to hyperfine state 兩 ␬ 典 . In general Eq. (5.39)
these results generalize to the more physical case in
will have more than one solution (see below), and for
which the ␹ o ␣ (r), ␹ o ␤ (r) are nonuniform and in general
the ground state we must choose that corresponding to
different (though see below). In this case we write the
the lowest value of the total energy.
order parameter in the more general form
Let us now consider the case in which the interaction
term commutes with the single-particle part of the ⌿ ␣ 共 rt 兲 ⫽⌿ ␣ 共 r兲 exp⫺i ␮ ␣ t/ប,
Hamiltonian, so that the numbers of particles in states ␣ (5.48)
⌿ ␤ 共 rt 兲 ⫽⌿ ␤ 共 r兲 exp⫺i ␮ ␤ t/ប,
and ␤ are separately constant. Expressed in terms of the
matrix elements U ␬ ␭ ␮ ␯ , this means that the only allowed and the generalized form of Eq. (5.47) [obtained by a
terms correspond to the choices ␬ ⫽ ␮ , ␭⫽ ␯ or ␬ ⫽ ␯ , ␭ substitution of (5.48) into (5.39)] is

冉 冊
⫽ ␮ . Introducing the notation
ប2 2
U ␣␣␣␣ ⬅U ␣␣ 共 ⬅4 ␲ ប 2 a s( ␣␣ ) /m 兲 , (5.40) ⫺
2m
ⵜ ⫹V ␬ 共 r兲 ⌿ ␬ 共 r兲 ⫹
␭⫽ ␣ , ␤

U ␬ ␭ 兩 ⌿ ␭ 共 r兲 兩 2 ⌿ ␬ 共 r兲

U ␤␤␤␤ ⬅U ␤␤ 共 ⬅4 ␲ ប 2 a s( ␤␤ ) /m 兲 , (5.41) ⫽ ␮ ␬ ⌿ ␬ 共 r兲 共 ␬ ⫽ ␣ , ␤ 兲 (5.49)


1 1

U ␣␤␣␤ 共 ⬅U ␣␤␤␣ ,etc.兲 ⬅ U ␣␤ ⬅ 4 ␲ ប 2 a s( ␣␤ ) /m ,
2 2 冊 —a form that is often used in the literature for cases in
which the species numbers n ␣ ,n ␤ are separately con-
(5.42) served.
It is actually possible to generalize the description
we find that Eq. (5.39) reduces to
(5.45) even to cases in which the interaction term does

冉 E ␣⫹ 兺
␭⫽ ␣ , ␤

nU ␣ ␭ 兩 C ␭ 兩 2 C ␣ ⫽ ␮ C ␣ (5.43)
not commute with the single-particle Hamiltonian, since
in that case also Eq. (5.47) [or in the spatially inhomo-
geneous case (5.49)] has more than one solution, and
and a similar equation with ␣ → ␤ for C ␤ . It is straight- these can be superposed to generate a solution analo-
forward to verify that the (lowest) solution of Eq. (5.42) gous to Eq. (5.45). However, this generalization does
minimizes the total energy per particle, not appear to be of much practical interest in the con-
text of current experiments, and in view of the nota-
E 共 C ␣ ,C ␤ 兲 ⫽ 兺
␬⫽␣,␤
E ␬兩 C ␬兩 2 tional complications I do not pursue the question here.
Returning to the simple case corresponding to Eq.
1 (5.45), we reemphasize that this equation describes
⫹ 兺
2 ␬ ,␭⫽ ␣ , ␤
共 nU ␬ ␭ 兲 兩 C ␬ 兩 2 兩 C ␭ 兩 2 , (5.44) simple BEC (in a time-dependent state). However, pro-
vided that we are at no time interested in measuring
with respect to arbitrary variations of C ␣ and C ␤ subject operators that change the hyperfine state, it is clear that
to the normalization constraint 兩 C ␣ 兩 2 ⫹ 兩 C ␤ 兩 2 ⫽1 [i.e., in in the thermodynamic limit (N→⬁) all experimental
effect, with respect to arbitrary variation of 具 n ␣ 典 ⫺ 具 n ␤ 典 ]. properties of the state (5.45) are identical to those of
However—and here is the crunch—if processes that the state
change 具 n ␣ 典 ⫺ 具 n ␤ 典 are really absent as we have as- n␣ N

兿 兿
sumed, then the physically interesting states of the
many-body system, even at T⫽0, are by no means re-
⬘ ⬅S
⌿N 兩 ␣ i典 兩 ␤ j典 , (5.50)
i⫽1 j⫽n ␣ ⫹1
stricted to states of the form (5.38). Let us in fact con-
sider states of the more general form where S denotes the symmetrization operator and n ␣ is
(the nearest integer to) 兩 C ␣ 兩 2 N. It is clear that the
⌿ N ⫽⌸ i 兵 C ␣ 兩 ␣ i 典 exp⫺i ␮ ␣ t/ប⫹C ␤ 兩 ␤ i 典 exp⫺i ␮ ␤ t/ប 其 single-particle density matrix corresponding to the state
(5.45) (5.50) has two eigenvalues, n ␣ and n ␤ , which are each of
where we do not assume that ␮ ␣ ⫽ ␮ ␤ . This corresponds order N and thus that ⌿ ⬘N displays multiple BEC (a frag-
to an order parameter of the form mented condensate). We shall see in Sec. VII.C that ⌿ N ⬘

Rev. Mod. Phys., Vol. 73, No. 2, April 2001


Anthony J. Leggett: Bose-Einstein condensation in the alkali gases 329

can be represented (up to an overall phase) as a super- The chemical potential, the central density ␳ o (0), and
position of states of the generic form (5.45), i.e., of the the condensate radius r o are given in terms of ␭ by the
form formulas
N
1
兿 兵 兩 C ␣兩 • 兩 ␣ i 典 ⫹ 兩 C ␤兩 exp i ␸ 兩 ␤ i 典 其 ,
i⫽1
(5.51) ␮ ⫽ ប ␻ o •␭ 2 ,
2
(5.54)

with random relative phase ␸. Since in the absence of r o ⫽␭a ho , (5.55)


processes that change the hyperfine states no physical ⫺3
quantities can depend on ␸, the properties of the state ␳ 共 0 兲 ⫽␭ ⫺3 Na ho . (5.56)
(5.50) are identical (in the thermodynamic limit) to At finite temperatures the results for the condensate are
those of Eq. (5.45). the same except that N is replaced by N o and the defi-
A more general discussion of BEC in a two-state sys- nition of ␭ is similarly modified. In practice the condi-
tem is given in Sec. VII. tion ␭Ⰷ1 is met until temperatures quite close to
It is clear that the above discussion recovers most of
T c , since it turns out that the free-gas formula N o
the results of that given in Sec. IV.E (which was pre-
⫽N 关 1⫺(T/T c ) 3 兴 is not a bad approximation for the real
cisely for the case of conservation of the hyperfine in-
trapped gases. What of the thermal (uncondensed, nor-
dex). However, even in the T⫽0 case (and a fortiori at
finite temperatures) one might worry about the effects mal) component? Following Dalfovo et al. (1999), let us
of the momentum-exchange term discussed there. At T define a dimensionless parameter ␩ ⬅ ␮ (0)/k B T c
⫽0, in the homogeneous case or more generally where ⬃(N 1/6a s /a ho ) 2/5. Then ␩ lies in the range 0.3–0.4 for
the two condensate wave functions ␹ o ␣ (r), ␹ o ␤ (r) are most existing experiments.25 At T c , the density n of the
identical, this term vanishes, but in general it does not: gas and hence the mean-field energy per atom nU o are
its physical effect is, crudely, to scatter an atom initially of the order of ␩ 3/2⬃0.1 of their values at T⫽0; how-
in state ␹ o ␣ (r) into the component of ␹ o ␤ (r), which is ever, the competition of the mean field is not, as at T
orthogonal to ␹ o ␣ (r) and vice versa. By making the ⫽0, with the zero-point energy (ប ␻ o /2) but with the
Gross-Pitaevskii (Hartree) approximation one of course thermal energy k B T c , which is a factor ⬃100 larger [see
ignores such processes, and I believe that, at least at Eq. (3.7)]. Thus, in the normal phase (T⬎T c ) and in fact
zero temperature, if one is going to take them into ac- even below T c , the mutual interactions of the atoms of
count one should simultaneously take into account the the normal component are negligible. In fact, since at
terms taken at the next (Bogoliubov) level of approxi- T c ␭ DB is of order n ⫺1/3, the dilute-gas condition nⰆ1
mation (see Sec. VIII); indeed, they may be regarded as automatically guarantees that nU 0 ⰆkT for T⭓T c .
simply a special case of the latter; see Sec. VIII.F. At However, below T c the interaction between the normal
finite temperature the effects of momentum-exchange component and the condensate can have a non-
collisions with the normal component may be more sig- negligible effect on the condensate fraction, reducing
nificant (Oktel and Levitov, 1999, see Sec. 4.6). the latter according to the estimate of Dalfovo et al. by
as much as 20% for T/T c ⫽0.6.
The application of the time-dependent Gross-
Pitaevskii equation (5.13) to a single-species gas in a
trap has been studied by many authors, both to calculate
E. Applications
the collective excitation spectrum and to explore the
The applications of both the time-independent and nonlinear behavior; I particularly draw attention to the
the time-dependent Gross-Pitaevskii equation to a elegant scaling solution first constructed by Kagan and
single-species alkali gas in a harmonic trap are excel- co-workers (1996) and by Castin and Dum (1997) for a
lently reviewed by Dalfovo et al. (1999), so I just reca- harmonic trap with arbitrary time variation of the trap-
pitulate here the barest essentials, confining myself for ping frequency ␻ o , which is exact in the Thomas-Fermi
simplicity to the isotropic case. A critical dimensionless limit ␭Ⰷ1 (and exact in two dimensions for all ␭). See
parameter that controls the qualitative behavior is the Dalfovo et al. (1999), Sec. IV.D, and for an exhaustive
quantity discussion of the collective excitations, including effects
associated with the normal component, Griffin (2000).
␭⬅ 共 15Na s /a ho 兲 1/5, (5.52) Turning to the multispecies case, a number of authors
where a ho ⬅(ប/m ␻ o ) is the oscillator length. In most
1/2 (e.g., Ho and Shenoy, 1996; Esry and Greene, 1999)
experiments to date ␭ lies in the range 5–10, and in the have used Eq. (5.49) to study the equilibrium of two
following I shall treat it as large compared to unity. In different hyperfine species in a trap. A rather generic
this limit we can neglect, in the T⫽0 Gross-Pitaevskii result of such calculations is that even if the external
equation, the kinetic-energy term except very close to potential is identical for the two species, phase separa-
the edge of the condensate (the Thomas-Fermi approxi- tion will occur if, as in the case of the F⫽2, m F
mation), and the form of the ground-state density distri- ⫽⫹1(‘‘ 兩 2 典 ’’) and F⫽1, m F ⫽⫺1(‘‘ 兩 1 典 ’’) states of
bution is then an inverted parabola:
␳ o 共 r 兲 ⫽ ␳ o 共 0 兲共 1⫺r 2 /r o2 兲 . (5.53) 25
Note that the maximum value of na s3 is ␩ 6 /(15) 3/5.

Rev. Mod. Phys., Vol. 73, No. 2, April 2001


330 Anthony J. Leggett: Bose-Einstein condensation in the alkali gases

87 measurements of Kapitza and of Allen and Misener,


Rb, the scattering lengths are all positive and satisfy
the inequality a 122
⬎a 11a 22 (or, what is approximately the what was observed was the ability of the liquid to flow
same condition for close values of the a ij , a 12⬎(1/2) without apparent friction through channels so narrow
⫻(a 11⫹a 22). In such a case, in a typical harmonic trap, that any ordinary liquid (including the He-I phase of
one species (in this case 兩 1 典 ) will form a ring or shell 4
He above the ␭ temperature) would be almost totally
around the other, while the total density behaves (for clamped by viscous effects.
close a ij ) almost identically to that of a single species. As so often occurs in physics, the original experiment
Such behavior has been verified in the experiments of is not the most conceptually clear-cut demonstration of
Hall, Matthews, Ensher, et al. (1998); see their Fig. 1. the phenomenon. In fact, from a modern point of view,
An even richer situation was realized in the experi- superfluidity is not a single phenomenon but a complex
ments of Stenger et al. (1998a), in which all three sub- of phenomena, and the picture becomes clearer if one
states of the F⫽1 ground multiplet were simultaneously considers not a channel between two bulk reservoirs, as
confined in an optical trap and their distribution studied in the experiments of Kapitza and Allen and Misener,
as a function of parameters such as the (spatially vary- but rather an annular geometry.26 Let us then consider a
ing) magnetic field. In subsequent experiments the same hollow cylinder of height h, inner radius R⫺d/2 and
group investigated metastable configurations of systems outer radius R⫹d/2, where dⰆR; for the moment we
containing two of the three species (Miesner et al., 1999)
assume that deviations from exact cylindrical symmetry
and their relaxation by quantum tunneling of the con-
are small (but nonzero). Then, if the cylinder is filled
densate (Stamper-Kurn et al., 1999). It appears that a
with 4He, we can observe two conceptually distinct
reasonably good quantitative account of all these experi-
ments is given by calculations based on the time- (though related) phenomena. The first is sometimes
independent Gross-Pitaevskii equation (5.49). An inter- called the Hess-Fairbank effect: the system appears to
esting difference from the case of 87Rb (with F⫽2 and come out of equilibrium with its rotating container. To
F⫽1 components) is that the ‘‘accidental’’ near degen- amplify this definition, we all know that if we take our
eracy of the various scattering lengths observed in the annulus filled with water and set it on an old-fashioned
latter case appears to be absent: the difference in the gramophone turntable which we then set into rotation,
K⫽2 and K⫽0 scattering lengths in the F⫽1 ground the water will come (after a delay of maybe ⬇1 min)
multiplet (see Sec. V.D) is of order 10% relative to their into rotation with the annulus and will thereafter rotate
mean (see Burke et al., 1998), with a 2 ⬎a 0 . As a result, with it as long as the turntable continues to rotate. When
phase separation apparently occurs between the 0 and we stop the rotation, the water also gradually comes to
⫾1 components, while the latter two are mutually mis- rest. Imagine now that we do the same experiment with
cible (Stenger et al., 1998a). He, starting above the ␭ point and rotating very slowly
The time-dependent equation (5.35) has similarly (for a 1-cm-radius annulus, the angular velocity would
been applied by a number of groups (e.g., Ho, 1998; have to be ⱗ10⫺4 rad/sec to see the specific behavior to
Ohmi et al., 1998; Ohberg and Stenholm, 1999; cf. Col- be described; in practice smaller radii are used and the
son and Fetter, 1978) both to calculate collective excita- criterion is not quite so stringent). The He behaves in
tions and to study the process of achievement of equi- exactly the same way as the water, coming into rotation
librium when the relative concentration of the species is with the container. Now suppose that, while still rotating
suddenly changed. In the latter context I particularly call with this low angular velocity, we cool the system
attention to the analysis by Sinatra et al. (1999) of the through the ␭ temperature. The He then appears to
experiment of Hall et al. [Hall, Matthews, Ensher, et al. gradually come out of equilibrium with the container,
(1998) and Hall, Matthews, Wieman, and Cornell i.e., to cease (prima facie) to rotate even though the con-
(1998)], on which see also Sec. VII.D. A salient conclu- tainer is still rotating! In fact, as T falls to zero, the He
sion is that because of the very small difference in the appears at first sight to come to rest in the laboratory
a ij ’s for 87Rb, equilibration will take a time (⬃60 msec) frame (or, to be more precise, rather in the frame of the
long compared to that associated with the average fixed stars). It is clear that this behavior cannot simply
mean-field energy (⬃2 msec), as indeed observed. An- reflect very long relaxation times, since the liquid has
other interesting conclusion is that the approach to equi- come out of equilibrium with the container: the‘‘nonro-
librium is nonmonotonic (as observed) and that the phe- tating’’ state must be the true thermodynamic state. This
nomenologically observed damping is a result of Hess-Fairbank effect is the exact analog of the Meissner
strongly chaotic behavior (see Sec. IX.E). effect in a superconductor. It is conventional to define
the superfluid density ␳ s (T) [or superfluid fraction
VI. ROTATIONAL PROPERTIES: SUPERFLUIDITY ␳ s (T)/ ␳ , where ␳ is the total density] in terms of the
experimentally observed value of the temperature-
A. Phenomenology of superfluidity in liquid 4He

Fritz London’s original suggestion in 1938 that liquid 26


All the ensuing considerations have close analogs in the
4
He in its He-II phase below the ␭ temperature exhibits theory of superconductivity, which indeed is nothing but su-
BEC was prompted primarily by the observation in that perfluidity occurring in a charged system; see, for example,
phase of the property of superfluidity. In the original Vinen (1969) or Leggett (1995b).

Rev. Mod. Phys., Vol. 73, No. 2, April 2001


Anthony J. Leggett: Bose-Einstein condensation in the alkali gases 331

dependent moment of inertia I(T)⬅L/ ␻ relative to its Consider a system that, when the ‘‘container’’ is sta-
classical value I cl ⬅NmR 2 : ␳ s (T)/ ␳ ⬅1⫺I(T)/I cl . tionary in the laboratory frame,27 is described by the
The second phenomenon is the following: Again put generic Hamiltonian
the He, above T ␭ , in the annulus and set the latter into 1
rotation, but this time much faster. This time, as we cool Ĥ lab ⫽
i

兵 共 p 2i /2m 兲 ⫹V ext 共 ri 兲 其 ⫹
2 ij

U 共 兩 ri ⫺rj 兩 兲 .
through T ␭ , we see very little change: to all intents and (6.1)
purposes the liquid continues to rotate with the con- We wish to know how to apply statistical mechanics
tainer. Now stop the container. The He continues to ro- when the container is rotated at angular velocity ␻
tate, apparently indefinitely. One can show rigorously around an axis which by convention we shall choose as
that for the container stationary the rotating state can- passing through the origin and along direction ẑ:
not be the thermodynamic equilibrium one, so what we V ext 共 ri 兲 →V ext 共 ri :t 兲 ⬅V 关 r⬘i 共 t 兲兴 , (6.2)
are seeing here is an example of an extremely long-lived
metastable state. I shall refer to this phenomenon as where r⬘i (t)⬅(x i cos ␻t⫹yi sin ␻t,⫺xi sin ␻t⫹yi cos ␻t,zi).
metastability of superflow. At first sight neither the Hess- In the following it is not required that V ext (r) have
Fairbank effect nor the metastability of superflow is ob- exact28 or even approximate symmetry under rotation
viously equivalent to the phenomenon originally ob- around the axis ␻, but it is essential that the interparticle
interaction have this property [for which the centrality
served in 1938; I discuss the relation to this and yet other
of U specified in Eq. (6.1) is sufficient though not strictly
possible definitions of superfluidity elsewhere (Leggett, speaking necessary]; see below.
2000a). The standard result is now the following: Consider a
A simple phenomenological understanding of both time ␶ ⫽2 ␲ n/ ␻ (n integral) such that the instantaneous
the Hess-Fairbank effect and the stability of supercur- configuration of the potential is its original stationary
rents may be obtained if we assume (a) that in analogy one. Then the thermodynamic equilibrium state of the
to the electrons in an atom, the atoms of the condensate system is determined by minimizing the expectation
(which, we recall, must all behave as one) can have only value of the quantity Ĥ eff ⫺TS, where
integral values l ប of their angular momentum, corre-
sponding in the annular geometry to an angular velocity Ĥ eff ⬅Ĥ lab ⫺ ␻ •L̂, (6.3)
of rotation l ប/mR 2 ⬅ l ␻ c , and (b) that (in distinction where L̂ is the operator of total angular momentum. At
to the electrons in an atom) the passage of an atom (or other times the density matrix of the system is modified
rather of the condensate atoms as a whole) from one in such a way as to make it, and thus all physical prop-
value of l to another is impeded by a high free-energy erties of the system, time independent when viewed
barrier. Then it is intuitively plausible that on cooling from the rotating frame. (Note that this does not imply
through T ␭ with ␻ Ⰶ ␻ c the condensate will prefer to that the system is necessarily stationary in that frame.)
come to rest. On the other hand, if the angular velocity The derivation of Eq. (6.3) will fail if the interatomic
potential has a part that is not invariant under rotation
of the container is Ⰷ ␻ c , say n ␻ c where n is in general
around the axis of ␻, as is in general the case when
not integral, then on cooling through T c the condensate dipolar interactions are taken into account (assuming
will simply ‘‘choose’’ the value of l that most closely the magnetic field is not rotated, as in the experiment of
matches its angular velocity to that of the container; in Madison et al., 2000). In such a case there is strictly
particular, if the latter is Ⰷ ␻ c /2, the condensate will sim- speaking no frame of reference in which the Hamil-
ply choose the integer l closest to n, and the difference tonian is time independent, and thus no thermodynamic
between l and n will be barely observable, so that the equilibrium state: in fact, the situation is similar to that
liquid appears to continue to rotate with the container. realized in Couette flow, and the best we can do is to
However, when the rotation stops the free-energy barri- find a steady state. Fortunately, in realistic geometries
ers will prevent relaxation to l ⫽0. such as that of Madison et al. (2000) such effects appear
likely to be a small perturbation.
In the remainder of this section I shall be discussing
the spinless case unless explicitly otherwise stated; for
(other) possible effects of the hyperfine degree of free-
B. Rotating frames of reference
dom, see the last paragraph of Sec. VI.D.
In order to formulate a quantitative account of super- C. Equilibrium of a BEC system in a rotating container
fluidity, whether in liquid 4He or in the BEC alkali
gases, we need to know how to do statistical mechanics Let us first apply formula (6.3) to a free Bose gas
where the potential that confines the system (be it a above the transition temperature in a nearly cylindri-
physical container or a set of magnetic and/or laser
fields) is rotating. This problem is discussed in textbooks
of statistical mechanics (e.g., Landau and Lifshitz, 1969, 27
Strictly speaking, that of the fixed stars; I shall ignore the
Sec. 26), and I consider it in the specific alkali-gas con- difference.
text elsewhere (Leggett, 2000a); here I merely quote the 28
In fact, the limit of exact cylindrical symmetry needs to be
standard result. approached with some care; see Leggett (2000a).

