Nobel Lecture: Bose-Einstein condensation in a dilute gas...

40
Nobel Lecture: Bose-Einstein condensation in a dilute gas, the first 70 years and some recent experiments* E. A. Cornell and C. E. Wieman JILA, University of Colorado and National Institute of Standards and Technology, and Department of Physics, University of Colorado, Boulder, Colorado 80309-0440 (Published 19 August 2002) Bose-Einstein condensation, or BEC, has a long and rich history dating from the early 1920s. In this article we will trace briefly over this history and some of the developments in physics that made possible our successful pursuit of BEC in a gas. We will then discuss what was involved in this quest. In this discussion we will go beyond the usual technical description to try and address certain questions that we now hear frequently, but are not covered in our past research papers. These are questions along the lines of: How did you get the idea and decide to pursue it? Did you know it was going to work? How long did it take you and why? We will review some our favorites from among the experiments we have carried out with BEC. There will then be a brief encore on why we are optimistic that BEC can be created with nearly any species of magnetically trappable atom. Throughout this article we will try to explain what makes BEC in a dilute gas so interesting, unique, and experimentally challenging. 1 The notion of Bose statistics dates back to a 1924 pa- per in which Satyendranath Bose used a statistical argu- ment to derive the black-body photon spectrum (Bose, 1924). Unable to publish his work, he sent it to Albert Einstein, who translated it into German and got it pub- lished. Einstein then extended the idea of Bose’s count- ing statistics to the case of noninteracting atoms (Ein- stein, 1924, 1925). The result was Bose-Einstein statistics. Einstein immediately noticed a peculiar fea- ture of the distribution of the atoms over the quantized energy levels predicted by these statistics. At very low but finite temperature a large fraction of the atoms would go into the lowest energy quantum state. In his words, ‘‘A separation is effected; one part condenses, the rest remains a saturated ideal gas’’ 2 (Einstein, 1925). This phenomenon we now know as Bose-Einstein con- densation. The condition for this to happen is that the phase-space density must be greater than approximately unity, in natural units. Another way to express this is that the de Broglie wavelength, l dB , of each atom must be large enough to overlap with its neighbor, or more precisely, n l dB 3 .2.61. This prediction was not taken terribly seriously, even by Einstein himself, until Fritz London (1938) and Las- zlo Tisza (1938) resurrected the idea in the mid 1930s as a possible mechanism underlying superfluidity in liquid helium 4. Their work was the first to bring out the idea of BEC displaying quantum behavior on a macroscopic size scale, the primary reason for much of its current attraction. Although it was a source of debate for de- cades, it is now recognized that the remarkable proper- ties of superconductivity and superfluidity in both he- lium 3 and helium 4 are related to BEC, even though these systems are very different from the ideal gas con- sidered by Einstein. The appeal of the exotic behavior of superconductiv- ity and of superfluidity, along with that of laser light, the third common system in which macroscopic quantum behavior is evident, provided much of our motivation in 1990 when we decided to pursue BEC in a gas. These three systems all have fascinating counterintuitive be- havior arising from macroscopic occupation of a single quantum state. Any physicist would consider these phe- nomena among the most remarkable topics in physics. In 1990 we were confident that the addition of a new member to the family would constitute a major contri- bution to physics. (Only after we succeeded did we real- ize that the discovery of each of the original Macro- scopic Three had been recognized with a Nobel Prize, and we are grateful that this trend has continued!) Al- though BEC shares the same underlying mechanism with these other systems, it seemed to us that the prop- erties of BEC in a gas would be quite distinct. It is far more dilute and weakly interacting than liquid-helium superfluids, for example, but far more strongly interact- ing than the noninteracting light in a laser beam. Per- haps BEC’s most distinctive feature (and this was not *The 2001 Nobel Prize in Physics was shared by E. A. Cor- nell, Wolfgang Ketterle, and C. E. Wieman. 1 This article is our ‘‘Nobel Lecture’’ and as such takes a rela- tively personal approach to the story of the development of experimental Bose-Einstein condensation. For a somewhat more scholarly treatment of the history, the interested reader is referred to E. A. Cornell, J. R. Ensher, and C. E. Wieman, ‘‘Experiments in dilute atomic Bose-Einstein condensation in Bose-Einstein Condensation in Atomic Gases,’’ Proceedings of the International School of Physics ‘‘Enrico Fermi’’ Course CXL, edited by M. Inguscio, S. Stringari, and C. E. Wieman (Italian Physical Society, 1999), pp. 15 66, which is also avail- able as cond-mat/9903109. For a reasonably complete technical review of the three years of explosive progress that immedi- ately followed the first observation of BEC, we recommend reading the above article in combination with the correspond- ing review from Ketterle, cond-mat/9904034. 2 English translation of Einstein’s quotes and the historical in- terpretation are from Pais (1982), Subtle is the Lord .... REVIEWS OF MODERN PHYSICS, VOLUME 74, JULY 2002 0034-6861/2002/74(3)/875(19)/$35.00 © The Nobel Foundation 2001 875