Rev. Mod. Phys., Vol. 73, No. 2, April 2001


332 Anthony J. Leggett: Bose-Einstein condensation in the alkali gases

cally symmetric annular container. With the above re- BEC leads to the Hess-Fairbank effect. Moreover, this
placement, the Bose-Einstein distribution is simply feature persists in more general geometries and in the
presence of interactions. To see this, let us specialize for
n n ⫽ 关 exp ␤ 共 ⑀ n ⫺ប ␻ l n ⫺ ␮ 兲 ⫺1 兴 ⫺1 (6.4) simplicity to the case T⫽0 and assume that the time-
and the angular momentum is independent Gross-Pitaevskii theory will give at least a
qualitative guide to the behavior. Thus we write the
L⫽ 兺n n n ប l n ⫽ 兺n l n 关 exp ␤ 共 ⑀ n ⫺ប ␻ l n ⫺ ␮ 兲 ⫺1 兴 ⫺1 . many-body wave function in the Gross-Pitaevskii (Har-
tree) form and minimize the effective Hamiltonian (6.3).
(6.5) Since the angular momentum operator of the ith particle
can be written as simply ⫺iប ⳵ / ⳵ ␪ i where ␪ i is the polar
Now for a cylindrically symmetric geometry in the limit
angle relative to the rotation axis, the resulting Gross-
dⰆR n can be taken as some pair of quantum numbers
Pitaevskii equation has the form
j,k describing the transverse motion (their detailed na-
ture is irrelevant) plus the angular momentum quantum
number l : associated with the latter is an ‘‘angular’’ ki-
netic energy l 2 ប 2 / 2 mR 2 . Associating with (j,k) the en-
冉 ⫺
ប2 2
2m
ⵜ ⫹iប ␻

⳵␪ 冊
⌿ 共 r兲 ⫹V ext 共 r兲

ergy ˜⑀ jk , we thus have 4␲ប2


⫹ a 兩 ⌿ 共 r兲 兩 2 ⌿ 共 r兲 ⫽ ␮ ⌿ 共 r兲 , (6.9)

再 冋 冉 冊册 冎
l 2ប 2 ⫺1 M s
L⫽ 兺
jk
兺l ប l exp ␤ ˜⑀ jk ⫹
2mR 2
⫺ប ␻ l ⫺ ␮ ⫺1 . and this must be solved subject to the ‘‘single-
(6.6) valuedness’’ boundary condition
The crucial point is that for any realistic geometry and ⌿ 共 ␪ ⫹2 ␲ 兲 ⫽⌿ 共 ␪ 兲 . (6.10)
temperature the quantity ␤ ប 2 /mR 2 ⬅ប 2 /(mR 2 k B T) will The expectation value of the angular momentum is
be extremely small (for example, even for Na in a 100␮ given (at T⫽0 within the Gross-Pitaevskii approxima-
annulus at 1 ␮K it is still only ⬇10⫺6 ), so the sum over l tion) by the expression

冉 冊
can be replaced by an integral and the origin shifted:
l ⬘ ⫽ l ⫺mR 2 ␻ ប. It is then clear from symmetry that the ⳵
具 L̂典 ⬅ẑ 兺 ⌿ N ,⫺iប ⌿
result is simply i ⳵␪i N
L⫽NmR 2 ␻ ⬅I cl ␻ ,
where I cl ⬅NmR is the classical moment of inertia.29
2
(6.7)
⫽ẑ 冕 ⌿ * 共 r兲 ⫺iប冉 ⳵
⳵␪冊⌿ 共 r兲 dr. (6.11)
Thus the gas rotates exactly in pace with the container,
In general there will be more than one solution of Eq.
and ␳ s (T)⫽0 from its definition.
(6.9) that satisfies the boundary condition (6.10), and we
What happens below T ␭ ? Consider for simplicity the
must then choose the one that minimizes the expecta-
case T⫽0. Now all the atoms must be condensed into
tion value of 具 H eff 典 . The above formulation applies in-
the state with lowest ‘‘effective’’ energy, i.e., into the
dependently of the container geometry30 and the sign
single-particle state that minimizes the quantity
(and a fortiori the strength) of the interaction term.
l 2ប 2 For a first qualitative discussion it is convenient to
E ang 共 l 兲 ⫽ ⫺ប ␻ l , l integral. (6.8) separate the cases of toroidal and simply connected ge-
2mR 2
ometries. In the toroidal case let us provisionally assume
It is clear that the value of l that minimizes E ang ( l ) that ⌿(r) is finite and the derivatives of its phase ␸ (r)
is the nearest integer to ␻ / ␻ c , where ␻ c ⬅ប/mR 2 . Thus exist everywhere in the torus; then we can write Eqs.
for 0⬍ ␻ ⬍ ␻ c /2 the condensate stays in the original (6.10) and (6.11), respectively, in the form31
ground state ( l ⫽0), and thus the system has zero an-
gular momentum; for ␻ c /2⬍ ␻ ⬍3 ␻ c /2 the condensate
occupies the state l ⫽1 and the total angular momen- 冖 ⳵␸⳵ ␪ 共 r兲
d ␪ ⫽2n ␲ , n⫽0,⫾1,⫾2 . . . , (6.12)


tum is Nប, and so on. At finite temperatures the angular
momentum is the sum of that of the condensate, which ⳵␸
具 L典 ⫽ẑប 兩 ⌿ 共 r兲 兩 2 dr, (6.13)
behaves similarly to that described but with N ⳵␪
→N o (T), and that of the normal component, which be- where in Eq. (6.12) the integral is taken along any path
haves exactly as in the normal phase, i.e., rotates with that goes around the torus once, staying always within it.
the container and contributes an angular momentum
关 N n (T)/N 兴 I cl ␻ . 关 N n (T)⬅N⫺N o (T). 兴 Thus in this
simple model, ␳ s (T)/ ␳ ⫽N o (T)/N. 30
However, in the following I shall implicitly assume that the
We see, then, that in a cylindrically symmetric con- axis of rotation lies within the container cross section, at least
tainer, even for a noninteracting Bose gas, the onset of at some values of z (otherwise the problem becomes rather
trivial and uninteresting).
31
In the He literature it is conventional to express the ensuing
29
There are slight complications, which I shall ignore, con- argument in terms of the superfluid velocity vs (r) defined by
cerning the small shift in ␮ and the fact that (for d/R finite) a Eq. (3.15); the equation corresponding to Eq. (6.12) is the
meniscus tends to form. Onsager-Feynman quantization condition (3.17).

Rev. Mod. Phys., Vol. 73, No. 2, April 2001


Anthony J. Leggett: Bose-Einstein condensation in the alkali gases 333

n is conventionally called the winding number. From question of the metastability or not of certain excited
Eqs. (6.12) and 6.13), plus the normalization condition states is taken up in the next subsection.
on ⌿(r) (see Sec. V.A), we see that unless the geometry
of the trap [and hence 兩 ⌿(r) 兩 2 ] departs very far from
cylindrical symmetry, the value of 具 L典 in a state charac- D. Metastability of superflow
terized by a given value of n(⫽0) in Eq. (6.12) has the
order of magnitude 1. General considerations
具 L典 ⬃Nnបẑ, (6.14) Let us turn at this point to the types of generic behav-
although this relation becomes an equality only for exact ior possible for a system described by an effective
cylindrical symmetry (in which case it gives ␳ s ⫽ ␳ at T Hamiltonian of the form (6.3). We shall take the angular
⫽0, just as in the noninteracting case). A similar argu- velocity ␻ along the z axis and denote by L the expec-
ment shows that the ‘‘extra’’ kinetic energy in a state tation value of the z component of angular momentum.
with n⫽0 [associated with the angular term in the La- We first consider the effective (␻-dependent) moment of
placian in Eq. (6.9)] is of order n 2 ប 2 /2mR 2 . From these inertia I( ␻ ) defined by
conditions it follows that for any reasonable toroidal ge- I 共 ␻ 兲 ⬅ ⳵ L/ ⳵␻ . (6.17)
ometry there will be a finite range of ␻ around ␻ ⫽0 for
By varying ␻ around some initial ␻ o ( ␻ ⬅ ␻ o ⫹ ␦ ␻ ) and
which the effective Hamiltonian Ĥ eff , Eq. (6.3), is mini-
treating the term ⫺ ␦ ␻•L̂ in Eq. (6.3) as a perturbation,
mized by the choice n⫽0. For finite ␻ the angular mo-
we see that I is in fact the static response function cor-
mentum of this state is not zero (except for exact cylin-
responding to ␦␻. For a mechanically stable state all
drical symmetry), but for reasonable geometries it is
static response functions must be ⬎0, and thus we con-
substantially less than the classical value NmR 2 ␻ . Thus,
clude that I( ␻ ) is positive for any stable state of the
very generically, we expect the system to show a Hess-
system.
Fairbank effect.
Let us now consider the lab-frame free energy F lab
In the simply connected case it is clear that if ⌿(r) is
⬅ 具 H lab 典 ⫺TS, whose minimization determines the pos-
everywhere finite, the winding number n must be zero
sible (meta)stable states in the absence of rotation.
for any closed circuit that stays within the volume occu-
From the condition that L( ␻ ) is determined from the
pied by the system. The simplest configuration that al-
minimum of F eff ⬅F lab ⫺ ␻•L, we find ⳵ 2 F lab / ⳵ L 2
lows nonzero n is a vortex; if for illustration we consider
⫽ ⳵␻ / ⳵ L⫽I ⫺1 ( ␻ ). Now it is clear that while F lab (L)
a single-quantized (n⫽1) vortex centered on the axis of
always has a stable minimum33 at L⫽0, the existence of
rotation, then in terms of the standard cylindrical polar
one or more metastable minima at finite values of L
coordinates r, z, ␪ the order parameter has the generic
requires ⳵ 2 F lab / ⳵ L 2 to be negative for some range of L.
form
If L is a continuous function of ␻, this cannot happen
⌿ 共 r兲 ⬅⌿ 共 ␳ ,z, ␪ 兲 ⫽f 共 ␳ ,z, ␪ 兲 exp关 i ␸ 共 ␪ 兲兴 , (6.15) (see above). Consequently, we conclude that a sufficient
where the real functions f and ␸ satisfy the conditions condition for the absence of metastability is that the
function L( ␻ ) calculated by taking the (absolute) mini-
f 共 0,z, ␪ 兲 ⫽0, (6.16a) mum of F lab ⫺ ␻•L be continuous.


It is somewhat more difficult to prove that the condi-
共 ⳵␸ / ⳵ ␪ 兲 d ␪ ⫽2 ␲ (6.16b) tion is necessary as well as sufficient. A zeroth-order
argument is that if the function L( ␻ ) is discontinuous,
if the circuit in Eq. (6.16b) encircles the line ␳ ⫽0 once this must mean that L ‘‘jumps’’ as a function of ␻, i.e., a
in the clockwise direction. [In the case of approximate macroscopically different state has become the ground
cylindrical symmetry, ␸ ( ␪ )⬇ ␪ .] More complicated ar- state. Barring pathologies (of which one case is the non-
rays of vortices, not necessarily centered on the axis of interacting Bose gas in exactly cylindrical geometry; see
rotation, are also possible.32 All the generic consider- below), it seems plausible that such a macroscopically
ations given above for the toroidal geometry apply with different state will be metastable relative to the true
appropriate modifications also to the simply connected ground state for at least some finite range of ␻ (though
case, and in particular one always expects a Hess- not necessarily for ␻⫽0.) This consideration is related
Fairbank effect. to the fact that the winding number n is a topological
It should be emphasized that all the considerations invariant; that is, it cannot be changed unless the mag-
addressed in this subsection relate to the thermodynami- nitude of the order parameter ⌿(r) is suppressed to
cally stable state of the system, in general in the pres- zero over an appropriate cross section of the terms,
ence of rotation of the container. The quite different something that for positive g is resisted by the interac-
tion term (see the next subsection).

32
However, the case 兩 n 兩 ⬎1 is not of much practical interest,
33
since such multiply quantized vortices are rather generically This follows, for a Bose system in arbitrary geometry, from
unstable against dissociation into several singly quantized the fact that the many-body ground-state wave function can
ones. always be chosen real.

Rev. Mod. Phys., Vol. 73, No. 2, April 2001


334 Anthony J. Leggett: Bose-Einstein condensation in the alkali gases

2. A toy model

These considerations may be illustrated by means of a


toy model, which, while it would be quite difficult to
realize experimentally for the BEC alkalis, is neverthe-
less quite a useful guide to the behavior in more realistic
cases. In particular, it illustrates rather straightforwardly
the fact that hysteresis appears only for a sufficiently
strong repulsion. Consider a narrow annulus with thick-
ness dⰆ radius R, and with weak deviations from exact
symmetry, and imagine it filled with a weakly interacting
Bose gas at T⫽0 and rotated with angular velocity ␻.
More specifically, let ␺ o and ␺ 1 denote, respectively, the
lowest-lying s and p states (that is, the exact eigenstates
of L̂ z ) of a single particle in the torus (with the depen-
dence on the azimuthal and polar coordinates self-
FIG. 2. The shape of the energy curve E( ␹ ) as given by Eq.
consistently determined). The overall phase of ␺ 1 is cho-
(6.24): (a) the nonhysteretic case g⬍V 0 ; (b) the hysteretic case
sen relative to ␺ 0 so as to make V o [Eq. (6.19) below]
g⬎V 0 (schematic). The zeros of energy are arbitrary. The
real and positive. Then define three characteristic ener- heavy dot indicates the position (value of ␹) of the system in
gies, namely, an experiment in which ␦ ␻ is swept from positive to negative
(a) the single-particle quantization energy values; note that, according to Eq. (6.25), the value of ␹ / ␲ is a
rough qualitative measure of the angular momentum per par-
ប 2 /mR 2 ⬅ប ␻ c , (6.18)
ticle.
(b) the asymmetry energy

V o ⬅⫺ 冕 ␺* 1 共 r兲 V ext 共 r兲 ␺ o 共 r兲 dr⬎0, (6.19) 冉


⌿ N ⫽ a o⫹ cos

2
exp共 i⌬ ␸ /2兲

where V ext (r) is the external potential, and


(c) the mean interaction energy per particle, in (say)
the s-wave state, which is given by
⫹a ⫹
1 sin

2
exp共 ⫺i⌬ ␸ /2兲 冊 N
兩 vac典 . (6.23)

It is clear that the only term dependent on ⌬␸ is that in


g⬅
4 ␲ ប 2 Na s
m
冕 dr兩 ␺ o 共 r兲 兩 .
4
(6.20) V o , and thus the optimum choice is to take ⌬ ␸ ⫽0 [so
that, by Eq. (6.19), ⌿(r) is smallest, roughly speaking,
The conditions of weak asymmetry and interaction where V ext is the most repulsive]. The dependence of
the energy and angular momentum per particle on ␹ is
are then explicitly
then given up to terms of relative order N ⫺1 by
兩 V o 兩 Ⰶប ␻ c , 兩 g 兩 Ⰶប ␻ c , 0⬍ ␻ ⬍ ␻ c . (6.21)
34
Note that g can have either sign. Since under these g
E 共 ␹ 兲 /N⫽ ␦ ␻ cos ␹ ⫺V o sin ␹ ⫹ sin2 ␹ , (6.24)
conditions the promotion of particles to any states other 2
than the s and p states is negligible, and since for weak
asymmetry 兩 ␺ 1 (r) 兩 2 ⬵ 兩 ␺ 0 (r) 兩 2 almost everywhere, the ef- 1
fective Hamiltonian takes the form, apart from a con- L/Nប⫽ 共 1⫺cos ␹ 兲 . (6.25)
2
stant,
It is clear that the qualitative behavior of the curve
Ĥ eff ⫽⫺ប ␦ ␻ 共 a ⫹ ⫹ ⫹
1 a 1 ⫺a o a o 兲 ⫺V o 共 a o a 1 ⫹H.c. 兲 E( ␹ ), and thus the presence or absence of metastability
(hysteresis), is determined by the ratio g/V o (see Fig. 2).
⫹g 共 a o⫹ a o a ⫹
1 a1兲, (6.22)
If this quantity is smaller than 1 (which of course is au-
where ␦ ␻ ⬅( ␻ ⫺ 21 ␻ c ) (so 兩 ␦ ␻ 兩 ⬍ 21 ␻ c ) and the last term tomatically the case for g⬍0) then E( ␹ ) is monotonic
is the difference of the interaction term [i.e., Eq. (6.20) for any value of ␻ and the stable solution for ␦ ␻ ⫽0 is
with ␺ o (r) replaced by ␺ (r)] from its value in the pure s ␹ ⫽ ␲ /2 (superposition of s and p states with equal
or p state. [Note carefully that this term enters only be- weight). In this case, if we vary ␦ ␻ as a function of time
cause of the factor of 2 (Sec. IV.D).] and assume that the system can exchange energy and
We look for a solution of the Gross-Pitaevskii type: angular momentum with its environment, it will follow
the (unique) L( ␻ ) curve smoothly. There is no difficulty
with the fact that in going from the s state to the p state
34
In general, negative values of g are apt to lead to collapse in the winding number changes, because it is easily verified
position space (see Dalfovo et al., 1999, Sec. III.C), but for any that at some stage in the process (when ␦ ␻ ⫽0, in fact)
given geometry it is possible to choose g small enough that this the condensate wave function has a node at some point
does not occur. in the torus (typically where V ext is largest). Note, how-

Rev. Mod. Phys., Vol. 73, No. 2, April 2001


Anthony J. Leggett: Bose-Einstein condensation in the alkali gases 335

ever, that for small ␻ the first term (⫺ប ␻ c cos ␹) domi- We get a rough idea of the degree of metastability by
nates the other two, so cos ␹ ⯝1 and L⯝0; we do get a repeating the above argument replacing the p state by
(nearly complete) Hess-Fairbank effect, though it is not the state l ⫽n, so that for ␻ ⫽0 the coefficient of cos ␹
complete (i.e., ␳ s ⫽ ␳ ) because of the term in sin ␹. in Eq. (6.24) becomes n 2 ប 2 /2mR 2 . Then the term in V o
The opposite case, g/V o ⬎1, is more interesting. In can be neglected to a first approximation, and we find
this case, for ␦ ␻ ⫽0, there is still a stationary point of (a) that the allowed values of L/N are always very close
E( ␹ ) at ␹ ⫽ ␲ /2, but it is a maximum; the minima occur to nប, and (b) that the condition for metastability of the
at the pair of points ␹ ⫽sin⫺1 (Vo /g). If we now make ␦␻ state with winding number n when the container is sta-
nonzero, one minimum is pushed up and the other tionary is g⬎n 2 ប 2 /mR 2 . When rewritten in terms of the
down, and the barrier between them disappears entirely superfluid velocity v s ⬅nប/mR of the metastable state
when and the velocity of sound c s ⬅ 冑g/m, this reads v s ⬍c s ,
which is just the famous Landau criterion for the meta-
兩 ␦ ␻ 兩 2/3⫽g 2/3⫺V o2/3 . (6.26) stability of superfluid flow.
Actually, the Landau prediction has been verified ex-
Thus, if for example we start rotating with ␻ ⫽ ␻ c and perimentally in 4He only in the prima facie rather differ-
gradually decrease the rotation, the system will stay in ent case of the mobility of ions moving through the liq-
the ␹ ⬎ ␲ /2 minimum not only in the regime where it is uid (Allum et al., 1977; for an analogous experiment in
the globally stable one ( ␻ ⬎ ␻ c /2) but also for a finite the alkali-gas case, see Chikkatur et al., 2000); in a tor-
range of ␻ below ␻ c /2 where it is globally unstable. oidal geometry the critical velocities observed in prac-
However, because of our choice (6.21) of the orders of tice are often orders of magnitude less than the Landau
magnitude of the parameters, there is no metastability in value (see Donnelly, 1967, Sec. 2.9). It is suspected that
this case for ␻ ⫽0. Note that in the context of the in most cases the reason for the discrepancy has to do
present discussion the case of the noninteracting Bose with the presence of a ‘‘tangle’’ of vortices formed in
gas in an exactly cylindrical geometry (g⫽0, V o ⫽0) is nonequilibrium processes during the quench through the
rather pathological: it shows a discontinuous L( ␻ ) curve BEC transition temperature. However, under favorable
but no hysteresis. conditions it is possible to observe an apparently intrin-
The above argument needs to be supplemented in one sic decay of superflow even at velocities less than c s
important respect: In Eq. (6.23) it is implicitly assumed (Kukich et al., 1968). The classic paper on the theory of
that all the relevant states of the system are of the this process is that by Langer and Fisher (1967); they
Gross-Pitaevskii (coherent) type, i.e., involve only postulate that the mechanism involves the formation of
simple BEC. But what of the role of possible Fock a vortex ring and its thermally activated expansion (a
states, that is, those of the form process that is apparently not possible for the simple toy
model discussed above). Related considerations have
⌿ Fock ⬃ 共 a o⫹ 兲 M 共 a ⫹
1 兲
N⫺M
兩 vac典 , (6.27) been given for the alkali gases by Mueller et al. (1998).
One further question concerns the possible role of the
where M and N⫺M are both macroscopic (general hyperfine degree of freedom. As regards the Hess-
BEC)? Actually it is straightforward to show that in the Fairbank effect, there seems no reason to assume that
limit N→⬁ the expectation value of the interaction term this makes any appreciable difference. However, as re-
is identical to that in the Gross-Pitaevskii state with the gards the question of metastability of superflow, Ho
same value of 具 a o⫹ a o 典 , while the expectation value of the (1982) obtained the following surprising and beautiful
asymmetry energy corresponding to Eq. (6.19) is zero. result: Consider a system of bosons each with total
Thus for any given value of ␦␻ the ‘‘best’’ Gross- atomic spin F and with a Hamiltonian that is invariant
Pitaevskii state [i.e., that corresponding to the choice under spin rotation in a toroidal geometry which is ap-
⌬ ␸ ⫽0 and the best choice of ␹ in Eq. (6.23)] always proximately cylindrically symmetric. Then for any given
does better than any Fock state. It is amusing that the value of the winding number n there exists a path that
fact that the interaction energy has a nonmonotonic be- changes n by 2F and involves no energy barrier; i.e.,
havior not only on the coherent path (6.23) but on the under zero rotation no state with n⬎F can be meta-
Fock path (6.27) is a direct consequence of the factor of stable. For further discussion of this result and its pos-
2 discussed at length in Sec. IV. sible applications to the alkali gases, and also for further
discussion of the Landau criterion, see Leggett (2000a).
3. Further remarks
E. Real-life BEC alkali gases in harmonic traps
In real experiments with liquid helium, and probably
though not certainly future experiments with the BEC The analysis of this section so far (which largely rests
alkali gases in a toroidal geometry, the (repulsive) inter- on concepts developed in the context of liquid helium)
action is typically strong enough that g is large com- needs to be modified in two ways before it can be ap-
pared to both ប ␻ c and V o (which may, however, be mu- plied to a realistic alkali-gas problem (even if we neglect
tually comparable). Under these conditions, it is the hyperfine degree of freedom, on which see above).
intuitively rather obvious that even for ␻ ⫽0, not only The first has to do with the geometry of realistic traps,
the p state but many higher-l states will be metastable. which is usually of the anisotropic simple harmonic-

Rev. Mod. Phys., Vol. 73, No. 2, April 2001


336 Anthony J. Leggett: Bose-Einstein condensation in the alkali gases

oscillator type.35 A number of papers in the literature m F ⫽⫹1 ( 兩 2 典 ) (see Sec. VII.B). As mentioned in Sec. V,
have studied the properties of the solutions of Eq. (6.9) the equilibrium of such a binary combination in a har-
in such a geometry [with the rotation generally taken monic trap corresponds to a ring or shell of 兩 1 典 sur-
around the axis of the (not quite exact) cylindrical sym- rounding a core of 兩 2 典 . Following a suggestion of Will-
metry]; the most detailed known to me is that of Butts iams and Holland (1999), Matthews et al. (1999) were
and Rokhsar (1999; see also, for example, Benakli et al., able by an appropriate rotation and detuning of the laser
1999). In contrast to the case of a narrow annulus, in this field to produce a single vortex in an outer component
geometry a major role is played by centrifugal effects; in that could be either 兩 1 典 or 兩 2 典 at will; the presence of the
fact, for a rotational velocity ␻ larger than the radial vortex was confirmed by measuring the relative phase of
frequency ␻ r of the trap, the gas flies apart. For ␻ ⬍ ␻ r the two condensates as a function of angle by an appro-
the behavior is qualitatively reminiscent of the hysteretic priate ␲/2 pulse (see Sec. VII.E). As the authors ob-
regime of the toy model of Sec. VI.D: as ␻ inreases from serve, it is noteworthy that in this experiment, while the
zero, the equilibrium value of L jumps between different effective equilibrium geometry of the 兩 2 典 species is es-
smooth curves L l ( ␻ ) characterizing different numbers sentially that of the harmonic trap, that of the 兩 1 典 species
l of vortices, which in general (for l ⫽1) are not situ- resembles more the toroidal (annular) geometry of Sec.
ated at the center of the trap. As a result, the angular VI.D. It is therefore perhaps not surprising that the ob-
momentum per particle is in general not integral. For served lifetime of a vortex in the 兩 1 典 species was consid-
general ␻ there is usually some degree of hysteresis in erably longer than that of one in 兩 2 典 .
the transitions between the states corresponding to dif- A second experiment, that of Madison et al. (2000),
ferent l ; however, for ␻ ⫽0 none of the finite-l states is adopts a scheme closer to that envisaged in Secs. VI.B
metastable, because it turns out always to be energeti- and VI.C, with a single species in a slightly anisotropic
cally advantageous to move a vortex out towards the harmonic trap uniformly rotated. The vortices were de-
edge of the condensate, where it eventually disappears. tected by observing (after trap expansion) the ‘‘holes’’ in
For details the reader is referred to Butts and Rokhsar the density distribution due to their cores. As the angu-
(1999) and the references cited therein. For new effects lar velocity of rotation is varied from zero up to the
which may arise when the healing length becomes com- radial trap frequency, Madison et al. report observation
parable to or larger than the cloud size, see Wilkin and of first no vortices, then successively one, two, three, and
Gunn (2000). four (at higher velocities the system flies apart); the gen-
The second respect in which the considerations devel- eral pattern is similar to the predictions of Butts and
oped so far in this section need qualification concerns Rokhsar (1999). Very recently, the same group has
the effects of the kinetics (see Sec. II.E). In liquid 4He, a made direct measurements of the angular momentum of
strongly interacting system, the characteristic time scale the system as a function of the angular velocity of the
for collisions of an atom both with other atoms and (in- trap (Chevy et al., 2000) and in particular confirmed that
elastically) with the container walls is so small that it has below the critical value of ␻ for the entry of the first
very little effect on the decay of a metastable state, vortex the system appears to be at rest in the laboratory
which is totally dominated by energetic considerations. frame, as the above discussion would lead us to expect.
By contrast, in the alkali gases it is entirely conceivable
that even though a particular process may be ‘‘downhill
all the way’’ energetically, the unstable state still appears VII. BEC IN A TWO-STATE SYSTEM: JOSEPHSON-TYPE
stable over times of the order of perhaps seconds, be- EFFECTS, PHASE DIFFUSION
cause of the difficulty of energy exchange (and also ex-
change of angular momentum) not just between atoms In traditional condensed-matter physics, the term
but, even more importantly, with the container. As an Josephson effect is used generically to refer to a situation
example, if we have created, by rotation, a vortex at the in which a large number N of identical bosons are
center of a harmonic trap and then, by stopping the ro- restricted to occupy the same two-dimensional single-
tation, rendered it unstable, it may still persist for mac- particle Hilbert space. Well-known examples are
roscopically long times (Fedichev et al., 1999; cf. Mat- the eponymous effect,36 originally predicted by Joseph-
thews et al., 1999). son (1962) for superconductors, and the related effect
which occurs when two baths of 4He (or 3He) are con-
nected through a narrow channel, or superleak (Ander-
F. The experimental situation

Recently two experimental groups have reported the 36


observation of vortices in a BEC alkali gas. Matthews It is a matter of purely historical interest that in the original
realization of the Josephson effect in superconductivity (a) the
et al. (1999) started with 87Rb in a pure F⫽1, m F ⫽⫺1
‘‘bosons’’ (Cooper pairs) are composite rather than elemen-
state ( 兩 1 典 ) and used a two-photon microwave transition tary objects, and (b) the system is coupled to external current
to produce superpositions of 兩 1 典 and the state F⫽2, leads and therefore cannot be described by the Hamiltonian
(7.5) as it stands. The closed problem described by Eq. (7.5) is
actually conceptually simpler than the original version (and
35
There are further complications when more than one hyper- nowadays can be realized in a superconducting context, e.g., in
fine species is involved. See Sec. VI.F. mesoscopic systems).

Rev. Mod. Phys., Vol. 73, No. 2, April 2001


Anthony J. Leggett: Bose-Einstein condensation in the alkali gases 337

son, 1966; Avenel and Varoquaux, 1988). I will call these ␩ ⬅2N ⫺2 Tr ␳ˆ 21 ⫺1, (7.2)
effects ‘‘external,’’ since the two single-particle states in
question are spatially separated. A condensed-matter where ␳ˆ 1 is the (2⫻2) single-particle density matrix; ␩
example of an ‘‘internal’’ Josephson effect, in which the evidently takes values from 0 to 1.
two states differ in some internal (nongeometrical) A complete set of number-conserving Hermitian op-
quantum number, is the longitudinal nuclear magnetic erators describing the N-particle system is the set
resonance in superfluid 3He-A (Leggett, 1975; Wheat-
1
ley, 1975); in this particular case the relevant quantum M̂⬅ 共 a ⫹ a ⫺a ⫹
2 a2兲,
number is the hyperfine (nuclear spin) index of a Coo- 2 1 1
per pair. A general review of the Josephson effect in
these and other physical systems is given by Barone 1
(1999). Q̂ 1 ⬅ 共 a ⫹ a ⫹a ⫹
2 a1兲, (7.3)
2 1 2
In the context of the BEC alkali gases, the interest of
the generic Josephson situation is twofold: it can serve 1
as at least a starting model for a number of situations of Q̂ 2 ⬅⫺ i 共 a ⫹ ⫹
1 a 2 ⫺a 2 a 1 兲
practical experimental interest (see Sec. VII.B), but per- 2
haps even more importantly, it provides a simple and and thus the Hamiltonian can be expressed as a function
sometimes exactly soluble toy model for the examina- of these three operators. At this point it is convenient to
tion of questions (such as corrections to the Gross- make a definite choice of basis. In most Josephson-type
Pitaevskii description, or the generation and destruction problems there is an intuitively natural choice; for ex-
of BEC) that arise in a less tractable form in the general ample, in an extended problem the natural basis is one
theory of BEC in a spatially nonuniform system. in which 兩 1 典 and 兩 2 典 refer to the states physically local-
A. General formulation: choice of basis ized in the two bulk reservoirs, while for an internal
effect (see below) it is the two single-particle hyperfine
We consider a system of N identical bosons, each of eigenstates in the absence of laser coupling. Formally,
which is restricted to occupy a Hilbert space spanned by we can often (though not always) select a basis by re-

two orthonormal eigenvectors 兩 1 典 and 兩 2 典 , which for the quiring (a) that any terms of fourth order in the a 1,2 that
moment we choose arbitrarily; we denote the corre- 2
are not simply proportional to M be small and not in-
sponding single-particle creation operators a ⫹ ⫹
1 ,a 2 . The volve M, and (b) that there be no term linear in Q̂ 2 (this
Hilbert space of the N-particle system is thus N⫹1 di- fixes the relative phase of the basis vectors 兩 1 典 and 兩 2 典 ).
mensional and spanned by the eigenvectors in the Fock With this specification, the most general Hamiltonian
basis can be written in the form (up to terms that are func-

冋冉 冊冉 冊册 ⫺1/2
tions only of N and hence constant)
N N
兩M典⬅ ⫹M ! ⫺M !
2 2 1
Ĥ⫽ KM̂ 2 ⫺⌬ ␮ •M̂⫺EJ Q̂ 1 ⫹f 共 M̂,Q̂ 1 ,Q̂ 2 兲
⫻共 a⫹ N/2⫹M ⫹ N/2⫺M
兩 vac典 , 2
1 兲 共a2 兲

M⫽⫺N/2,⫺N/2⫹1 ¯N/2, N even: ⫹ 兺


i,j⫽1,2
␰ ij Q̂ i Q̂ j , (7.4)

⫽⫺N/2⫺1/2,⫺N/2⫹1/2 ¯N/2⫺1/2, N odd, where f contains terms of third order and higher in its
(7.1) arguments and all ␰ ij are ⰆK. In a number of cases of
where 兩 vac典 denotes the vacuum; it is clear that M is the practical interest the last two terms can be either ne-
(nearest integer to)37 half the difference in the numbers glected or incorporated into a shift in the origin of M̂
of particles in states 1 and 2. This model is then isomor- (see below), and in this case I shall call the resulting
phic to the N⫹1-dimensional representation of the expression the ‘‘canonical Josephson Hamiltonian’’: ex-
group SU(2). plicitly (in the former case),
It is clear that for such a system the occurrence of
BEC as defined in Sec. III.A is automatic; the only ques- 1 ⌬␮ ⫹
Ĥ canon ⫽ K 共 a ⫹ ⫹
1 a 1 ⫺a 2 a 2 兲 ⫺
2
共 a 1 a 1 ⫺a ⫹
2 a2兲
tion is whether it is simple or general. A useful invariant 8 2
measure of the degree of simplicity (coherence) of the
BEC is the quantity EJ ⫹
⫺ 共 a a ⫹a ⫹
2 a1兲. (7.5)
2 1 2
37 We shall see below that the canonical form (7.5) lends
This choice of definition for M, while technically conve-
itself to a particularly simple analysis, while preserving
nient, means that some of the formulas below, which are writ-
ten explicitly for even N, need correction by factors ⬃(N ⫺1 ) most (though not all) of the features of the most general
for odd N. This affects none of the conclusions and will not be problem. It is amusing that the Ĥ eff used in Sec. VI.D
noted explicitly below. [Eq. (6.22)] is, up to a constant, equivalent to Eq. (7.5)

Rev. Mod. Phys., Vol. 73, No. 2, April 2001


338 Anthony J. Leggett: Bose-Einstein condensation in the alkali gases

although the interpretation of the various terms there is approximation but the approximation of it by Eq. (7.5)
different from that discussed below. is likely to fail badly, both for the (relatively well-
known) reason that for the single well ␮ is a highly non-
B. Realizations in the BEC alkali gases linear function of N and for the (less widely appreciated
but at least equally important) one that once M be-
The scheme developed in Sec. VII.A can form, at
comes comparable to N the relevant WKB integral B,
least, a starting point for the discussion of two effects in
the BEC alkali gases that may be reasonably character- and hence the tunneling amplitude EJ , is substantially
ized, respectively, as an external and internal Josephson modified from its M⫽0 value.40
effect. However, it should be emphasized that in each Turning now to the internal Josephson effect, the
case the simple canonical Hamiltonian (7.5) is appli- nearest approximation to this realized to date in the
cable, if at all, only for a very restricted region of the BEC alkali gases seems to be the experiments of Hall
experimentally accessible parameter space.38 Space pre- and co-workers (Hall, Matthews, Ensher, et al., 1998;
cludes a discussion here of the issues involved in the Hall, Matthews, Wieman, and Cornell, 1998) in which a
justification of Eq. (7.5), so I shall simply state the re- set of 87Rb atoms in a (single-well) magnetic trap were
sults and some caveats, and refer the reader to Sols driven by a microwave coupling into a linear superposi-
(1999), Leggett (1999a, 2000a), and references therein tion of two different hyperfine states and the degree of
for further details; see also Meier and Zwerger (2000). phase coherence between them measured by a Ramsey-
Let us start with the external effect. The simplest way fringe technique (see below for details). To analyze this
to create the necessary ‘‘two-well’’ potential would seem experiment along the lines of Sec. VII.A, we for the
to be to use the initial setup of the MIT interference moment pretend that the equilibrium condensate wave
experiment (Andrews et al., 1997; see Sec. 7.5), with an function ␹ o (r) is independent of the hyperfine index,
elongated cylindrical trap split down the center by a and moreover note that, since the microwave source in-
strongly detuned laser beam, and to turn the laser power jects very large numbers of photons, the electromagnetic
down to a value substantially less than that used in the field can be treated as a completely classical object.
published experiments39 so that tunneling between the Then we can identify the two basis states 兩 1 典 and 兩 2 典 of
two wells becomes non-negligible. Very crudely speak- Sec. VII.A with (say) the F⫽1, m F ⫽⫺1 and F⫽2, m F
ing, we expect the single-particle tunneling matrix ele- ⫽⫹1 hyperfine states, respectively. Since the laser fields
ment to be of order ប ␻ o exp⫺B, where ␻ o is a typical are strongly time dependent, it is useful in this case to
in-well frequency and B an appropriate WKB exponent consider the Hamiltonian in the rotating frame, that is,
which is Ⰷ1 for a high laser barrier. One’s immediate the frame in which the laser field is constant over the
instinct is that it should be possible to handle the two- time of the pulse (see, for example, Cohen-Tannoudji,
well problem by a simple application of the time- 1992). Within the standard rotating-wave approximation
dependent Gross-Pitaevskii equation (see, for example, (ibid.) this Hamiltonian can then be cast, for any one
Ostrovskaya et al., 2000), and indeed for sufficiently pulse, into the canonical form (7.5), with the identifica-
strong tunneling this is almost certainly a good approxi- tions
mation. However, it is of interest to consider also the
opposite limit BⰇ1. In that case it is plausible to trun- ⌬ ␮ ⫽⫺ ␦ ⫹ 共 4N ␲ ប 2 /m 兲 ˜␩ • 共 a 11⫺a 22兲 , (7.6)
cate the possible states of the atoms to a two-
dimensional subspace corresponding to the self- EJ ⫽ប⍀ R , (7.7)
consistent ground states in the two wells separately,
which we identify with the basis states 兩 1 典 , 兩 2 典 of Sec.
4␲ប2
VII.A. Then, provided that (a) any bias between the two K⫽ ˜␩ 共 a 11⫹a 22⫺2a 12兲 , (7.8)
wells is very small compared to the barrier height, and m
(b) the typical asymmetry M is very small compared to where
the total number N, it is possible to justify the canonical
Hamiltonian (7.5), with EJ an appropriate single-atom
tunneling amplitude through the barrier, ⌬ ␮ the bias,
and K essentially the ‘‘bulk modulus’’ ⳵ ␮ / ⳵ N for a
˜␩ ⬅ 冕␹兩 o共 r 兲兩
4
dr. (7.9)

single well, which up to a numerical factor is given by In Eqs. (7.6) ␦ is the detuning of the lasers from reso-
ប ␻ o ␭ 2 /N where ␭ is defined in Eq. (5.52). Once one nance (appropriately corrected for the laser-induced
goes to larger values of the bias and/or asymmetry, the shifts of the atomic levels), ⍀ R is the standard two-
generic two-state Hamiltonian (7.4) may still be a good photon Rabi frequency, and a ij is the s-wave scattering
amplitude of hyperfine species i and j (see Sec. IV.E; for
details, see Leggett, 2000a).
38
This situation is to be contrasted with that obtained in tra-
ditional condensed-matter systems, where physical consider-
ations generally restrict the motion to the regime when Eq. 40
It may be shown that in the WKB limit (BⰇ1) the M de-
(7.5) is an excellent approximation. pendence of EJ resulting from this effect is much greater than
39
Subsequent experiments may in fact have attained this re- the ‘‘kinematic’’ dependence explicitly appearing in Eq. (7.21)
gime (see Rohrl et al., 1997). below.