Transcript of Nobel Lecture: Bose-Einstein condensation in a dilute gas...

Page 1: Nobel Lecture: Bose-Einstein condensation in a dilute gas ...duine102/summerschoolTP/BECNobel.pdf · experimental Bose-Einstein condensation. For a somewhat more scholarly treatment

REVIEWS OF MODERN PHYSICS, VOLUME 74, JULY 2002

Nobel Lecture: Bose-Einstein condensation in a dilute gas,the first 70 years and some recent experiments*

E. A. Cornell and C. E. Wieman

JILA, University of Colorado and National Institute of Standards and Technology,and Department of Physics, University of Colorado, Boulder, Colorado 80309-0440

(Published 19 August 2002)

Bose-Einstein condensation, or BEC, has a long and rich history dating from the early 1920s. In thisarticle we will trace briefly over this history and some of the developments in physics that madepossible our successful pursuit of BEC in a gas. We will then discuss what was involved in this quest.In this discussion we will go beyond the usual technical description to try and address certain questionsthat we now hear frequently, but are not covered in our past research papers. These are questionsalong the lines of: How did you get the idea and decide to pursue it? Did you know it was going towork? How long did it take you and why? We will review some our favorites from among theexperiments we have carried out with BEC. There will then be a brief encore on why we are optimisticthat BEC can be created with nearly any species of magnetically trappable atom. Throughout thisarticle we will try to explain what makes BEC in a dilute gas so interesting, unique, and experimentallychallenging.1

The notion of Bose statistics dates back to a 1924 pa-per in which Satyendranath Bose used a statistical argu-ment to derive the black-body photon spectrum (Bose,1924). Unable to publish his work, he sent it to AlbertEinstein, who translated it into German and got it pub-lished. Einstein then extended the idea of Bose’s count-ing statistics to the case of noninteracting atoms (Ein-stein, 1924, 1925). The result was Bose-Einsteinstatistics. Einstein immediately noticed a peculiar fea-ture of the distribution of the atoms over the quantizedenergy levels predicted by these statistics. At very lowbut finite temperature a large fraction of the atomswould go into the lowest energy quantum state. In hiswords, ‘‘A separation is effected; one part condenses,the rest remains a saturated ideal gas’’ 2 (Einstein, 1925).This phenomenon we now know as Bose-Einstein con-densation. The condition for this to happen is that thephase-space density must be greater than approximatelyunity, in natural units. Another way to express this is

*The 2001 Nobel Prize in Physics was shared by E. A. Cor-nell, Wolfgang Ketterle, and C. E. Wieman.

1This article is our ‘‘Nobel Lecture’’ and as such takes a rela-tively personal approach to the story of the development ofexperimental Bose-Einstein condensation. For a somewhatmore scholarly treatment of the history, the interested readeris referred to E. A. Cornell, J. R. Ensher, and C. E. Wieman,‘‘Experiments in dilute atomic Bose-Einstein condensation inBose-Einstein Condensation in Atomic Gases,’’ Proceedings ofthe International School of Physics ‘‘Enrico Fermi’’ CourseCXL, edited by M. Inguscio, S. Stringari, and C. E. Wieman(Italian Physical Society, 1999), pp. 15–66, which is also avail-able as cond-mat/9903109. For a reasonably complete technicalreview of the three years of explosive progress that immedi-ately followed the first observation of BEC, we recommendreading the above article in combination with the correspond-ing review from Ketterle, cond-mat/9904034.

2English translation of Einstein’s quotes and the historical in-terpretation are from Pais (1982), Subtle is the Lord . . . .

0034-6861/2002/74(3)/875(19)/$35.00 875

that the de Broglie wavelength, ldB , of each atom mustbe large enough to overlap with its neighbor, or moreprecisely, nldB

3 .2.61.This prediction was not taken terribly seriously, even

by Einstein himself, until Fritz London (1938) and Las-zlo Tisza (1938) resurrected the idea in the mid 1930s asa possible mechanism underlying superfluidity in liquidhelium 4. Their work was the first to bring out the ideaof BEC displaying quantum behavior on a macroscopicsize scale, the primary reason for much of its currentattraction. Although it was a source of debate for de-cades, it is now recognized that the remarkable proper-ties of superconductivity and superfluidity in both he-lium 3 and helium 4 are related to BEC, even thoughthese systems are very different from the ideal gas con-sidered by Einstein.

The appeal of the exotic behavior of superconductiv-ity and of superfluidity, along with that of laser light, thethird common system in which macroscopic quantumbehavior is evident, provided much of our motivation in1990 when we decided to pursue BEC in a gas. Thesethree systems all have fascinating counterintuitive be-havior arising from macroscopic occupation of a singlequantum state. Any physicist would consider these phe-nomena among the most remarkable topics in physics.In 1990 we were confident that the addition of a newmember to the family would constitute a major contri-bution to physics. (Only after we succeeded did we real-ize that the discovery of each of the original Macro-scopic Three had been recognized with a Nobel Prize,and we are grateful that this trend has continued!) Al-though BEC shares the same underlying mechanismwith these other systems, it seemed to us that the prop-erties of BEC in a gas would be quite distinct. It is farmore dilute and weakly interacting than liquid-heliumsuperfluids, for example, but far more strongly interact-ing than the noninteracting light in a laser beam. Per-haps BEC’s most distinctive feature (and this was not

© The Nobel Foundation 2001

Page 2: Nobel Lecture: Bose-Einstein condensation in a dilute gas ...duine102/summerschoolTP/BECNobel.pdf · experimental Bose-Einstein condensation. For a somewhat more scholarly treatment

876 E. A. Cornell and C. E. Wieman: BEC in a dilute gas

something we sufficiently appreciated, in 1990) is theease with which its quantum wave function may be di-rectly observed and manipulated. While neither of uswas to read C. E. Hecht’s prescient 1959 paper (Hecht,1959) until well after we had observed BEC, we surelywould have taken his concluding paragraph as ourmarching orders:

The suppositions of this note rest on the possibility ofsecuring, say by atomic beam techniques, substantialquantities of electron-spin-oriented H, T and D atoms.Although the experimental difficulties would be greatand the relaxation behavior of such spin-oriented at-oms essentially unknown, the possibility of opening arich new field for the study of superfluid properties inboth liquid and gaseous states would seem to demandthe expenditure of maximum experimental effort.3

In any case, by 1990 we were awash in motivation. Butthis motivation would not have carried us far, had wenot been able to take advantage of some key recent ad-vances in science and technology, in particular, theprogress in laser cooling and trapping and the extensiveachievements of the spin-polarized-hydrogen commu-nity.

However, before launching into that story, it is per-haps worthwhile to reflect on just how exotic a system ofindistinguishable particles truly is, and why BEC in a gasis such a daunting experimental challenge. It is easy atfirst to accept that two atoms can be so similar one tothe other as to allow no possibility of telling them apart.However, confronting the physical implications of theconcept of indistinguishable bosons can be troubling.For example, if there are ten bosonic particles to be ar-ranged in two microstates of a system, the statisticalweight of the configuration with ten particles in onestate and zero in the other is exactly the same as theweight of the configuration with five particles in onestate, five in the other. This 1:1 ratio of statistical weightsis very counterintuitive and rather disquieting. The cor-responding ratio for distinguishable objects, such assocks in drawers, that we observe every day is 1:252,profoundly different from 1:1. In the second of Ein-stein’s two papers (Einstein, 1925; Pais, 1982) on Bose-Einstein statistics, Einstein comments that ‘‘The . . .molecules are not treated as statistically indepen-dent . . . , and the differences between distinguishableand indistinguishable state counting . . . express indi-rectly a certain hypothesis on a mutual influence of themolecules which for the time being is of a quite myste-rious nature. This mutual influence is no less mysterioustoday, even though we can readily observe the variety ofexotic behavior it causes such as the well-known en-hanced probability for scattering into occupied statesand, of course, Bose-Einstein condensation.’’