Rev. Mod. Phys., Vol. 73, No. 2, April 2001


Anthony J. Leggett: Bose-Einstein condensation in the alkali gases 339

The canonical Hamiltonian (7.5), with time- With this mapping we see that in the internal Josephson
independent parameters given by Eqs. (7.6)–(7.9), is ad- effect the laser coupling given by Eq. (7.10) is equivalent
equate to describe the behavior during a single mono- to the application of a magnetic field along an axis in the
chromatic laser pulse. In real life, both the amplitude of xy plane whose orientation angle relative to the x axis is
the laser field and its phase ␨ (t) in the rotating frame given by the phase ␨ (t). This angular momentum repre-
may vary in time.41 In this case, if one wishes to use a sentation of the BEC two-state system is exact and is
time-independent choice of axes, it is necessary to gen- sometimes quite helpful to one’s intuition. Notice in par-
eralize the term in EJ in Eq. (7.4) to read ticular that in a physical situation where the quantity M
⫺EJ 共 t 兲关 cos ␨ 共 t 兲 •Q̂ 1 ⫺sin ␨ 共 t 兲 Q̂ 2 兴 is well defined (or has small fluctuations relative to N),
(7.10) the quantity Ĵ 2x ⫹Ĵ 2y is fixed at the value (N/2⫹1)⫺M 2 ;
关 ⬅⫺EJ 共 t 兲共 e i ␨ (t) a ⫹
1 a 2 ⫹H.c. 兲兴 . thus, while the phase angle of the xy component of an-
The above schematic description is unfortunately still gular momentum may or may not be well defined (see
too simple to apply to most existing experiments. In real below), its magnitude certainly is, and in particular for
life not only is the external potential (and hence the M⬵0 is close to N/2. This feature is a direct result of the
condensate wave function, even in the absence of inter- BEC; it is easy to see that in a system described by many
actions) different for the two species, but according to different single-particle labels ␭ in addition to the two-
the results of Sec. V.E the condensate wave functions state one (e.g., a noncondensed mixture of two hyperfine
will themselves be functions of M. In the adiabatic limit, species) the angular momentum J whose components
where the Rabi frequency ⍀ R is small compared to the are defined by
characteristic inverse ‘‘rearrangement time’’ of the con- 1 ⫹
densates, it should still be possible to use the generic J i⬅ 兺␭ J i␭ , J x␭ ⬅ 共 a 1␭
2

a 2␭ ⫹a 2␭ a 1␭ 兲 , etc. (7.12)
two-state Hamiltonian (7.4) [but not Eq. (7.5)], with the
dependence on M given simply by that of the global does not have a corresponding property, since the differ-
ground-state energy for that M value. However, in ex- ent J␭ can interfere destructively.
isting experiments (Hall, Matthews, Wieman, and Cor- Since the magnitude N/2 of the angular momentum is
nell, 1998) we are quite far away from the adiabatic fixed, the motion is, intuitively speaking, described by
limit.42 only two variables corresponding to its direction, and it
For completeness it should be mentioned that there is is tempting to ask whether they can be chosen to be
one other experiment in the literature that is sometimes canonically conjugate. The most obvious suggestion is to
described as being of the Josephson type (in our classi- try to define an operator ␸ˆ corresponding to the relative
fication, of the external variety), namely, that of Ander- phase, i.e., in the angular momentum picture to the
son and Kasevich (1998), which detects the coherence angle of Ĵ in the xy plane, such that (a) ␸ˆ is canonically
between the output beams from an array of wells at dif-
conjugate to Ĵ z , i.e., it satisfies the commutation relation
ferent heights in the Earth’s gravitational field. In this
experiment the system is not closed, so it does not fall 关 Ĵ z , ␸ˆ 兴 ⫽⫺i (7.13)
under the general scheme of Sec. VI.B, and I shall there-
fore not discuss it here. (The analysis is in fact quite and (b) the coherent (Gross-Pitaevskii) states, that is,
straightforward and is given in the cited reference.) states of the form
⌿ N ⫽ 共 N! 兲 ⫺1/2共 cos ␹ e i ␸ /2a ⫹
1 ⫹sin ␹ e
⫺i ␸ /2 ⫹ N
a 2 兲 兩 vac典
C. Kinematics of the Josephson effect: the Rabi, (7.14)
Josephson, and Fock regimes (where we denote the polar angle by ␹ rather than the
conventional ␪ to avoid subsequent confusion), are as
We first note that the three operators (7.3), which nearly as possible eigenstates of ␸ˆ with eigenvalue ␸. Is
form a complete set for our system, satisfy the commu- this possible?
tation relations of the angular momentum components The difficulties that arise in defining a relative phase
J i , and thus we have an exact mapping to the problem operator having the above properties are closely analo-
of a particle of (large) spin J in an arbitrary J-dependent gous to the well-known problems that arise in quantum
potential: optics in defining an absolute phase operator for a
single-mode electromagnetic field; see, for example,
M̂→Ĵ z ,Q̂ 1 →Ĵ x ,Q̂ 2 →Ĵ y ,N/2→J. (7.11)
Mandel and Wolf (1995), Sec. 10.7, and references cited
therein. However, generally speaking, those problems
41 become acute only when one is discussing states contain-
For example, the experiments of Hall, Matthews, Wieman,
and Cornell (1998) use two pulses of approximately equal am-
ing no or a few photons, and in our case it is important
plitude but shifted by a phase ␸ in the rotating frame. In this to remember that we are interested in large values of the
case EJ (t) is of course zero outside the pulses, and we can take total particle number N. Thus it is plausible that we will
␨ (t)⬅0 during the first pulse and ⬅ ␸ during the second. be able to define a satisfactory relative phase operator
42
This is at least partly due to the anomalously small value of provided that (i) the amplitude, in the state we are con-
the quantity (a 11⫹a 22⫺2a 12) for 87Rb and is not necessarily a sidering, of states with N/2⫺ 兩 M 兩 ⱗN 1/2 is small, and (ii)
generic characteristic of such experiments. we are prepared to neglect effects of relative order N ⫺1/2

Rev. Mod. Phys., Vol. 73, No. 2, April 2001


340 Anthony J. Leggett: Bose-Einstein condensation in the alkali gases

TABLE II. Correspondence between Josephson parameters in different superfluid systems.

System⫹effect ⌬␮ K EJ
4
He/BEC gases, external gravitational/magnetic potential bulk modulus tunneling through aperture/barrier
Superconducting grains voltage Coulomb interaction tunneling through aperture/barrier
3
He, internal (longitudinal NMR) external magn. field spin polarizability nuclear dipole force
BEC gases, internal (rotating frame) detuning difference in inter- and laser coupling
intra-species interaction

or smaller. Under these conditions, the most appropriate J x ⫾iJ y ⫽ 冑共 N/2兲 2 ⫺Ĵ 2z •exp⫾i ␸ˆ (7.20)
procedure43 is probably an adaptation of that of Carru-
thers and Nieto44 for the absolute case. Namely, we de- without worrying about the ordering of the operators Ĵ z
fine our operator Ê⬅exp i␸ˆ by and ␸ˆ . While the above discussion could no doubt be
expanded, it should be adequate for the purposes of this
Ê⬅ 关共 N/2⫺Ĵ z 兲共 N/2⫹Ĵ z ⫹1 兲兴 ⫺1/2共 Ĵ x ⫹iĴ y 兲 (7.15) section. To conclude it, we note that a Fock (fixed-M)
so that state can be generated (up to normalization) from a co-
herent state of the form (7.14) by multiplying by
Ê ⫹ ⬅ 共 Ĵ x ⫺iĴ y 兲关共 N/2⫺Ĵ z 兲共 N/2⫹Ĵ z ⫹1 兲兴 ⫺1/2. (7.16) exp(⫺iM␾) and integrating over ␾ from zero to 2␲.
That is, a Fock state can be regarded as a superposition
As in the quantum-optical case, the operator Ê⬅e i ␸ˆ so of coherent states with random phase.
defined is not unitary, so that ␸ˆ cannot be Hermitian. The commutation relation (7.13) together with the
However, we can if desired define sine and cosine op-
Hamiltonian Ĥ(M̂, ␸ˆ ) gives a complete description of
erators that are Hermitian (cf. Mandel and Wolf, 1995),
our problem, which is now one dimensional. In the most
and in any case the lack of unitarity in the present con-
text is not too serious; we have in fact general case the function Ĥ(M̂, ␸ˆ ) is arbitrary (within
certain generic constraints such as Hermiticity, etc.), but
ÊÊ † ⬅1, Ê † Ê⬅1⫺ 兩 N/2典具 N/2兩 ⫺ 兩 ⫺N/2典具 ⫺N/2典 . for the canonical case described by Eq. (7.5) it takes a
(7.17) relatively simple form: using Eqs. (7.3) and (7.19) we
Thus, provided that the state we are considering has have
only a small component of the extreme states, M⫽
⫾N/2, we may treat ␸ˆ itself as effectively Hermitian.
Further, it may be verified by direct calculation that we
1
Ĥ⫽⫺⌬ ␮ •M̂⫹ KM̂ 2 ⫺E J
2
冑 1⫺
4M̂ 2
N2
cos ␸ˆ ,

have (independent of this condition) (7.21)


E J ⬅NEJ .
关 Ĵ z ,Ê 兴 ⬅ 关 Ĵ z ,exp i ␸ˆ 兴 ⫽exp i ␸ˆ , (7.18) The Hamiltonian (7.21) can be applied to a variety of
Josephson-type situations both in the traditional super-
which is consistent with Eq. (7.13), and furthermore that
fluids (superconductors and the two isotopes of He) and
just as in quantum optics the large-amplitude coherent
in the BEC alkali gases. In Table II I tabulate the physi-
states are approximate eigenstates of the (absolute)
cal meaning of the various parameters for some of these
phase operator, so in our case the coherent states of the
cases. As we saw in the last subsection, in the case of the
form (7.14) satisfy the relation
external Josephson effect in an alkali gas, the truncation
Ê⌿ N ⫽exp i ␸ •⌿ N ⫹ ␦ •⌿⬜ (7.19) of the general form (7.4) to the special case described by
Eq. (7.5) and hence Eq. (7.21) is justified, if at all, only
when ⌿⬜ is a normalized state orthogonal to ⌿ N and the
amplitude ␦ is of order N ⫺1/2 provided that 兩 sin ␪兩 for small values of M̂; the same is actually true for the
ⰇN⫺1/2 [i.e., condition (i) above is fulfilled]. Thus, sub- external effect in 4He, and even for the internal effect in
3
ject to conditions (i) and (ii), our definition (7.15) of ␸ˆ He, but in these cases the physical conditions are al-
indeed satisfies the requirements (a) and (b). It is clear most invariably such that this is the only relevant case.
that to the same accuracy we can write the intuitively The generic behavior of an alkali-gas BEC system de-
plausible relation scribed by the Hamiltonian (7.5) [or the equivalent for-
mulation via Eq. (7.14)] and the commutation relation
(7.11) has been discussed in a number of papers in the
43 literature.45 It is useful for our purposes to distinguish
A possible alternative approach (Javanainen and Ivanov,
1999) is to take over the definition of phase used by Pegg and three principal regimes, which are characterized by dif-
Barnett (1988) to the two-state problem. While this automati- ferent orders of magnitude of the ratio K/E J , and to
cally guarantees the Hermiticity of ␸ˆ , the relation to the co- attach names to them (Leggett, 1999a):
herent state (7.14) is not so direct.
44
While these authors also discuss the status of an angle vari-
45
able, it is the angle of a coordinate vector rather than of the These include, for example, Milburn et al., 1997; Smerzi
angular momentum itself, and we therefore cannot simply take et al., 1997; Javanainen and Ivanov, 1999; Leggett, 1999a;
over their discussion verbatim. Marino et al., 1999; Raghavan et al., 1999; Sols, 1999.

Rev. Mod. Phys., Vol. 73, No. 2, April 2001


Anthony J. Leggett: Bose-Einstein condensation in the alkali gases 341

• (1) ‘‘Rabi,’’ K/E J ⰆN ⫺2 , simply that of a simple pendulum. Within this (K/E J
• (2) ‘‘Josephson,’’ N ⫺2 ⰆK/E J Ⰶ1, ⰇN ⫺2 ) parameter regime the Josephson and Fock re-
• (3) ‘‘Fock,’’ 1ⰆK/E J . gimes are distinguished by the relative importance of
quantum fluctuations, which may be measured (see be-
We examine below the ways in which the behavior of low) by the rms value of ␸ in the ground state of Eq.
the system differs qualitatively in these three regimes. (7.14); in the Josephson limit this is Ⰶ1, and as we shall
First, however, it is useful to enquire in what regime see this means that the pendulum can for most (though
we are likely to find ourselves for the various realiza- not all) purposes be treated classically, while in the Fock
tions of the Josephson effect listed above. It turns out limit it is Ⰷ1 (i.e., ⬃2 ␲ , since ␸ is by definition a peri-
that the traditional superfluids lie overwhelmingly in the odic variable) and, conversely, the spread in M̂ in the
Josephson regime, though the Fock regime is attainable, ground state is Ⰶ1, i.e., the ground state corresponds
for instance, in ultrasmall superconducting grains; the approximately to the Fock state (relative number eigen-
Rabi regime is never attainable. As regards the external state) (a L ) (a R ) 兩 vac典 . Note that the simplicity pa-
† N/2 † N/2

effect in the BEC alkali gases, it follows from the con- rameter ␩ defined in Eq. (7.2) is 1 in the extreme Rabi
sideration of Sec. VI.B that the Rabi regime is never limit and close to 1 in both the Rabi and Josephson
attainable under conditions in which Eq. (7.20) remains regimes, but 0 in the Fock limit for M⫽0. The same
valid. Since the order of magnitude of the quantity K is considerations may be applied qualitatively to the biased
given by the single-well bulk modulus, while the Joseph- case, where ⌬ ␮ ⫽0, or to the case 兩 M 兩 ⬃N/2 [but (N/2
son energy E J is exponentially sensitive to the barrier ⫺ 兩 M 兩 ⰇN 1/2)]; the obvious procedure in this case is to
height (see, for example, Zapata et al., 1998), it is rela- replace M̂ by M̂ ⬘ ⬅M̂⫺M eq , where M eq is the equilib-
tively easy to adjust the latter so as to go continuously rium value of M defined by ⌬ ␮ /K; evidently the com-
from the Josephson to the Fock regime; it appears prob- mutation relation of M ⬘ with ␸ is still the canonical one,
able that the initial state in the MIT interference experi- Eq. (7.11), and the Josephson coupling energy may be
ment (Andrews et al., 1997) corresponded to the ex- expanded in M̂ around M eq rather than around zero.
treme Fock limit (though see footnote 39). The only point to watch is that if ⌬ ␮ /KⲏN/2, so that
The most interesting case is that of the internal effect the equilibrium value of M̂ is close to N/2 (or if we start
in the BEC alkali gases. Since in this case the Josephson from such a state), then it will be impossible to neglect
coupling E J is solely due to the action of the microwave the M̂ dependence of the Josephson coupling even if for
pumps, we can trivially attain the Fock regime simply by ⌬ ␮ ⫽0 we are well on the Josephson side.
switching the pumps off. As to the Rabi regime, from To summarize the conclusions of the above discussion
Eqs. (7.6) and the results of Sec. V.E we see that its in slightly different language: The Rabi regime corre-
attainment requires, crudely speaking, the condition sponds to a pendulum with a variable length depending
⍀ Rⲏ ␻ ¯ (N ⑀ a s /d) 2/5, where ⑀ is the factor by which the on its angular momentum (see below and Marino et al.,
difference a 11⫹a 22⫺2a 12 is reduced from a typical value 1999), while in the Josephson and Fock regimes the
a s of the a ij . Unfortunately, the expression on the right- length is fixed. The Fock regime corresponds to a
hand side of this inequality is just of the order of the strongly quantum pendulum, while in the Rabi and Jo-
spatial rearrangement time, so that according to the dis- sephson regimes the behavior is (semi)classical. Accord-
cussion in Sec. VI.B the spatial degrees of freedom can- ingly, the theory of the Josephson effect in the Joseph-
not be ignored. It seems that this was the situation dur- son parameter regime which corresponds to most of its
ing the pulse sequences in the experiment of Hall et al. traditional realizations in condensed-matter physics is
(Hall, Matthews, Ensher, et al., 1998; Hall, Matthews, just the theory of a simple classical pendulum (though
Wieman, and Cornell, 1998). see below). As we have seen, the external effect in the
Let us now examine the ways in which the behavior BEC alkali gases may correspond to either the Joseph-
differs qualitatively in the three different regimes de- son or the Fock regime, and the internal effect to any of
fined above, starting with the case of an unbiased sys- the three regimes, although in the Rabi regime the
tem, ⌬ ␮ ⫽0, and a starting value of M that is small com- simple two-state model is inadequate.
pared to N/2. We then see that the main difference The most detailed discussion known to me in the ex-
between the Rabi regime on the one hand and the Jo- isting literature that covers the whole of the semiclassi-
sephson (and Fock) regimes on the other lies in the cal region (Rabi and Josephson regimes) is that of
source of the principal dependence on M: for the Rabi Raghavan et al. (1999), while an extended discussion of
case this comes principally from the factor the simple-pendulum region that allows for Fock-limit
冑1⫺(4M̂) 2 /N 2 in the Josephson coupling energy (E J ), effects such as recurrences is given by Milburn et al.
while for the Josephson and Fock cases it arises mostly (1997). Here I shall concentrate on the Josephson and
from the explicit dependence of the interaction term Fock regimes and just mention the extra effects that
1 2 arise as we go towards the Rabi limit.
2 KM̂ . Consequently, in the extreme Rabi regime we
can neglect the latter term and the behavior will be ap- D. The Josephson regime: Josephson resonance and
proximately that of a simple set of Bose-condensed two- macroscopic quantum self-trapping
level systems; conversely, far into the Josephson (or
Fock) regime we can neglect the M̂ dependence of the The general mapping of the Josephson effect to the
Josephson term, and the Hamiltonian then becomes problem of a pendulum is obtained by the replacements

Rev. Mod. Phys., Vol. 73, No. 2, April 2001


342 Anthony J. Leggett: Bose-Einstein condensation in the alkali gases

␸ˆ → ␪ˆ , (7.22) The ground state of the pendulum problem is not of this


nature; rather, it is a coherent superposition of states
M̂→L̂/ប, 兩 ␹ , ␸ 典 with ␹ approximately ␲/4 but ␸ varying over a
range of order ␸ rms , i.e.,
N→2L max /ប,
E J →mg l , ⌿ N⫽ 冕␺ ␸共 兲兩 ␸ 典 d ␸ ,

K→ 共 m l 2 兲 ⫺1 ⬅I ⫺1 , 1
(7.25)
␺ 共 ␸ 兲 ⬃exp⫺ 共 ␸ 2 / ␸ rms
2
兲.
⌬ ␮ →A ␪ /ប, 4

where l is the length of the pendulum and ␪ˆ the angle We see therefore that once we introduce interactions
made with the vertical, and A ␪ is a fictitious ‘‘vector the simple Gross-Pitaevskii description of the ground
state cannot, strictly speaking, be consistent. Of course,
potential,’’ which couples to the angular momentum.
provided we are well on the Josephson side of the
Thus in the Josephson or Fock regimes we have in the
Josephson-Fock boundary (so that ␸ rms Ⰶ1) this does
pendulum analogy, adding for convenience a c-number
not usually matter very much, because typically physical
term proportional to A 2␪ , quantities depend on ␸ only on the scale of unity (or ␲)
关 L̂⫺A ␪ 共 t 兲兴 2 and so the spread in ␸ does not affect them. However, as
Ĥ→ ⫺mg l cos ␪ˆ , we have seen above, replacement of the true ⌿ N by the
2I Gross-Pitaevskii form (7.24) (with ␹ , ␸ ⫽ 0) would lead
关 L̂, ␪ˆ 兴 ⫽⫺iប. (7.23) to a serious overestimate of 具 M̂ 2 典 and thus of quantities
such as the dephasing rate, which may depend on the
We specialize to the case of the Josephson regime and latter (see Sec. VII.E).
moreover assume A ␪ ⰆL max , i.e., the bias, if any, is such The same remarks apply to the dynamics generated
that the equilibrium value of M is much less than N by the Hamiltonian (7.21). In the Josephson-Fock re-
(‘‘nearly symmetric’’ case). We have dropped the L̂ de- gime with A ␪ ⫽0, the quantum-mechanical equations of
pendence of the potential term, so that the Hamiltonian motion for the operators L̂ and ␪ˆ are
is, as already noted, that of a simple (but in general
6 ⫽⫺ ⳵ H/ ⳵ ␪ˆ ⫽⫺mgl sin ␪ˆ ,
L (7.26)
driven) pendulum. For the moment we do not replace
the quantum-mechanical commutator by a classical Pois-
son bracket. ␪N ⫽ ⳵ H/ ⳵ L̂⫽L̂/2I. (7.27)
Let us first put A ␪ ⬅0 and study the quantum- Generally speaking, a necessary (and usually suffi-
mechanical ground state. We assume that the rms value cient) condition to replace the operators L̂ and ␪ˆ by the
of ␪ will be small and subsequently confirm that this is corresponding classical variables is that 具 sin ␪ˆ 典⬵sin具␪ˆ 典,
consistent. In that case, the Hamiltonian effectively re- where the pointed brackets indicate averages over the
duces to that of a simple harmonic oscillator with fre- quantum-mechanical wave function ␺ ( ␪ ) [i.e., ␺ ( ␸ )].
quency ␻ J , thus V( ␪ )⬵ 21 m ␻ 2J ␪ 2 where ␻ J ⬅ 冑g/l, or in Provided only that the spread of ␸ is not much greater in
the language of our original problem ប ␻ J ⬅ 冑E J K; ␻ J is order of magnitude than the ground-state rms spread
known (because of its original realization in the super- ␸ rms , this condition is automatically fulfilled in the Jo-
conducting problem) as the ‘‘Josephson plasma reso- sephson regime, and thus we can usually treat the dy-
nance frequency.’’ The zero-point excursion in ␪ is of namics semiclassically in this regime. However, this ar-
order (ប/m ␻ J ) 1/2, which in the original language is sim- gument does not, as it stands, cover the case in which,
ply 冑K/E J . Provided therefore that we are indeed in the for instance, the initial state is nearly a pure Fock state
Josephson regime, the rms value of ␪ (⫽ ␸ ) is indeed (eigenstate of M̂) and thus has a spread ⬃2 ␲ in ␸, and
Ⰶ1; in fact, we have ␸ rms ⬃(K/E J ) 1/4, (M) rms further arguments are necessary in such a case to justify
⬃(E J /K) 1/4; note that provided we are well out of the a semiclassical approximation; see Sec. IX.E.
Rabi regime, this latter quantity is still much smaller Once we have established that in the Josephson re-
than the value (⬃N 1/2) it would have for a Bose conden- gime the dynamics are (usually) those of a simple clas-
sate of simple two-state systems—a point that is of cru- sical pendulum, we can immediately apply our classical
cial importance in discussing the nonequilibrium dynam- intuition. However, because the most directly observ-
ics that arise when E J changes as a function of time; see able quantity in the BEC case (the quantity M, which
Sec. VII.E below. can be measured, for example, by optical absorption
The latter point may be put in a different way: The techniques; see Sec. II) corresponds to the angular mo-
ground state of the interacting system is not a simple mentum rather than the displacement of the pendulum,
Gross-Pitaevskii state. Such a state may be written in the the results are sometimes quite striking. Consider first
general case in the form (7.14), that is, the small oscillations about equilibrium; if for present
⌿ N ⫽ 关 a †1 cos ␹ exp共 i ␸ /2兲 ⫹a †2 sin ␹ exp共 ⫺i ␸ /2兲兴 N 兩 vac典 purposes we neglect quantum fluctuations, the latter
corresponds to ␸ ⫽0 ( ␪ ⫽0). Taking the classical version
⬅兩␸典. (7.24) of Eqs. (7.26) and (7.27), we see that the small oscilla-

Rev. Mod. Phys., Vol. 73, No. 2, April 2001


Anthony J. Leggett: Bose-Einstein condensation in the alkali gases 343

tions are sinusoidal with the Josephson frequency ␻ J E. The Fock regime: phase diffusion
⬅(KE J ) 1/2; they should in principle be observable as a
sinusoidal oscillation of the atomic density between the One advantage of studying BEC in a two-state system
two wells (or two hyperfine states). is that it allows us to formulate, in a very clean and
A more interesting experiment is one in which we ini- unambiguous way, the problem of phase diffusion. In
tially hold ⌬ ␮ fixed at a small but finite value and then, this context ‘‘phase’’ always means relative phase. For
at time zero, reduce it to zero. It is easy to verify that the present purposes I shall consider only the very simplest
equilibrium ground state for t⭐0 has ␸ ⫽0 but M possible case, a system that is for the duration of the
⫽⫺⌬ ␮ /K⫽0, i.e., in the pendulum analogy, it has finite relevant part of the experiment in the extreme Fock
canonical angular momentum [but zero angular velocity limit (i.e., the Josephson coupling E J is simply set to
␻, since we recall that in the presence of a finite A ␪ ␻ is zero) and where the initial (and conserved) expectation
given not by L/I but by (L⫺A ␪ )/I]. For t⭓0 the dy- value of the relative number M is zero. The dispersion
namics are given by the classical version of the (un- in M is for the moment taken to be ⰆN but otherwise
arbitrary. Such a situation is realized in the ‘‘Ramsey-
forced, that is A ␪ ⫽0) Eqs. (7.26) and (7.27), i.e., the
fringe’’ experiment of Hall, Matthews, Wieman, and
simple pendulum equations, but the initial condition is
Cornell (1998); in this case, the two states in question
that L(t⫽0⫹)⫽A ␪ (so that the pendulum is given an
are two different hyperfine states of 87Rb. A superposi-
impulsive torque at t⫽0), or in the original BEC lan- tion state of the type (7.28) below is created by the ini-
guage M(t⫽0⫹)⫽⫺⌬ ␮ /K. It is clear from our intu- tial ␲/2 two-photon pulse47; the system is then allowed to
ition regarding the classical pendulum that there is a evolve freely (E J ⫽0) until a second ␲/2 pulse, arbi-
critical value of L(t⫽0⫹) or equivalently a critical trarily phased with respect to the first. Finally, the num-
value M c of M, given by M c ⫽ 冑2E J /K [note that in the ber of atoms in each of the two hyperfine states is mea-
Josephson region, by definition, this value is much sured.
greater than 1, and in fact much greater than M rms Consider initially a normalized Gross-Pitaevskii-type
⬃(E J /K) 1/4, but much smaller than N.] For M⬍M c state of the form [a special case of Eq. (7.14)]
(L⬍L c ) the pendulum performs simple librational (os-
关 2 ⫺1/2共 e i ␸ /2a ⫹
1 ⫹e
⫺i ␸ /2 ⫹ N
a 2 兲兴
cillatory) motion, which for LⰆL c is nearly simple har- ⌿ N⫽ 兩 vac典
monic at the Josephson plasma frequency ␻ J . As M 冑N!


approaches M c from below the motion becomes strongly
N N
anharmonic, and for M⬎M c the pendulum rotates in- ⫹M! ⫺M!
2 2
stead of librating. It is clear that in the rotating state the
angular momentum stays always positive definite,46 and
⬅ 兺
M N!
exp iM ␸ 兩 M 典 . (7.28)
analogously, for the Josephson system M does the same,
In this state the dispersion of M is clearly of order N 1/2.
i.e., the system never reaches the equilibrium value It is convenient to introduce the operator
(zero) of M (in the absence of damping). This behavior
has been seen spectacularly in the longitudinal nuclear 2
Q̂⬅ 共 Q̂ 1 ⫹iQ̂ 2 兲 ⬅a ⫹
1 a 2 / 共 N/2 兲 共 ⬵ exp i ␸ˆ 兲 .
magnetic resonance of 3He-A (Wheatley, 1975, Sec. N
VIII). In the case of the BEC alkali gases, it has not at (7.29)
the time of writing been seen experimentally but has The Gross-Pitaevskii state (7.28) is an approximate
been analyzed in the theoretical literature under the eigenstate of Q̂, with eigenvalue exp i␸; equivalently, in
name ‘‘macroscopic quantum self-trapping.’’ the 3D angular momentum picture of Eqs. (7.11), the
There are a number of other interesting phenomena angular momentum lies in the xy plane at an angle ␸
that can be seen in the semiclassical regime of the Jo- with respect to the x axis (for example, if the magnetic
sephson effect, particularly near the Rabi-Josephson field corresponding to the initial ␲/2 pulse lies along the
boundary: see Smerzi et al. (1997) and Marino et al. x axis, J lies initially along the y axis).
(1999). Before proceeding to calculate the subsequent devel-
The question of dissipation in the Josephson effect is opment of the expectation value 具 Q(t) 典 of Q̂, it is im-
of considerable interest, both in its own right and as a portant to discuss the relation of this quantity to what is
possible guide to the more general characteristics of re- actually measured experimentally. At this point it may
laxation in a BEC system. However, to discuss it we be helpful to make a brief digression to discuss the cel-
clearly have to go beyond the simple conservative ebrated MIT interference experiment (Andrews et al.,
Hamiltonian (7.4), for instance, by introducing extra de- 1997), since although the latter involves some complica-
grees of freedom corresponding to the normal compo- tions that are not present in the toy version of the
nent (see, for example, Ruostekoski and Walls, 1998; Ramsey-fringe experiments that I shall consider, the is-
Zapata et al., 1998; Meier and Zwerger, 2000). I shall sues concerning the relation of what is calculated to
therefore not attempt to discuss it here.

47
As discussed in Sec. VII.C, during the pulse itself the sys-
46
In the classical approximation. tem is in the Rabi limit.

Rev. Mod. Phys., Vol. 73, No. 2, April 2001


344 Anthony J. Leggett: Bose-Einstein condensation in the alkali gases

what is actually measured is essentially identical. The and since the functions ␺ L (r,t) (obtained by solving the
ensuing discussion is highly schematic and is essentially time-dependent Gross-Pitaevskii equation) correspond
a synopsis of considerations already thoroughly dis- approximately, at least locally, to oppositely propagating
cussed in the existing literature on this experiment (e.g., plane waves of relatively well-defined wavelength ␭, the
Javanainen and Yoo, 1996; Naraschewski et al., 1996; last term in Eq. (7.34) produces density oscillations of
Castin and Dalibard, 1997; Hegstrom, 1998); it makes no
period ␭/2, leading to the corresponding fringes in the
claim to any originality.
At the initial stage of the experiment, BEC is pro- optical transmission pattern. Note that if the initial con-
duced separately in two sets of atoms, confined in two ditions on each run of the experiment are identical
separate traps; the number in each trap is macroscopic (something that would be difficult to achieve in practice)
(⬃5⫻106 ) and approximately, though in practice not then not only the spacing but the ‘‘offset’’ (absolute po-
exactly, equal. In the published experiments the (laser- sition) of the fringe pattern should be reproducible from
induced) energy barrier between the two traps is so high run to run.
that the probability of tunneling between them should Now, what happens if the initial state is (as it almost
be completely negligible, so that the system is firmly in certainly was in the published experiments) the Fock
the extreme Fock limit according to the classification of state (7.30)? If we carry out a direct calculation of
Sec. VII.C and, if we denote the single-atom ground 具 ␳ (r,t) 典 on the basis of the many-body wave function
state in the left and right wells by ␺ L and ␺ R , respec- (7.31), we find the rather disappointing result
tively, the initial many-body state should be schemati-
具 ␳ 共 r,t 兲 典 Fock ⫽N L 兩 ␺ L 共 r,t 兲 兩 2 ⫹N R 兩 ␺ R 共 rt 兲 兩 2 (7.35)
cally of the form
with no interference term. Yet the photographs ob-
⌿ N ⬃ 关 ␺ L 共 r兲兴 N L 关 ␺ R 共 r兲兴 N R , N L ⫹N R ⫽N, (7.30) tained in the experiments show spectacular interference
corresponding to a definite number N L or N R of atoms fringes! (See Andrews et al., 1997, Fig. 2.)
in the left or right well, or in the language of Sec. VII.A The resolution of the apparent paradox lies in the ob-
to a Fock state. The overlap of ␺ L (r) and ␺ R (r) is com- servation that the predictions of quantum mechanics al-
pletely negligible. Next, the barrier is removed, so that ways refer to statistical averages over many experimen-
the atoms can expand freely into the region between the tal runs. In condensed-matter physics one is used to the
two traps where they can overlap. At this stage, and idea that when the quantity one is measuring is macro-
before any measurement is made, a natural schematic scopic and is the sum of contributions from many micro-
description is of the form scopic entities (in this case, the atoms) then it is usually
adequate to replace averages over runs by averages over
⌿ N ⬃ 关 ␺ L 共 r,t 兲兴 N L 关 ␺ R 共 r,t 兲兴 N R , (7.31)
atoms, i.e., to assume that the result obtained on any
where ␺ L (r,t) is the time-dependent state that evolves one run will be close to the run-averaged result. But this
from ␺ R (r) under the action of the time-dependent assumption may fail spectacularly if there is a strong
Gross-Pitaevskii equation.48 In general, ␺ L (r,t) and degree of correlation between the atoms—and the phe-
␺ R (r,t) will both be nonzero in the overlap region be- nomenon of BEC represents in some sense the strongest
tween the wells. The final step is to measure, by optical correlation conceivable, since every atom must be doing
means, the atomic density ␳ (r,t) in this region, and the exactly the same thing at the same time! Indeed, when
interesting question is: Do we see fringes (density oscil- applied to the experiment of Andrews et al. (1997) the
lations) corresponding to interference between the two quantum-mechanical prediction (7.35) is entirely
condensates? correct—but all that it tells us is that if we take a pho-
If we had started not from the Fock state (7.30) but tograph of the density distribution on each individual
from a coherent state of the form run of many runs and then lay the photographs on top of
one another, the pattern will be correctly described by
⌿ N ⬃ 关 a ␺ L 共 r兲 ⫹b ␺ R 共 r兲兴 N 共 兩a兩⬃兩b兩 兲, (7.32)
Eq. (7.35). It tells us nothing (much) about what we will
then after the expansion we would have schematically see in each individual photograph, and to find this out it
instead of Eq. (7.31) the state is necessary to calculate higher-order correlation func-
tions such as the quantity
⌿ N ⬃ 关 a ␺ L 共 r,t 兲 ⫹b ␺ R 共 r,t 兲兴 N . (7.33)
P 共 rr⬘ :t 兲 ⬅ 具 ␳ 共 r,t 兲 ␳ 共 r⬘ ,t 兲 典 . (7.36)
In such a state the expectation value of the density
␳ (r,t) is If we calculate P on the basis of the wave function
(7.31), we find that it has strong oscillations as a function
具 ␳ 共 r,t 兲 典 coh ⫽ 兩 N 关 兩 a 兩 兩 ␺ L 共 r,t 兲 兩 ⫹ 兩 b 兩 兩 ␺ R 共 r,t 兲 兩
2 2 2 2
of z⫺z ⬘ , with period ␭/2 (where z is the relevant coor-
⫹2 Re兵 ab ␺ L
* 共 r,t 兲 ␺ R 共 r,t 兲 其 兴 , (7.34) dinate, roughly speaking that corresponding to the axis
separates the two wells). Thus the conditional probabil-
ity that if one sees a high atomic density at position z
48
This result is of course only approximate, and in particular one will also see a high density at a point n␭/2 away
ignores important questions concerning the effect of inter- [and a low one at (n⫹1/2)␭/2] is high. One can evi-
atomic interactions in the overlap region (which fortunately dently extend the argument to higher correlation func-
turns out to be fairly negligible), but it will be adequate for our tions, and the net result is that quantum mechanics
present purpose. makes a unique prediction about the fringe spacing but

Rev. Mod. Phys., Vol. 73, No. 2, April 2001


Anthony J. Leggett: Bose-Einstein condensation in the alkali gases 345

no prediction about the offset. What this means is that necessarily well defined and whose apparent value may
on each run we expect to see a definite fringe pattern, therefore vary from run to run (see the discussion above
with spacing ␭/2 but with random offset. This is exactly of the MIT interference experiment). Then, if the prob-
what is seen in the experiments. When the photographs ability of the system realizing a given value of ␪ at time
from many different experiments are superimposed, the t is P t ( ␪ ), the probability of obtaining a given value of
randomness of the offsets means that the average den- J z 关 ⬅ 21 (n 1 ⫺n 2 ) 兴 after a second pulse at time t with
sity thus seen conforms to Eq. (7.35), i.e., shows no phase angle ␦ relative to the first will be proportional to
fringes. An alternative way of describing the situation P t ( ␪ o ) where cos(␪o⫺␦)⫽Jz , and the expectation value
that may make the analogy to the Ramsey-fringe case of J z will be


clearer is that any Fock state of the form (7.30) may be
represented as a linear superposition of coherent states 具 J z 典 ⫽const d ␪ P t 共 ␪ 兲 cos共 ␪ ⫺ ␦ 兲 . (7.39)
of the form (7.32) with equal amplitudes for each value
of the relative phase of a and b, and that according to Consider a series of experiments such that we first select
Eq. (7.34) measurement of the atomic density consti- a given value of t and of ␦, measure the value of J z on
tutes a measurement of this relative phase; then accord- each of an ensemble of runs with this t and thus a value
ing to the measurement axioms of quantum mechanics of 具 J z 典 (t), then choose a new value of t (but the same ␦)
such a measurement must have a definite outcome, but and repeat the procedure, and so on. The possible be-
that outcome is random from run to run and unpredict- havior varies between two extreme cases:
able in advance. [For a further discussion of this point in
the general Josephson context, see Hegstrom and Sols (a) If ␪ (t) is well defined and processes with angular
(1995), and for the specific application to BEC alkali frequency ␻, then the value J z ⬅ 21 i 具 (n 1 ⫺n 2 ) 典 mea-
gases see the references cited above.] sured on each run of the ensemble of experiments
Returning now to the Ramsey-fringe experiment, the conducted with time delay t is unique and given by
second ␲ /2 pulse, at time t, corresponds to application the expression
of a pulsed magnetic field along some axis in the xy J z 共 t 兲 ⬅ 具 J z 共 t 兲 典 ⫽ 共 N/2兲 cos共 ␻ t⫺ ␦ 兲 . (7.40)
plane rotated relative to that of the first pulse by some
phase shift ␦, and this is then followed by a measure- (b) If ␪ (t) is completely random (ill defined), then the
ment (by optical absorption) of the relative population value of J z measured on a given run of the relevant
of states 兩1典 and 兩2典, i.e., of J z . Suppose first that at time ensemble should vary at random from ⫺N/2 to
t the system is still in an eigenstate of Q̂ with eigenvalue ⫹N/2, and the run-averaged value of 具 J z 典 (t) should
exp i关␸⫹␪(t)兴, i.e., the angular momentum has simply be zero. In the intermediate case 具 J z 典 (t) should be
precessed around the z axis through an angle ␪ (t). Then given, crudely speaking, by a formula of the type
it is clear that on each run of an ensemble of experi- A(t)(N/2)cos(␻t⫺␦) where the envelope function
ments involving the same initial conditions and the same A(t), which under appropriate conditions is slowly
time difference t, the measured value of J z will be re- varying over a period 2 ␲ / ␻ , is less than 1 and is
producible and equal to cos关␸(t)⫺␦兴, 关 ␸ (t)⬅ ␸ ⫹ ␪ (t) 兴 equal under the stated conditions to 兩 具 Q̂(t) 典 兩
[cf. the predictions above for the coherent case de- ⫽ ␩ 1/2(t).
scribed by Eq. (7.32)].
However, in general we shall find that the behavior of Let us now turn to the calculation of 具 Q̂(t) 典 , and thus
of ␩ (t). Consider an initial state that is a generalization
具 Q̂(t) 典 does not correspond to a unique precession.
What does this mean for the expected experimental re- of Eq. (7.22), namely,
sults? We have seen in Sec. VII.C that provided the sys-
tem behaves as a simple BEC two-state system, the op- ⌿ N共 0 兲 ⫽ 兺
M
␭ M exp共 iM ␸ 兲 兩 M 典 , (7.41)
erator J 2x ⫹J 2y always has under these conditions a value
of approximately (N/2) 2 , which means that the magni- where the coefficients ␭ M are real, positive, slowly vary-
tude of Q(t), could it be directly observed in the experi- ing between neighboring values of M, such that the dis-
ment, should always be 1: persion in M is ⰆN, and satisfy the normalization con-
dition 兺 M ␭ M2
⫽1. These conditions are satisfied, for
具 Q̂ † 共 t 兲 Q̂ 共 t 兲 典 ⫽1. (7.37) example, by a normalized Gaussian distribution with
On the other hand, the calculation below will show that width ⰆN (but Ⰷ1). Because of the slow variation con-
in general dition, ⌿ N (0) is an approximate eigenstate of Q̂ with
eigenvalue exp i ␸ , and ␩ (0)⫽1.
兩 具 Q̂ 共 t 兲 典 兩 2 ⫽ ␩ 共 t 兲 ⬍1, (7.38)
Consider now the time evolution of 具 Q̂ 典 and hence of
where to obtain the equality we used the definition (7.2) ␩. For orientation let us first consider the case in which
of the degree of coherence ␩ (t), Eq. (7.29), and the fact all interactions are small, so that the Hamiltonian is
that 具 J z 典 ⫽0. What this means is that on any given run of simply ⫺⌬ ␮ •M. In the following, it is convenient to
the experiment the system behaves ‘‘as if’’ the (vector) work in the rotating frame as defined by the initial and
angular momentum lies in the xy plane and has its full final ␲ /2 pulses, so that ⌬ ␮ is fairly small. It is immedi-
magnitude N/2, but with a phase angle ␪ (t) that is not ately clear that the phase accumulated over time t by the

Rev. Mod. Phys., Vol. 73, No. 2, April 2001


346 Anthony J. Leggett: Bose-Einstein condensation in the alkali gases

state 兩 M 典 is just ⌬ ␮ •Mt, so that the many-body state at as recombination and trap evaporation. What happens if
time t is just of the form (7.28) with ␸ replaced by we relax these assumptions?
␸ ⫹⌬ ␮ •t. Thus we have To the extent that the environment can be modeled
by a classical random field which couples linearly to M̂
具 Q̂ 共 t 兲 典 ⫽exp i 共 ␸ ⫹⌬ ␮ t 兲 (7.42)
关 H sys⫺en v ⬃M̂V(t), where V(t) is random], it is clear
and ␩ (t) stays equal to unity for all t. that it gives rise to dephasing such that

冏 冕 冏
Next let us consider the effect of a term in Ĥ of the 2
t
form 21 KM̂ 2 . The effect of this is to add to the evolution ␩ 共 t 兲 ⫽ exp⫺i V 共 t 兲 dt/ប , (7.47)
of the state 兩 M 典 an extra phase ⫺ 21 KM 2 t. However, o

since all contributions to 具 Q̂ 典 come from the matrix el- where the bar indicates an average over the form of
ements between states 兩 M 典 and 兩 M⫹1 典 , it is the deriva- V(t); the same should be true within a quantum-
tive of this expression that enters 具 Q̂(t) 典 : mechanical description of the environment, if the cou-
ˆ ⍀̂ with ⍀̂ some operator of the
pling is of the form M
具 Q̂ 共 t 兲 典 ⫽exp i 共 ␸ ⫹⌬ ␮ t 兲 兺 ␭ M
2
exp⫺iKMt. (7.43) latter and the average is replaced by a trace. A second
M
example of dephasing that may be important in practice
Suppose the initial packet is a Gaussian of width is the uncertainty, in a given experimental run, in the
(⌬M) o , i.e., value of N and hence in the constant K. Finally, a more
subtle example is recombination (Sinatra and Castin,
1
␭ M⬃ exp⫺M 2 /4共 ⌬M 兲 o2 . (7.44) 1998). Let us suppose for the moment that a single
⌬M o l -body recombination process takes place at definite
Then the initial time dependence of 具 Q̂(t) 典 can be ob- time t⫺T (where t is, as above, the time of the second
tained by replacing the sum over M by an integral: ␲ /2 pulse and is fixed by the experimenter). The effect
of this recombination is to replace each of the coeffi-
具 Q 共 t 兲 典 ⫽exp i 共 ␸ ⫹⌬ ␮ •t 兲 exp⫺K 2 共 ⌬M 兲 o2 t 2 (7.45) cients C M (t⫺T) of the state 兩 M 典 by C M⫹m (t⫺T). This
and so means that the relative phase of neighboring C M at time
t is of the form
␩ 共 t 兲 ⫽exp⫺2K 2 共 ⌬M 兲 o2 t 2 . (7.46)
⌬ ␸ M ⫽ 共 ⌬ ␮ ⫺KM 兲共 t⫺T 兲 ⫹ 关 ⌬ ␮ ⫺K 共 M⫺ l 兲兴 T
Thus the degree of coherence ␩ (t) falls to zero over a
time scale (call it ␶ decoh ) of the order of ប/(K•⌬M o ). ⫽ 共 ⌬ ␮ ⫺KM 兲 t⫹ l KT, (7.48)
Since in an experiment of the JILA type ⌬M o ⬃N 1/2 i.e., it adds an M-independent term to ⌬ ␮ M , whose
and K is proportional, in a harmonic trap, to N ⫺3/5, value is determined by the unknown time T. The argu-
[see Eq. (7.7)], we see that ␶ decoh actually increases with ment can obviously be generalized to the case of many
N, though only as N 1/10. Since, as we have seen, K is of recombinations; although it is classical in the sense that
order E times the mean-field energy, which is typically the decays are conceived as taking place at definite (but
of order 20 kHz, we see that for a sample of, say, random) times, it seems extremely probable that a fully
106 atoms ␶ decoh can be of the order of a few seconds.49 quantum-mechanical treatment would give the same re-
The transition from a sum over M to an integral, how- sult.
ever, obscures one phenomenon that is of potential in- All the above considerations are confirmed quantita-
terest, namely, that of recurrences. We see from Eq. tively in an elegant calculation due to Sinatra and Castin
(7.43) that at times which are exact multiples of ប/K (1998). If ␭ ⫺1 is the characteristic time for the first re-
⬅t rec all the different states 兩 M 典 come back into phase,
combination to occur, they find, for the quantity we have
and ␩ (t) recovers its original value of unity. The quan-
tity t rec is longer than ␶ decoh by a factor (⌬M o ) ⫺1 , and called Q̂(t), the result for tⰆt c [their Eq. (64)]
thus for the JILA experiment would be of the order of
hours and almost certainly unobservable.50
All the above assumes not only isolation of the system
兩 具 Q̂ 共 t 兲 典 兩 ⫽exp⫺
␹ 2t 2 1
冉 4 2
2 ⫹ l ␭t ,
8 共 ⌬␸ 兲0 3 冊 (7.49)

from any ‘‘environment’’ but neglect of processes such where (⌬ ␸ ) 20 is the mean-square spread in relative
phase (as they define it) in the initial state. Since this
quantity is proportional to (⌬M) ⫺2 0 and the ␹ in Eq.
49
Note that had we started with a wave packet (such as the (7.49) is proportional to our K, it is clear that the first
ground state of a harmonic well) that had a finite spread in ␸
factor in the brackets is identical in structure to that in
and hence a value of (⌬M) o much smaller than N 1/2, the de-
Eq. (7.45), while the second indicates the effect of aver-
coherence time would have been much longer. Cf. Leggett and
Sols, 1998; Javanainen and Wilkins, 1997. aging over the recombination processes. Note that since
50
Inclusion in Ĥ of cubic and higher terms in M̂ will tend to in a JILA-type experiment (⌬ ␸ ) o2 ⬃N 1/2, the second
reduce the amplitude of the recurrences even for an isolated term becomes comparable to the first only at times
system, but a dimensional estimate shows that the effect is at (⬃N␭ ⫺1 ) so long that the whole sample has already
most of order unity and may be smaller. decayed appreciably.

Rev. Mod. Phys., Vol. 73, No. 2, April 2001


Anthony J. Leggett: Bose-Einstein condensation in the alkali gases 347

In the real-life JILA experiments, an important role is


played by the spatial rearrangement processes following 兺l 共 E l ⫺E 0 兲 兩 具 l 兩 ␳ q兩 0 典 兩 2 ⫽nប 2 q 2 /2m (8.3)
the initial ␲ /2 pulse. These processes are discussed by
Sinatra et al. (1999) and their effect on the phase coher- (n⬅N/V). Moreover, since the energy per unit volume
ence is noted by Sinatra and Castin (1998); the latter in the Gross-Pitaevskii ground state is simply 21 U 0 n 2 ,
conclude that an important role is played by the dephas- the compressibility ␹ 0 is just U ⫺1
0 ; writing out the
ing due to uncertainty in the initial value of N, and that second-order perturbation-theory expression for ␹ 0 , we
when this is taken into account the agreement of theory have
and experiment is reasonable.
兺l 兩 具 l 兩 ␳ q兩 0 典 兩 2共 E l ⫺E 0 兲 ⫺1 ⫽n/2U 0 . (8.4)

VIII. THE BOGOLIUBOV APPROXIMATION We can now use the Cauchy-Schwartz inequality to
derive from Eqs. (8.3) and (8.4) the result
Let us confine ourselves for the moment to the case of
具 ␳ q ␳ ⫺q 典 0 ⬅ 兺 兩 具 l 兩 ␳ q 兩 0 典 兩 2 ⭐ 共 n 2 ប 2 q 2 /2mU 0 兲 1/2⬅nq ␰
a spinless gas at zero temperature. Then we have seen in
Sec. V that in the dilute limit appropriate to the alkalis a l
very reasonable description of most of the properties, (8.5)
whether time dependent or not, is given by the Gross-
[where we used definition (5.6) of the healing length ␰].
Pitaevskii (Hartree) ansatz (5.14) for the many-body
However, we can also evaluate the quantity 具 ␳ q ␳ ⫺q 典 0
wave function, i.e., a simple product of single-particle
directly:
functions with no two-particle or higher correlations
(though see footnote 18). Now in fact it is clear that Eq.
具 ␳ q␳ ⫺q典 0 ⬅ 兺 具 a p⫹q/2

(5.14) can be regarded as the first of a sequence of trial

a p⫺q/2a p⬘ ⫺q/2a p⬘ ⫹q/2典 0
pp⬘
functions, of which the next is one that builds in two- but
not three-particle or higher correlations:
N
⫽ 兺p 具 n p⫹q/2典 0 具 1⫺n p⫺q/2典 0 ⫽n. (8.6)

⌿ N 共 ri ¯rN :t 兲 ⫽S 兿
i⬍j
␸ 共 ri ,rj :t 兲 , (8.1) For q ␰ ⭐1 the relations (8.4) and (8.5) are clearly mutu-
ally inconsistent, and the only possible conclusion is that
where S denotes the operation of symmetrization be- the Gross-Pitaevskii ground state cannot be the true
tween the particles. Just as in the Gross-Pitaevskii ap- ground state of the many-body system.
proximation all particles occupy the same single-particle Some insight as to what has gone wrong may be ob-
state, in the approximation (8.1) all pairs of particles tained from our experience in Sec. VII.D with the two-
occupy the same two-particle state. Equation (8.1) is state system in the Josephson regime. We saw there that
nothing but the particle-number-conserving version of in the presence of a term in the energy proportional to
the celebrated Bogoliubov approximation. In the litera- M 2 , it is energetically advantageous to allow the relative
ture this approximation is, almost without exception, in- phase of the condensate to fluctuate somewhat around
troduced either by explicitly relaxing the constraint of its mean value, thereby lowering 具 M 2 典 , which would
particle number conservation, as in the original work of otherwise be ⬃N. In a similar way, in the bulk system,
Bogoliubov (1947), or by writing down and analyzing one would like to build in fluctuations so as to reduce
operator equations of motion. In my view neither of the value of 具 ␳ q␳ ⫺q典 .
these techniques adequately exhibits the simple under- Another way of looking at what is essentially the same
lying structure (8.1) of the many-body wave function, so point is that, as we saw in Sec. IV.C, the total interaction
I shall take in this section a more direct approach. First, energy is proportional to the sum over i and j of the
however, I give some motivation for going beyond the ⬃

Gross-Pitaevskii approximation for the many-body quantity 兩 ⌿(r ij → 0) 兩 2 , and it is therefore energetically
ground state. favored to reduce the latter. It is intuitively plausible to
expect that the distance in 兩 r ij 兩 over which the distortion
A. Inconsistency of the Gross-Pitaevskii approximation from the simple Gross-Pitaevskii ansatz occurs will be,
just as in the case of repulsion by a single-particle poten-
Consider a system of N spinless bosons in free space. tial such as a hard wall, of the order of the healing length
A direct way of seeing that the Gross-Pitaevskii ground ␰, and we shall see that this is correct.
state cannot be the true ground state of the system is to
use the sum rules obeyed by the density fluctuation op-
B. The Bogoliubov ground state in the
erator translation-invariant case

␳ q⬅V ⫺1 兺i e iq•r ⬅V ⫺1 兺p a p⫹q/2


i †
a p⫺q/2 . (8.2) Consider the case of a system moving in a constant
external potential and subject to periodic boundary con-
The potential term in the Hamiltonian can be written ditions, so that the ground-state wave function of the
in the form const 兺 q␳ q␳ ⫺q , and thus commutes with ␳ q ; condensate is simply a zero-momentum plane wave (i.e.,
this allows us to derive the ‘‘f-sum rule’’ ␹ 0 ⫽const). The upshot of the arguments of Sec. VIII.A

Rev. Mod. Phys., Vol. 73, No. 2, April 2001


348 Anthony J. Leggett: Bose-Einstein condensation in the alkali gases

is that we need to modify the Gross-Pitaevskii ground


state so as to build in the effects of long-wavelength K⫽ 兺
q⫽0
⑀ q a q† a q , (8.10a)
(q ␰ ⱗ1) density fluctuations. Now, the simplest way (in-
deed the only way) of creating a density fluctuation of (b) the Hartree terms
wave vector q starting from the Gross-Pitaevskii ground
1 U0 1 N2
state is to take a particle out of the state 0 (⬅ ␹ 0 ) and
put it into a state q. However, a single process of this
H H⫽ † †

a a a a ⫽
2 V qq ⬘ q q ⬘ q ⬘ q 2 V o
U , (8.10b)

kind would violate conservation of momentum. The sim-


(c) the Fock terms
plest momentum-conserving procedure is to create pairs
of density fluctuations of opposite momentum, i.e., to 1 U0
operate on the Gross-Pitaevskii ground state with the H F⫽ 兺†
a†a a a ,
2 V q⫽q ⬘ q q ⬘ q ⬘ q
(8.10c)
operator a q† a ⫺q

a 0 a 0 ⬅⌳
ˆ q . Thus we should expect that a
good approximation to the true many-body ground state (d) the ‘‘pairing’’ terms
might be of the generic form 1 U0
H P⫽ 兺
a†a† a a .
2 V q⫽q ⬘ q ⫺q ⫺q ⬘ q ⬘
(8.10d)
⌿ N ⬃const 兺 c 兵n 其 兿 ⌳
j
ˆ n 兩0典,
q
j
(8.7)
j j
兵 j其
n In the following I shall assume that the total ‘‘deple-
where 兩 0 典 is the Gross-Pitaevskii ground state and the tion’’ of the condensate, that is, the quantity
c 兵 n j 其 are arbitrary complex coefficients. N ⫺1 兺 q⫽0 具 a q† a q 典 , is much smaller than 1 (this will be
Actually, however, the ansatz (8.7) turns out to be confirmed below); this then allows us consistently to ne-
more general than we need; it can in fact be proved glect, in Eqs. (8.10c) and (8.10d), all terms except those
(Leggett, 1999b: the proof is too cumbersome to be in which either q or q⬘ is zero. Then if we take the zero
given here) that in the limit N→⬁ and small depletion of energy at the Gross-Pitaevskii value 21 N 2 U 0 /V, the
Hamiltonian reduces to a sum of contributions from dif-
(see below) the member of the class (8.7) that minimizes
ferent values of q:
the energy is actually a member of a subclass, namely,
the subclass of states corresponding to the ansatz (8.1).
In second-quantized notation the latter reads in general, Ĥ⫽ 兺q Ĥ q , (8.11)
apart from normalization,

⌿ N ⫽ 共 N! 兲 ⫺1/2 冉冕 冕 dr dr⬘ Ĥ q ⬅Eq a q† a q ⫹


1 U0 † †
共 a a a a ⫹H.c.兲 ,
2 V q ⫺q 0 0
(8.12)

⫻K 共 rr⬘ 兲 ␺ † 共 r兲 ␺ † 共 r⬘ 兲 冊 N/2
兩 vac典 , (8.8)
where
Eq ⬅ ⑀ q ⫹n 0 U 0 , n 0 ⬅N 0 /V⬇N/V. (8.13)
and for the translation-invariant case we must have To calculate 具 H q 典 over the wave function (8.9) for any
K(rr⬘ )⫽K(r⫺r⬘ ). Then, taking out the macroscopically given q, we make a binomial expansion of Eq. (8.9):
occupied zero-momentum state explicitly and taking
N/2
Fourier transforms, we have (still apart from normaliza-
tion) ⌿ N ⫽N! ⫺1/2
兺 共 a 0 a 0 ⫺c q a q† a ⫺q
N/2 † †
Cm †
兲m

冉 冊
m⫽0

冉兺 冊
N/2
⌿ N ⫽N! ⫺1/2 a †0 a †0 ⫺ 兺 c q a q† a ⫺q
q⬎0

兩 vac典 , (8.9)
⫻ ⫺c k a k† a ⫺k
N/2⫺m
兩 vac典 , (8.14)
k⫽q,0
where the minus sign is introduced for subsequent con- N/2
venience. Equation (8.9) is the ground state of a with C m the binomial coefficient, and note that the sec-
translation-invariant Bose system with fixed particle ond factor contributes only an overall probability factor
to 具 H q 典 . Moreover, we can check that the only values of
number N in the Bogoliubov approximation.51
m contributing appreciably in Eq. (8.14) are close to N.
It remains only to determine the coefficients c q . To
Since the value of 具 H q 典 for the normalized state corre-
do this, we argue as follows: It is clear that the only
sponding to m is only weakly sensitive to m (see below),
terms in the second-quantized Hamiltonian that have
it follows that it is legitimate, in calculations of 具 H q 典 , to
nonzero expectation values in the state (8.9) are (ignor-
replace the full wave function (8.9) by the expression,
ing terms of relative order N ⫺1 )
now correctly normalized,
(a) the kinetic-energy terms
⌿ (q) ⫽ 共 N! 兲 ⫺1/2共 1⫺ 兩 c q 兩 2 兲 1/2共 a †0 a †0 ⫺c q a q† a ⫺q

兲 N/2兩 vac典
51
For related number-conserving formulations see, for ex- 共 兩 c q 兩 ⬍1 兲 . (8.15)
ample, Girardeau et al. (1959) and Gardiner (1997), and for
the standard (number-nonconserving) approach see Huang It is now extremely straightforward to calculate the
(1987), Chap. 13. quantity 具 H q 典 ⬅ 具 ⌿ (q) 兩 Ĥ q 兩 ⌿ (q) 典 , where Ĥ q is given by

Rev. Mod. Phys., Vol. 73, No. 2, April 2001


Anthony J. Leggett: Bose-Einstein condensation in the alkali gases 349

Eq. (8.12). Omitting the simple algebra, I just quote the C. Properties of the Bogoliubov ground state: elementary
result52 for the total Hamiltonian Ĥ: excitations

具 H 典 ⫽ 兺 兵 Eq sinh2 ␪ q ⫺n 0 U 0 sinh ␪ q cosh ␪ q 其 , Once one has the correct form (8.17) of the coeffi-
q⫽0 cients c q in the many-body ground state (8.9), the calcu-
(8.16) lation of the correlations and the elementary excitations
where I have introduced the notation c q ⬅tanh ␪q . Mini- is straightforward and the algebra involved is closely
mization of Eq. (8.16) with respect to ␪ q gives parallel to that encountered in the ‘‘textbook’’ (particle-
tanh(2␪q)⫽n0U0 /Eq , or equivalently nonconserving) approach (see, for example, Huang,
1987, Chap. 13); I therefore simply quote the results be-
1
c q⫽ 共 E ⫺E q 兲 共 ⬎0 兲 , (8.17) low without derivation.
n 0U 0 q [Note added in proof. At a very late stage in the pro-
where for subsequent convenience we introduced the cessing of this manuscript, I have realized that because
notation of ambiguities in the precise definition of the ‘‘Gross-
E q ⬅ 共 Eq2 ⫺n 20 U 20 兲 1/2⬅ 关 ⑀ q 共 ⑀ q ⫹2n 0 U 0 兲兴 1/2 Pitaevskii’’ state (cf. footnote 18), both the (counterin-
tuitive) sign of the right-hand side of Eq. (8.23) and the


⬅ ប 2 c s2 q 2 ⫹ 冉 冊册ប 2q 2
2m
2 1/2
. (8.18)
precise interpretation of Eq. (8.24) deserve considerably
more detailed discussion than given here. (For some of
the issues involved, see e.g., Cherny and Shanenko,
Insertion of Eq. (8.17) into Eq. (8.9) completely deter-
mines the many-body ground state, which, moreover, 2000.) I thank Willi Zwerger for correspondence which
from Eq. (8.16) has the energy (relative to the Gross- drew my attention to this.]
Pitaevskii ground state) (1) The depletion of the Gross-Pitaevskii ground
1 state, that is, the quantity ␨ ⬅N ⫺1 兺 q 具 a q† a q 典 , is
具 H 典 0⫽ 兺 共 E q ⫺Eq 兲 . (8.19) given by the formula
q⫽0 2
The expression (8.19) is actually linearly divergent as it 8
␨⫽ 共 n 0 a s3 兲 1/2. (8.21)
stands for large q; for the resolution of this difficulty, see 3 冑␲
below.
Since in published experiments on the alkali gases
The above derivation, which is simply the particle-
the right-hand side of Eq. (8.21) has never been
conserving version of Bogoliubov’s original argument,
much greater than 0.01, it is for most purposes an
relies on the pairing of the plane-wave states q,⫺q,
excellent approximation to ignore the depletion,
which are not identical but are related by time reversal.
i.e., set n 0 ⫽n.
With a view to the generalization to inhomogeneous sys-
(2) In evaluating 具 H 0 典 one must beware of double-
tems (Sec. VIII.D), it is interesting to note that we could
counting that part of the two-particle interaction
equally well have paired atoms in the same state (in this
energy that has already been taken into account
case sine or cosine waves). In fact, an alternative way of
in defining the s-wave scattering length (see Sec.
writing the ansatz (8.9) for the many-body wave function
IV.C). The result of this consideration is to sub-
is

冉 冊
tract from the summation in Eq. (8.19) a term
N/2
⫺n 20 U 20 /2⑀ q , after which the sum is convergent
⌿ N ⫽N! ⫺1/2 a † 0 a †0 ⫺ 兺
q⬎0
† †
c q 共 a qc a qc ⫹a qs
† †
a qs 兲 兩 vac典 ,
and equal to the expression
(8.20)
64
where †
and
a q,s †
a q,c
create particles in sine and cosine 具 H 典 0⫽ N 共 n 0 U 0 兲 • 共 n 0 a s3 兲 1/2. (8.22)
15冑␲
states; e.g., the normalized amplitude of a q,c †
兩 vac典 at r is
1/2 Thus the correction (8.19) to the Gross-Pitaevskii
2 cos q•r. The algebra then proceeds exactly as above,
the only difference being that if one considers, in the ground-state energy is a fraction of order the
formula analogous to Eq. (8.15), only (say) the sine con- depletion ␨ of the latter, and thus again negligible
tribution, the normalization factor is (1⫺ 兩 c q 兩 2 ) 1/4 rather in the alkali gases under most current conditions.
than (1⫺ 兩 c q 兩 2 ) 1/2. However, Eq. (8.16) and all subse- (3) The correction to the two-particle correlation

quent formulas follow unchanged, with the sum over 具 ␳ (r) ␳ (r⬘ ) 典 ⬃ 兩 ⌿(r ij → 0) 兩 2 also involves a sub-
(all) q replaced by a sum over the positive half-space. traction to avoid double-counting the single-pair
The coefficients c q are numerically equal to those in Eq. effects. Once this is done, the result is
(8.9), and, needless to say, the results of Sec. III.C are
unaffected. ␦ 具 ␳ 共 r 兲 ␳ 共 r⬘ 兲 典
⫽⫺const共 n 0 a s3 兲 1/2F 共 兩 r⫺r⬘ 兩 / ␰ 兲 ,
n 20
(8.23)
52
To obtain Eq. (8.13) we replaced matrix elements such as
冑N 0 (N 0 ⫹1) by N 0 . Note that the sum over q now goes over where F(x) tends to 1 for x→0 and to zero for

all nonzero q. x→⬁. Thus the quantity 兩 ⌿(r ij → 0) 兩 2 is reduced

Rev. Mod. Phys., Vol. 73, No. 2, April 2001


350 Anthony J. Leggett: Bose-Einstein condensation in the alkali gases

by a factor of order ␨ from its Gross-Pitaevskii D. The inhomogeneous case


value n(⬇n 0 ) and the effect on the two-particle
wave function extends out to a distance of the
The case of a system moving in a spatially varying
order of the healing length ␰, as anticipated.
potential is considerably more complicated, because (a)
(4) Consider the operator ␣ q⫹ defined by
there is no a priori choice for the condensate wave func-
␣ q† ⬅N ⫺1/2
0 共 u q a q† a 0 ⫹ v q a †0 a ⫺q 兲 , (8.24) tion ␹ 0 (r) and one cannot assume a priori that it will
retain its original (Gross-Pitaevskii-level) form in the
u q ⬅cosh ␪ q , v q ⬅sinh ␪ q , 共 tanh ␪ q ⬅c q 兲 . presence of pairing, and (b) there is in general no simple
(8.25) analog of the orthogonality of the states k and ⫺k to
It may be verified by direct calculation that when one another and to 0. As a result of (a), the class of wave
applied to the ground state (8.15) [with the c q functions for the many-body system that is considered in
given their ground state values (8.25)], ␣ q⫹ gener- the Bogoliubov approximation is actually a subclass of
ates an excited state with momentum បq and en- the general class (8.1) of paired states, and thus may not
ergy E q given by Eq. (8.15), while its Hermitian correspond to the absolute minimum of the energy
conjugate ␣ q gives zero. Consequently, for weak within this class; however, this effect may be shown to
excitation the Hamiltonian can be written, rela- be of higher order in na s3 than the (na s3 ) 1/2-level effects
tive to the Bogoliubov ground-state value (8.19), considered in the Bogoliubov approximation and will
in the simple form (Bogoliubov, 1946) therefore not be discussed here.
As regards point (b), it is important to realize53 that
Ĥ⫽ 兺q E q ␣ q† ␣ q . (8.26) despite those complications the many-body ground state
can still be written54 in a form that is the obvious gener-
In the particle-nonconserving approach the quan- alization of Eq. (8.17), namely,
tities N ⫺1/2
0 a 0 , N ⫺1/2
0 a †0 are commonly replaced

冉 冊
by 1.
N/2
(5) Finally, consider the superfluid density ␳ s (0) (see 1
Sec. VI). This quantity is no longer simply related
⌿ N ⫽ 共 N! 兲 ⫺1/2 a o† a o† ⫺ 兺m c a† a†
2 m m m
兩 vac典 ,
to the condensate number N o ; in fact, inspection (8.29)
of the arguments of Sec. VI.C indicates that since,
in the Hess-Fairbank effect in a cylindrical geom-
etry for small but finite ␻, the condensate remains where the single-particle states m are orthogonal to one
in its original (s) state and the virtual Bogoliubov another and to the state 0, and, if one were interested in
excitations are created out of it with zero total the ground-state energy and the structure of ⌿ N , one
(angular) momentum, the latter contribute noth- could proceed by a generalized version of the argument
ing to the circulating current, which remains zero. of Sec. VIII.B.55 However, it turns out that in general
If so, then according to the definition in Sec. VI there is no simple relationship between the basis vectors
the superfluid fraction ␳ s / ␳ should remain unity m appearing in Eq. (8.29) and the elementary excita-
even though N o is no longer equal to N. This re- tions of the system, and an alternative approach is there-
sult is in fact correct, not only within the Bogoliu- fore standard in the literature. For the details of the
bov approximation but for arbitrary interaction derivation I refer the reader to Fetter (1999) and merely
strength, provided only that perturbation theory quote the principal results.
starting from the free Bose gas converges (Ga- Relaxing (for the only time in this review) our insis-
voret and Nozières, 1964; see also Leggett, 1998). tence on exact particle number conservation,56 we write
At finite but low temperature 关 T⬅(k B ␤ ) ⫺1 ⰆT c 兴 the the Bose destruction operator in the form ␺ˆ (r)
condensate is still not much depleted and the number ⫽â 0 ␹ 0 (r)⫹ ␸ˆ (r), expand the Hamiltonian up to second
具 n q 典 ⬅ 具 ␣ q† ␣ q 典 of Bogoliubov quasiparticles is given by order in ␸ˆ (r), and treat the quantity â 0 â 0 as a c number
the Bose distribution with E q given by Eq. (8.15), with a value equal to N 0 . In this way we obtain an ef-
具 n q 典 ⫽ 关 exp共 ␤ E q 兲 ⫺1 兴 ⫺1 , (8.27) fective Hamiltonian that is bilinear in ␸ˆ and ␸ˆ † and of
the general form
and a famous argument due to Landau (see Huang,
1987, Chap. 13) gives for the normal density ␳ n (T) ⬅ ␳
⫺ ␳ s (T) 53
To my knowledge, the first place in the literature where this
1 is (implicitly) pointed out is a paper by Fetter (1972), which is
␳ n共 T 兲 ⫽
3 兺q 共 បq 兲 2共 dn q /dE q 兲 . (8.28) probably less well known in the BEC community than it de-
serves to be.
54
Provided it is real, which will be true if the Hamiltonian is
For TⰆnU o this expression is proportional to T 4 , while
invariant under time reversal.
for nU o ⰆTⰆT c it has the same temperature depen- 55
I hope to discuss this question further elsewhere.
dence as (1⫺N o /N), namely T 3/2. For T⬃T c the deple- 56
For explicitly number-conserving formulations of the ensu-
tion is substantial and the theory becomes more compli- ing argument and that of Sec. VIII.E, see Gardiner (1996) or
cated. Castin and Dum (1997).

Rev. Mod. Phys., Vol. 73, No. 2, April 2001


Anthony J. Leggett: Bose-Einstein condensation in the alkali gases 351

Ĥ⫽ 冕 dr ␸ˆ † 共 r兲 ⫺ 再 ប2 2
2m
ⵜ ⫹V ext 共 r兲 u j 共 r兲 ⬃A j 共 r兲 exp i 冉冕 冊 k j 共 r兲 dr , (8.38a)


⫹2N 0 兩 ␹ 0 共 r兲 兩 2 ␸ˆ 共 r兲 v j 共 r兲 ⬃B j 共 r兲 exp ⫺i 冉 冕 冊 k j 共 r兲 dr , (8.38b)

where the functions A j , B j , and k j are slowly varying


⫹ 共 N 0 U 0 关 ␹ 0 共 r兲兴 2 ␸ˆ † 共 r兲 ␸ˆ † 共 r兲 ⫹H.c.兲 . (8.30)
functions of r, and the fast oscillation of the phase then
One then shows that the expression (8.30) can be di- means that u j and v j are indeed approximately orthogo-
agonalized (and minimized) by the substitution nal to one another and to the slowly varying function
␹ 0 (r).
␸ˆ 共 r 兲 ⫽ 兺j 关 ␣ j u j共 r兲 ⫹ ␣ †j v j共 r兲兴 (8.31)
E. Time-dependent Bogoliubov–de Gennes equations:
where the functions u j (r), v j (r) satisfy the celebrated connection with the time-dependent
Bogoliubov–de Gennes equations Gross-Pitaevskii equation

冉 ⫺
ប2 2
2m
ⵜ ⫹V ext 共 r兲 ⫹2N 0 兩 ␹ 0 共 r兲 兩 2 ⫺ ␮ u j 共 r兲 冊 In accordance with the considerations of Sec. VIII.B,
it is convenient at this stage to redefine the Bogoliubov
⫹N 0 U 0 ␹ 20 共 r兲v j* 共 r兲 ⫽E j u j 共 r兲 , (8.32a) quasiparticle operators ␣ ⫹ j [currently defined simply by
the inversion of Eq. (8.31) and its Hermitian conjugate]

冉 ⫺
ប2 2
ⵜ ⫹V ext 共 r兲 ⫹2N 0 兩 ␹ 0 共 r兲 兩 2 ⫺ ␮ v j 共 r兲 冊 so as to make them explicitly particle conserving:
2m
⫹N 0 U 0 关 ␹ *
0 共 r 兲兴 u j 共 r 兲 ⫽⫺E j v j 共 r 兲 .
2
(8.32b)
␣ †j ⬅N ⫺1/2
0 a0 冉 冕 dr u j 共 r兲 ␸ˆ † 共 r兲 ⫹a ⫹
0 冕 冊
dr v j 共 r兲 ␸ˆ 共 r兲 .
(8.39)
Equations (8.32) do not specify the normalization of According to Eq. (8.35), states of the general form (up
the u j and v j , and it is convenient to choose the latter so to normalization)
that

冕 关 兩 u j 共 r兲 兩 2 ⫺ 兩 v j 共 r兲 兩 2 兴 dr⫽1. (8.33)
兿j 共 ␣ †j 兲 n 兩 0 典 ,
j (8.40)

where 兩 0 典 is the Bogoliubov ground state, are energy


With this choice the operators ␣ j , ␣ †j satisfy the standard eigenstates and thus time independent. However, super-
Bose commutation relations, positions of such states will in general have a nontrivial
关 ␣ j , ␣ k 兴 ⫽ 关 ␣ †j , ␣ k† 兴 ⫽0, 关 ␣ j , ␣ k† 兴 ⫽ ␦ jk , (8.34) time dependence. In particular, consider a linear super-
position of the ground state and a small admixture of
and the effective Hamiltonian takes the form different states each containing a single quasiparticle of
index j. The time-dependent wave function has the gen-
Ĥ⫽ 兺j E j ␣ †j ␣ j ⫹E 0 , (8.35) eral form

E 0 ⫽⫺ 兺j Ej 冕 兩 v j 共 r 兲 兩 2 dr. (8.36)
⌿ N 共 t 兲 ⫽␭ ⫺1/2 1⫹ 冉 兺j ⑀ j ␣ †j e ⫺iE t/ប
j
冊 兩0典, (8.41)

The u j ’s and v j ’s satisfy the orthogonality properties ␭⬅1⫹ 兺j 兩 ⑀ j兩 2 , 共 ⑀ j Ⰶ1,᭙j 兲 . (8.42)


[with the normalization (8.33)]


On the other hand, let us go back to the time-
dr关 u j* 共 r兲 u k 共 r兲 ⫺ v j* 共 r兲v k 共 r兲兴 ⫽ ␦ jk , (8.37a) dependent Gross-Pitaevskii equation (5.13), which we
recall describes (in the case N 0 ⫽const) the evolution of

冕 dr关 u j* 共 r兲v k 共 r兲 ⫺u k* 共 r兲v j 共 r兲兴 ⫽0, (8.37b)


the condensate wave function ␹ 0 (rt). As we saw in Sec.
V.C, if we linearize this equation around the static
Gross-Pitaveskii ground state, we find that we get Eq.

冕 dr关 u j 共 r兲 ␹ 0 共 r兲 ⫺ v j ␹ *
0 共 r 兲兴 ⫽0. (8.37c)
(5.16), with ⌿ 0 (rt)⬅exp⫺i␮t/ប•⌿0(r), where ⌿ o (r)
⬅ 冑N o ␹ o (r) is the ground-state form of the order pa-
rameter. But it is now a straightforward exercise (see,
Note that in general a given u j is not orthogonal to the for example, Castin and Dum, 1998; Dalfovo et al., 1999)
corresponding v j , nor are either of them separately to to show that quite generally Eq. (5.16) is diagonalized by
␹ 0 . However, there are many cases of practical interest expanding ␦ ⌿(rt) and ␦ ⌿ * (rt), respectively, in terms
in which this is likely to be true to a good approxima- of the Bogoliubov functions u j (r), v j (r). In fact,
tion; for example, if ␹ 0 (r) is slowly varying in the semi-
classical limit one can choose u and v to have the ap- ␦ ⌿ 共 rt 兲 ⫽ 兺 ˜⑀j 关 u j 共 r兲 e ⫺iE j t/ប ⫹ v *j 共 r兲 e iE j t/ប 兴 , (8.43)
proximate forms j

Rev. Mod. Phys., Vol. 73, No. 2, April 2001


352 Anthony J. Leggett: Bose-Einstein condensation in the alkali gases

suggesting a close relation with Eq. (8.41). [Equations and expands about the Gross-Pitaevskii solution dis-
(5.19) and (5.20) correspond to the special case of Eq. cussed in Sec. V.D, similarly to what is done in Sec.
(8.42) with only one pair u j , v j nonzero.] In fact, if on VIII.D:
the wave function (8.41) we evaluate the quantity
具 ␸ † (r)a 0 典 (t) [where ␸ † (r) is the part of the field opera- ␺ˆ ␣ 共 r兲 →⌿ ␣ 共 r兲 ⫹ ␸ˆ ␣ 共 r兲 . (8.46)
tor ␺ † (r) orthogonal to ␹ 0 ], we find an expression iden- Note that when we do this, the terms of the form, for
tical to Eq. (8.43) (with ⑀ j →˜⑀ j ; Castin and Dum, 1998). example,
Thus there is a one-one correspondence between the
single-quasiparticle Bogoliubov eigenstates and the small ⌿ ␣* 共 r兲 ⌿ ␤* 共 r兲 ␸ˆ ␤ 共 r兲 ␸ˆ ␣ 共 r兲 (8.47)
oscillations of the condensate. [In the free-space case we contain inter alia the zero-temperature momentum-
could already see this by a comparison of Eqs. (5.23) exchange contributions mentioned at the end of Sec.
and (8.18).] V.D.
This correspondence should not surprise us: it is a Of course, if we are interested only in obtaining the
characteristic feature of systems that are in some sense elementary excitation spectrum, it is probably easier
‘‘quasiclassical.’’ Consider, for example, our old friend simply to linearize the time-dependent Gross-Pitaevskii
the simple harmonic oscillator. We can do one of two equation (5.35) around the ground state (since we ex-
things: (a) starting from a classical-approximation pect that the correspondence between the solutions of
ground state, ␺ 0 (x)⬃ ␦ (x), we can replace the argument this equation and the elementary excitations will go
x by x⫺x 0 (t), where x 0 (t) satisfies the classical equa- through just as in the spinless case discussed in the last
tion of motion with a small amplitude [much less than subsection). This is done for example by Ho (1998) and
(ប/m ␻ 0 ) 1/2]; (b) we can start from the true quantum- Ohmi and Machida (1998). Discussion of the Bogoliu-
mechanical ground state ␺ 0 (x) and mix in a small bov theory as such for the multicomponent case has, to
amount of the first excited state ␺ 1 (x). With a suitable the best of my knowledge, been limited to the spatially
correspondence of the parameters, these two proce- uniform case (Colson and Fetter, 1978; Huang and Gou,
dures, which are the analogs of Eqs. (5.13) and (8.41), 1999).
respectively, give precisely the same value for the physi-
cal observable 具 x(t) 典 [namely, A cos(␻t)]. However, it is
IX. FURTHER TOPICS
clear that they correspond to quite different assumptions
about the actual form of the wave functions. In the same
In this section I very briefly list and comment on a
way, despite the coincidence of their predictions for the
number of topics in the theory of the BEC alkali gases
quantity 具 ␸ † (r)a 0 典 , the ansätze (5.13) and (8.41) corre-
that I have had to omit from the main body of this re-
spond to totally different assumptions about the actual
view, either for sheer lack of space or because I believe
many-body wave function (and hence different predic-
the theory is still in a state of flux and therefore feel
tions for other possibly observable quantities).
unable to say anything definitive. The list is not com-
The above discussion refers to small disturbances
plete, and in particular does not include some important
around the ground state. The question of the application
topics, such as the collective excitations of a gas in a
of the Bogoliubov approximation to situations in which
harmonic trap, which are extensively discussed in the
the condensate wave function is itself strongly time de-
review of Dalfovo et al. (1999; see also Griffin, 2000).
pendent is a good deal more delicate, and for the rea-
sons sketched in Sec. IX.E, I shall not attempt to discuss A. Attractive interactions
it here.
For some systems, in particular 7Li and 85Rb in low
magnetic fields, the s-wave scattering length is negative
and thus, by the arguments of Sec. IV, the effective in-
teratomic interaction is attractive. Under these condi-
F. The multicomponent case
tions it is believed that the system will be unstable
against collapse in real space not only in free space but
In view of the weight given in the first seven sections
in a harmonic trap provided that the total number of
of this review to the case of several co-existing hyperfine
atoms exceeds a critical value (see Dalfovo et al., 1999,
species, it seems appropriate to sketch very briefly the
Sec. III.C); this seems consistent with existing experi-
generalization of the considerations of the present sec-
ments on 7Li (Bradley et al., 1997) and on 85Rb (Cornish
tion to this case. In fact, this generalization is rather
et al., 2000). However, the kinetics of the collapse pro-
obvious: one simply starts from the Hamiltonian
cess and the nature of the final state are controversial.
Ĥ⫽ 兺
␣␤
冕 冉
dr ␺ ␣† 共 r兲 ⫺ ␦ ␣␤
ប2
2m
⫹V ␣␤
(ext)

共 r兲 ␺ ␤ 共 r兲
See, for a theoretical consideration, Kagan et al. (1998)
and Sackett et al. (1998).
(8.44)


B. Optical properties
1
⫹ 兺 U
2 ␣␤␥ ␦ ␣␤␥ ␦
␺ ␣† 共 r兲 ␺ ␤† 共 r兲 ␺ ␥ 共 r兲 ␺ ␦ 共 r兲 dr
Most of the properties of the BEC alkali gases that
(8.45) are associated with genuinely ‘‘optical’’ transitions (i.e.,

Rev. Mod. Phys., Vol. 73, No. 2, April 2001


Anthony J. Leggett: Bose-Einstein condensation in the alkali gases 353

visible-wavelength, as distinct from the microwave tran- E. Kinetics, damping, relaxation, etc.
sitions discussed in Sec. VII.D) are believed to be sensi-
tive primarily to the behavior of the local atomic den- Probably the most glaring omission in the main body
sity, and therefore do not require separate of this review has been any discussion of the strongly
consideration. However, we have already seen one ex- nonequilibrium dynamics of a Bose-Einstein condensate
ception (Sec. IV.F). A second involves the phenomenon (except in the special case of the two-state system dis-
of super-radiance (see Inouye et al., 1999). There may cussed in Secs. VII.D and VII.E); this despite the fact
be other phenomena associated with spontaneous emis- that this problem is ubiquitous in the interpretation of
sion, which as we have seen is usually totally negligible experimental work (e.g., on the kinetics of the conden-
for microwave transitions but need not be so in the op- sation process, the damping of collective excitations, and
tical regime. In this context some intriguing proposals the decay of vortex states). The reason for the omission
related to the considerations of Sec. VII.E have been is partly sheer lack of space, but partly also that despite
made (Javanainen, 1996); to date there has been no ex- a proliferation of theoretical work in this area in the last
perimental realization. four years, I still do not feel that the situation is entirely
clear. The problem is not just one of uncertainty about
C. Coexistence of three hyperfine species how well particular approximation schemes work in
practice for particular experimental setups; there also
Under very special circumstances it is possible to tune seem to me to be major conceptual issues that are not
the parameters in a trap so that not only can three or completely resolved. One particularly glaring issue is
more different hyperfine species coexist in it (Stenger how far it is necessary to maintain the conceptual dis-
et al., 1998a) but collisions of the type 兩 1 典 ⫹ 兩 ⫺1 典 → 兩 0 典 tinction between a situation in which one has complete
⫹ 兩 0 典 , in which the initial and final hyperfine states are or nearly complete Bose condensation but the state is
different, can exactly or approximately conserve energy. highly chaotic and thus fluctuates randomly from run to
Law et al. (1998) pointed out that in the case of exact run, and one in which there is genuinely no condensa-
degeneracy the ground state need not necessarily be of tion. In neither case does the single-particle density ma-
the simple Gross-Pitaevskii type but could represent a trix, as calculated by the standard techniques, possess an
sort of spatially extended pseudomolecule, in which eigenvalue of order N (in the first case because it essen-
pairs of atoms condense into a state with, for example, tially predicts the average behavior over runs), and it is
total spin zero but no definite single-atom spin. How- therefore tempting to assume that in practice the distinc-
ever, it was subsequently pointed out by Ho and Yip tion is meaningless; however, the analysis of the MIT
(2000) that this kind of state is extremely fragile and that interference experiments shows that in that context an
even tiny perturbations, such as magnetic field gradients, analysis based on the single-particle density matrix can
would be likely to restabilize the Gross-Pitaevskii be highly misleading (Javanainen and Yoo, 1996; see
ground state. This is currently a very active area of dis- also Hegstrom, 1998). This issue is rendered increasingly
cussion. urgent by the increasing popularity in the BEC field of
the so-called ‘‘phase-space technique,’’ in which the den-
D. The ‘‘atom laser’’ sity matrix of an arbitrary initial state is expressed as a
sort of pseudomixture of Gross-Pitaevskii-type states,
While the experiment of Andrews et al. (1997) dem- each individually possessing BEC but with wildly vary-
onstrates that it is certainly possible to produce two ing condensate wave functions, the time evolution of the
beams of atoms that either have a definite relative phase latter calculated, and the results put together to describe
or acquire one on measurement, there has been to date the behavior of the real system (see Steel et al., 1999,
no realization of the precise atomic analog of an optical and for an early application of similar ideas to the kinet-
laser, that is, roughly speaking, a continuous-output ics of condensation, Kagan et al., 1994). My suspicion is
beam with a well-defined direction, frequency, intensity, that, as so often happens in physics, we are dealing here
and phase. Many theoretical papers have been devoted with a trick that is extremely computationally conve-
to the design of such an atom laser and an analysis of its nient and gives physically correct results provided our
likely properties; for an up-to-date review, see Ballagh measurements are in some sense sufficiently coarse
and Savage (2000). It should be noted that the discus- grained, but may be subtly misleading under special cir-
sion of the phase coherence properties of an atom laser cumstances. However, I feel at present unable to say
does not require us to introduce the concept of absolute anything clearly on this topic, and thus in accordance
phase (though this is often done in the literature), any with the motto on my title page must ‘‘keep silent’’ and
more than does the discussion of, for instance, weak- merely refer the reader to what I regard as some of the
localization effects in solids in the context of the simple most important recent papers on this general question
one-electron Schrödinger equation; in each case it is the (as distinct from applications to specific experimental
relative phase at different times that is at issue. The situations): Castin and Dum (1998), Kagan and Svis-
points raised in these discussions have much in common tunov (1998), Lopez-Arias and Smerzi (1998), Drum-
with the problem of diffusion of relative phase discussed mond and Corney (1999), Steel et al. (1999), Stoof
in Sec. VII.D, and are also closely connected with the (1999), Walser et al. (1999), Carusotto et al. (2000), and
next issue. Gardiner and Zoller (2000).

Rev. Mod. Phys., Vol. 73, No. 2, April 2001


354 Anthony J. Leggett: Bose-Einstein condensation in the alkali gases

F. Late-breaking developments Ballagh, R. J., and C. M. Savage, 2000, in Bose-Einstein Con-


densation: from Atomic Physics to Quantum Fluids, Proceed-
The first draft of this review was completed in January ings of the 13th Physics Summer School, edited by C. M.
2000, and over the intervening year there have been a Savage and M. Das (World Scientific, Singapore).
number of interesting developments, some but not all of Barone, A., 1999, Proceedings NATO-ASI Conference on
which are mentioned in the main text. Of these develop- Quantum Mesoscopic Phemena and Mesoscopic Devices in
ments, probably the one that will most obviously require Microelectronics, edited by I. O. Kulik (Kluwer Academic,
us to extend the conceptual basis outlined in this review Dordrecht/Boston), in press.
is the attainment of unprecedentedly high values of the Benakli, M. S. S. Raghavan, A. Smerzi, S. Fantoni, and S. R.
‘‘gas parameter’’ na 3 in experiments conducted close to Shenoy, 1999, Europhys. Lett. 46, 275.
a Feshbach resonance (Cornish et al., 2000) and the pro- Bogoliubov, N. N., 1947, J. Phys. (USSR) 11, 23 (reprinted in
duction of diatomic molecules in an alkali-gas conden- D. Pines, The Many-body Problem, W. A. Benjamin, New
sate by photoassociation (Wynar et al., 2000). These two York, 1961).
developments, which from a theoretical standpoint are Bradley, C. C., C. A. C. Sackett, and R. G. Hulet, 1997, Phys.
closely related, pose a major challenge to theory and will Rev. Lett. 78, 985.
very likely require us to go considerably beyond the Burke, J. P., C. H. Greene, and J. L. Bohn, 1998, Phys. Rev.
simple contact-interaction approximation of Sec. IV. For Lett. 81, 3355.
some relevant considerations, see Holland et al. (2000) Burt, E. A., R. W. Ghrist, D. J. Myatt, M. J. Holland, E. A.
and Heinzen et al. (2000) and references cited therein. Cornell, and C. E. Wieman, 1997, Phys. Rev. Lett. 79, 337.
Butts, D. A., and D. S. Rokhsar, 1999, Nature (London) 397,
327.
ACKNOWLEDGMENTS Carruthers, P., and M. M. Nieto, 1968, Rev. Mod. Phys. 40,
411.
This review could not have been written without im- Carusotto, I., Y. Castin, and J. Dalibard, 2000, preprint,
cond-mat/0003399.
portant input from many people. I would first like to
Castin, Y., and J. Dalibard, 1997, Phys. Rev. A 55, 4330.
thank my students Sahel Ashhab, Carlos Lobo, and
Castin, Y., and R. Dum, 1996, Phys. Rev. Lett. 77, 5315.
Sorin Paraoanu, and my UIUC colleague Gordon Baym Castin, Y., and R. Dum, 1997, Phys. Rev. Lett. 79, 3553.
and his students Minqiang Li and Erich Mueller, not Castin, Y., and R. Dum, 1998, Phys. Rev. A 57, 3008.
only for many incisive comments and questions but for Cherny, A. Yu., and A. A. Shanenko, 2000, Phys. Rev. E 62,
doing their best, sometimes vainly, to keep me abreast 1646.
of the BEC literature. Of the many other colleagues, Chevy, F., K. W. Madison, and J. Dalibard, 2000, Phys. Rev.
both experimental and theoretical, with whom I have Lett. 85, 2223.
enjoyed fruitful discussions, I would especially like to Chikkatur, A. P., A. Görlitz, D. M. Stamper-Kurn, S. Inouye,
acknowledge Yuri Kagan, Wolfgang Ketterle, Sigmund S. Gupta, and W. Ketterle, 2000, Phys. Rev. Lett. 85, 483.
Kohler, Alice Sinatra, Fernando Sols, Masahito Ueda, Chu, S., 1998, Rev. Mod. Phys. 70, 685.
Ivar Zapata, and, most particularly, Yvan Castin. I am Cohen-Tannoudji, C., 1992, in Fundamental Processes in
especially grateful to the last named and, even more, to Quantum Optics, edited by J. Dalibard, J-M. Raimond, and J.
Carlos Lobo for spending many hours patiently educat- Zinn-Justin (North-Holland, Amsterdam), pp. 1–164.
ing me on the subject of the nonlinear kinetics of a BEC Cohen-Tannoudji, C., 1998, Rev. Mod. Phys. 70, 707.
system; needless to say, the responsibility for any re- Colson, W. B., and A. L. Fetter, 1978, J. Low Temp. Phys. 33,
maining confusion is entirely my own. Finally, I am in- 231.
debted to several colleagues, in particular Willi Zwerger, Cornish, S. L., N. R. Claussen, J. L. Roberts, E. A. Cornell,
and C. E. Wieman, 2000, Phys. Rev. Lett. 85, 1795.
for very helpful comments on the draft manuscript. This
Corwin, K. L., S. J. M. Kuppens, D. Cho, and C. E. Wieman,
work was supported by the National Science Foundation
1999, Phys. Rev. Lett. 83, 1311.
under Grants No. NSF-DMR-96-14133 and NSF-DMR-
Dalfovo, F., S. Giorgini, L. P. Pitaevskii, and S. Stringari, 1999,
99-86199. Rev. Mod. Phys. 71, 463.
Dalibard, J., 1999, in Bose-Einstein Condensation in Atomic
Gases, International School of Physics ‘‘Enrico Fermi’’
REFERENCES Course 140, edited by M. Inguscio, S. Stringari, and C. Wie-
man (IOS Press, Amsterdam).
Allum, D., P. V. E. McClintock, A. Phillips, and R. M. Bowley, De Marco, B., and D. S. Jin, 1999, Science 285, 1703.
1977, Philos. Trans. R. Soc. London, Ser. A 284, 179. Donnelly, R. J., 1967, Experimental Superfluidity (Chicago
Anderson, B. P., and M. A. Kasevich, 1998, Science 282, 1686. University Press, Chicago), Sec. 2.9.
Anderson, P. W., 1966, Rev. Mod. Phys. 38, 298. Drummond, P.D., and J. F. Corney, 1999, Technical Digest,
Andrews, M. R., M-O. Mewes, N. J. van Druten, D. S. Durfee, summaries of papers presented at the Quantum Electronics
D. M. Kurn, and W. Ketterle, 1996, Science 273, 84. and Laser Science Conference, postconference edition (cat.
Andrews, M. R., H-J. Miesner, D. M. Stamper-Kurn, J. no. 99CH37012) (Optical Society of America, Washington,
Stenger, and W. Ketterle, 1999, Phys. Rev. Lett. 82, 2422. D.C.), pp. 38,39.
Andrews, M. R., C. G. Townsend, H-J. Miesner, D. S. Durfee, Dunningham, J. A., and K. Burnett, 1999, Phys. Rev. Lett. 82,
D. M. Kurn, and W. Ketterle, 1997, Science 275, 637. 3729.
Avenel, O., and E. Varoquaux, 1988, Phys. Rev. Lett. 60, 416. Esry, B. D., and C. H. Greene, 1999, Phys. Rev. A 59, 1457.

Rev. Mod. Phys., Vol. 73, No. 2, April 2001


Anthony J. Leggett: Bose-Einstein condensation in the alkali gases 355

Fedichev, P. O., and G. V. Shlyapnikov, 1999, Phys. Rev. A 60, Kagan, Yu., and E. L. Surkov, 1997, Phys. Rev. A 55, R18.
R1779. Kagan, Yu., E. L. Surkov, and G. V. Shlyapnikov, 1996, Phys.
Fetter, A. L., 1972, Ann. Phys. (N.Y.) 70, 67. Rev. A 54, R1753.
Fetter, A. L., 1999, in Bose-Einstein Condensation in Atomic Kagan, Yu., and B. V. Svistunov, 1994, Zh. Eksp. Teor. Fiz.
Gases, International School of Physics ‘‘Enrico Fermi’’ 105, 353 [Sov. Phys. JETP 78, 187 (1994)].
Course 140, edited by M. Inguscio, S. Stringari, and C. E. Kagan, Yu., and B. V. Svistunov, 1998, Pisma Zh. Eksp. Teor.
Wieman (IOS Press, Amsterdam). Fiz. 67, 495 [JETP Lett. 67, 521 (1998)].
Fried, D. G., T. C. Killian, L. Willmann, D. Landhuis, S. C. Kagan, Yu., B. V. Svistunov, and G. V. Shlyapnikov, 1985,
Moss, D. Kleppner, and T. J. Greytak, 1998, Phys. Rev. Lett. Pis’ma Zh. Eksp. Teor. Fiz. 42, 169 [JETP Lett. 42, 209
81, 3807. (1985)].
Gajda, M., and K. Rzazewski, 1997, Phys. Rev. Lett. 78, 2686. Kennedy, T., E. H. Lieb, and S. Shastry, 1988, Phys. Rev. Lett.
Gardiner, C. W., 1997, Phys. Rev. A 56, 1414. 61, 2582.
Gardiner, C. W., and P. Zoller, 2000, Phys. Rev. A 61, 033601. Ketterle, W., D. S. Durfee, and D. M. Stamper-Kurn, 1999, in
Gavoret, J., and P. Nozières, 1964, Ann. Phys. (N.Y.) 28, 349. Bose-Einstein Condensation in Atomic Gases, International
Girardeau, M. D., and R. Arnowitt, 1959, Phys. Rev. 113, 755. School of Physics ‘‘Enrico Fermi’’ Course 140, edited by M.
Gribakin, A., and V. V. Flambaum, 1993, Phys. Rev. A 48, 546. Inguscio, S. Stringari, and C. Wieman (IOS Press, Amster-
Griffin, A., 2000, in Bose-Einstein Condensation: from Atomic dam).
Physics to Quantum Fluids, Proceedings of the 13th Physics Kukich, G., R. P. Henkel, and J. D. Reppy, 1968, Phys. Rev.
Summer School, edited by C. M. Savage and M. Das (World Lett. 21, 197.
Scientific, Singapore). Landau, L. D., and E. M. Lifshitz, 1959, Quantum Mechanics
Hall, D. S., M. R. Matthews, J. R. Ensher, C. E. Wieman, and (Pergamon, London).
E. A. Cornell, 1998, Phys. Rev. Lett. 81, 1539. Landau, L. D., and E. M. Lifshitz, 1969, Statistical Mechanics
Hall, D. S., M. R. Matthews, C. E. Wieman, and E. A. Cornell,
(Addison-Wesley, Reading, MA).
1998, Phys. Rev. Lett. 81, 1543.
Langer, J. S., and M. E. Fisher, 1967, Phys. Rev. Lett. 19, 560.
Hau, L., B. D. Busch, Chien Liu, Z. Dutton, M. M. Burns, and
Law, C. K., H. Pu, and N. P. Bigelow, 1998, Phys. Rev. Lett.
J. A. Golovchenko, 1998, Phys. Rev. A 58, R54.
81, 5157.
Hegstrom, R. A., 1998, Chem. Phys. Lett. 288, 248.
Hegstrom, R. A., and F. Sols, 1995, Found. Phys. 25, 681. Leggett, A. J., 1975, Rev. Mod. Phys. 47, 331.
Heinzen, D. J., R. Wynar, P. D. Drummond, and K. V. Leggett, A. J., 1995a, in Bose Einstein Condensation, edited by
Kheruntsyan, 2000, Phys. Rev. Lett. 84, 5029. A. Griffin, D. W. Snoke, and S. Stringari (Cambridge Univer-
Ho, T-L., 1982, Phys. Rev. Lett. 49, 1837. sity Press, New York), pp. 452–462.
Ho, T-L., 1998, Phys. Rev. Lett. 81, 742. Leggett, A. J., 1995b, in Twentieth Century Physics, edited by
Ho, T-L., and V. B. Shenoy, 1996, Phys. Rev. Lett. 77, 2595. L. M. Brown, A. Pais, and A. B. Pippard (Institute of Physics,
Ho, T-L., and S. K. Yip, 2000, Phys. Rev. Lett. 84, 4031. Bristol, England).
Hohenberg, P. C., 1967, Phys. Rev. 158, 383. Leggett, A. J., 1998, J. Stat. Phys. 93, 927.
Holland, M., J. Park, and R. Walser, 2000, preprint Leggett, A. J., 1999a, in Proceedings of the 16th International
cond-mat/0005062. Conference on Atomic Physics, Windsor, Ontario, Canada,
Huang, K., 1987, Statistical Mechanics, 2nd edition (Wiley, Aug. 1998, edited by W. E. Baylis and G. F. Drake, AIP
New York). Conf. Proc. No. 477 (AIP, Woodbury, New York), pp. 154–
Huang, W-J., and S-C. Gou, 1999, Phys. Rev. A 59, 4608. 169.
Huse, D. A., and E. Siggia, 1982, J. Low Temp. Phys. 46, 137. Leggett, A. J., 1999b, Lecture notes, University of Illinois, un-
Inguscio, M., S. Stringari, and C. E. Wieman, 1999, Eds., Bose- published.
Einstein Condensation in Atomic Gases, International School Leggett, A. J., 2000a, in Bose-Einstein Condensation: from
of Physics ‘‘Enrico Fermi’’ Course 140 (IOS Press, Amster- Atomic Physics to Quantum Fluids, Proceedings of the 13th
dam).
Physics Summer School, edited by C. M. Savage and M. Das
Inouye, S., A. P. Chikkatur, D. M. Stamper-Kurn, J. Stenger,
(World Scientific, Singapore).
D. E. Pritchard, and W. Ketterle, 1999, Science 285, 571.
Leggett, A. J., 2000b, in preparation for submission to J. Math.
Javanainen, J., 1996, Phys. Rev. A 54, R4629.
Phys.
Javanainen, J., and M. Yu. Ivanov, 1999, Phys. Rev. A 60,
2351. Leggett, A. J., and F. Sols, 1991, Found. Phys. 21, 353.
Javanainen, J., and M. Wilkens, 1997, Phys. Rev. Lett. 78, Leggett, A. J., and F. Sols, 1998, Phys. Rev. Lett. 81, 1344.
4675. Lieb, E. H., 1999, Physica A 263, 491.
Javanainen, J., and S-M. Yoo, 1996, Phys. Rev. Lett. 76, 161. Lopez-Arias, T., and A. Smerzi, 1998, Phys. Rev. A 58, 526.
Josephson, B. D., 1962, Phys. Lett. 1, 251. Loss, D., P. Goldbart, and A. V. Balatsky, 1990, Phys. Rev.
Julienne, P. S., 1996, J. Res. Natl. Inst. Stand. Technol. 101, Lett. 65, 1655.
487. Madison, K. W., F. Chevy, W. Wohlleben, and J. Dalibard,
Julienne, P. S., F. H. Mies, E. Tiesinga, and C. J. Williams, 2000, Phys. Rev. Lett. 84, 806.
1997, Phys. Rev. Lett. 78, 1880. Mandel, L., and E. Wolf, 1995, Optical Coherence and Quan-
Kagan, Yu., 1995, in Bose Einstein Condensation, edited by A. tum Optics (Cambridge University Press, New York).
Griffin, D. W. Snoke, and S. Stringari (Cambridge University Marinescu, M., H. R. Sadeghpour, and A. Dalgarno, 1994,
Press, New York), pp. 202–225. Phys. Rev. A 49, 982.
Kagan, Yu., A. E. Muryshev, and G. V. Shlapnikov, 1998, Marino, I., S. Raghavan, S. Fantoni, S. R. Shenoy, and A.
Phys. Rev. Lett. 81, 933. Smerzi, 1999, Phys. Rev. A 60, 487.

Rev. Mod. Phys., Vol. 73, No. 2, April 2001


356 Anthony J. Leggett: Bose-Einstein condensation in the alkali gases

Matthews, M. R., B. S. Anderson, P. C. Haljan, D. S. Hall, C. Smerzi, A., S. Fantoni, S. Giovanazzi, and S. R. Shenoy, 1997,
E. Wieman, and E. A. Cornell, 1999, Phys. Rev. Lett. 83, Phys. Rev. Lett. 79, 4950.
2498. Sols, F., 1999, in Bose-Einstein Condensation in Atomic Gases,
Matthews, M. R., D. S. Hall, D. S. Jin, J. R. Ensher, C. E. International School of Physics ‘‘Enrico Fermi’’ Course 140,
Wieman, E. A. Cornell, F. Dalfovo, C. Minniti, and S. Strin- edited by M. Inguscio, S. Stringari, and C. Wieman (IOS
gari, 1998, Phys. Rev. Lett. 81, 243. Press, Amsterdam).
Meier, F., and W. Zwerger, 2000, preprint. Stamper-Kurn, D. M., H-J. Miesner, A. P. Chikkatur, S. In-
Mewes, M-O., M. R. Andrews, D. M. Kurn, D. S. Durfee, C. ouye, J. Stenger, and W. Ketterle, 1999, Phys. Rev. Lett. 83,
G. Townsend, and W. Ketterle, 1997, Phys. Rev. Lett. 78, 582. 661.
Mies, F. H., C. J. Williams, P. S. Julienne, and M. Krauss, 1996, Steel, M. J., M. K. Olsen, L. I. Plimak, P. D. Drummond, S. M.
Phys. Rev. Lett. 78, 1850.
Tan, M. J. Collett, and D. F. Walls, 1998, Phys. Rev. A 58,
Miesner, H-J., D. M. Stamper-Kurn, J. Stenger, S. Inouye, A.
4824.
P. Chikkatur, and W. Ketterle, 1999, Phys. Rev. Lett. 82,
Stenger, J., S. Inouye, M. R. Andrews, H-J. Miesner, D. M.
2228.
Stamper-Kurn, and W. Ketterle, 1999, Phys. Rev. Lett. 82,
Milburn, G. J., J. Corney, E. M. Wright, and D. F. Walls, 1997,
Phys. Rev. A 55, 4318. 2422.
Mueller, E. J., P. M. Goldbart, and Y. Lyanda-Geller, 1998, Stenger, J., S. Inouye, D. M. Stamper-Kurn, H-J. Miesner, A.
Phys. Rev. A 57, R1505. P. Chikkatur, and W. Ketterle, 1998a, Nature (London) 396,
Naraschewski, M., H. Wallis, A. Schenzle, J. I. Cirac, and P. 345.
Zoller, 1996, Phys. Rev. A 54, 2185. Stenger, J., D. M. Stamper-Kurn, M. R. Andrews, A. P.
Nozières, P., 1995, in Bose Einstein Condensation, edited by A. Chikkatur, S. Inouye, H-J. Miesner, and W. Ketterle, 1998b,
Griffin, D. W. Snoke, and S. Stringari (Cambridge University J. Low Temp. Phys. 113, 167.
Press, New York), pp. 15–30. Stoof, H. T. C., 1999, J. Low Temp. Phys. 114, 11.
Ohberg, P., and S. Stenholm, 1999, J. Phys. B 32, 1959. Stoof, H. T. C., M. Bijlsma, and M. Houbiers, 1996, J. Res.
Ohmi, T., and K. Machida, 1998, J. Phys. Soc. Jpn. 67, 1822. Natl. Inst. Stand. Technol. 101, 443.
Oktel, M. O., and L. Levitov, 1999, Phys. Rev. Lett. 83, 6. Tiesinga, E., C. J. Williams, P. S. Julienne, K. M. Jones, P. D.
Ostrovskaya, E. A., Yu. S. Kivshar, M. Lisak, B. Hall, F. Cat- Lett, and W. D. Phillips, 1996, J. Res. Natl. Inst. Stand. Tech-
tani, and D. Anderson, 2000, Phys. Rev. A 61, 031601. nol. 101, 505.
Parkin, A. S., and D. F. Walls, 1998, Phys. Rep. 303, 1. Vinen, W. F., 1969, in Superconductivity, Vol. I, edited by R.
Pegg, D. T., and S. M. Barnett, 1988, Europhys. Lett. 6, 483. D. Parks (Marcell Dekker, New York).
Penrose, O., and L. Onsager, 1956, Phys. Rev. 104, 576. Walser, R., J. Williams, J. Cooper, and M. Holland, 1999, Phys.
Pethick, C. J., and H. Smith, 2000, Bose-Einstein Condensa- Rev. A 59, 3878.
tion in Dilute Gases (Cambridge University Press, to be pub-
Weinberg, S., 1972, Gravitation and Cosmology (John Wiley
lished).
and Sons, New York).
Petrich, W., M. H. Anderson, J. R. Ensher, and E. A. Cornell,
Weiner, J., V. S. Bagnato, S. Zilio, and P. S. Julienne, 1999,
1995, Phys. Rev. Lett. 74, 3352.
Rev. Mod. Phys. 71, 1.
Phillips, W. D., 1998, Rev. Mod. Phys. 70, 721.
Pritchard, D. E., 1983, Phys. Rev. Lett. 51, 1336. Wheatley, J. C., 1975, Rev. Mod. Phys. 47, 415.
Raghavan, S., A. Smerzi, S. Fantoni, and S. R. Shenoy, 1999, Wilkin, N. K., and J. M. F. Gunn, 2000, Phys. Rev. Lett. 84, 6.
Phys. Rev. A 59, 620. Williams, J. E., and M. J. Holland, 1999, Nature (London) 401,
Rohrl, A., M. Naraschewski, A. Schenzle, and H. Wallis, 1999, 568.
Phys. Rev. Lett. 78, 4143. Woodgate, G. K., 1970, Elementary Atomic Structure
Ruostekoski, J., and D. F. Walls, 1998, Phys. Rev. A 58, R50. (McGraw-Hill, London/New York).
Sackett, C. A., H. T. C. Stoof, and R. G. Hulet, 1998, Phys. Wynar, R., R. S. Freeland, D. J. Han, C. Ryu, and D. J. Hei-
Rev. Lett. 80, 2031. nzen, 2000, Science 287, 1016.
Sinatra, A., and Y. Castin, 1998, Eur. Phys. J. D 4, 247. Yang, C. N., 1962, Rev. Mod. Phys. 34, 694.
Sinatra, A., P. O. Fedichev, Y. Castin, J. Dalibard, and G. V. Zapata, I., F. Sols, and A. J. Leggett, 1998, Phys. Rev. A 57,
Shlyapnikov, 1999, Phys. Rev. Lett. 82, 251. R28.

Rev. Mod. Phys., Vol. 73, No. 2, April 2001

You might also like