Not only does the Bose-Einstein phase transition of-fend our sensibilities as to how particles ought best todistribute themselves, it also runs counter to an unspo-

3Emphasis ours.

Rev. Mod. Phys., Vol. 74, No. 3, July 2002

ken assumption that a phase transition somehow in-volves thermodynamic stability. In fact, the regions im-mediately above and immediately below the transitionin dilute-gas experiments are both deep in the thermo-dynamically forbidden regime. This point is best madeby considering a qualitative phase diagram (Fig. 1),which shows the general features common to any atomicsystem. At low density and high temperature, there is avapor phase. At high density there are various con-densed phases. But the intermediate densities are ther-modynamically forbidden, except at very high tempera-tures. The Bose-condensed region of the n-T plane isutterly forbidden, except at such high densities that(with one exception) all known atoms or moleculeswould form a crystalline lattice, which would rule outBose condensation. The single exception, helium, re-mains a liquid below the BEC transition. However,reaching BEC under dilute conditions (say, at densities10 or 100 times lower than conventional liquid helium) isas thermodynamically forbidden to helium as it is to anyother atom.

Of course, forbidden is not the same as impossible;indeed, to paraphrase an old Joseph Heller joke, if itwere really impossible, they wouldn’t have bothered toforbid it. It comes down in the end to differing timescales for different sorts of equilibrium. A gas of atomscan come into kinetic equilibrium via two-body colli-sions, whereas it requires three-body collisions toachieve chemical equilibrium (i.e., to form moleculesand thence solids). At sufficiently low densities, the two-body rate will dominate the three-body rate, and a gaswill reach kinetic equilibrium, perhaps in a metastableBose-Einstein condensate, long before the gas finds itsway to the ultimately stable solid-state condition. Theneed to maintain metastability usually dictates a morestringent upper limit on density than does the desire tocreate a dilute system. Densities around 1020 cm23, forinstance, would be a hundred times more dilute than acondensed-matter helium superfluid. But creating such agas is quite impractical even at an additional factor-of-1000 lower density, say 1017 cm23, when metastabilitytimes would be on the order of a few microseconds;

FIG. 1. Generic phase diagram common to all atoms: dottedline, the boundary between non-BEC and BEC; solid line, theboundary between allowed and forbidden regions of thetemperature-density space. Note that at low and intermediatedensities, BEC exists only in the thermodynamically forbiddenregime.

Page 3: Nobel Lecture: Bose-Einstein condensation in a dilute gas ...duine102/summerschoolTP/BECNobel.pdf · experimental Bose-Einstein condensation. For a somewhat more scholarly treatment

877E. A. Cornell and C. E. Wieman: BEC in a dilute gas

more realistic are densities on the order of 1014 cm23.The low densities mandated by the need to maintainlong-lived metastability in turn make necessary theachievement of still lower temperatures if one is to reachBEC.

Thus the great experimental hurdle that must be over-come to create BEC in a dilute gas is to form and keepa sample that is so deeply forbidden. Since our subse-quent discussion will focus only on BEC in dilute gases,we shall refer to this simply as BEC in the sections be-low and avoid endlessly repeating ‘‘in a dilute gas.’’

Efforts to make a dilute BEC in an atomic gas weresparked by Stwalley and Nosanow (1976). They arguedthat spin-polarized hydrogen had no bound states andhence would remain a gas down to zero temperature,and so it would be a good candidate for BEC. Thisstimulated a number of experimental groups (Silveraand Walraven, 1980; Hardy et al., 1982; Hess et al., 1983;Johnson et al., 1984) in the late 1970s and early 1980s tobegin pursuing this idea using traditional cryogenics tocool a sample of polarized hydrogen. Spin-polarized hy-drogen was first stabilized by Silvera and Walraven in1980, and by the mid 1980s spin-polarized hydrogen hadbeen brought within a factor of 50 of condensing (Hesset al., 1983). These experiments were performed in a di-lution refrigerator, in a cell in which the walls werecoated with superfluid liquid helium as a nonstick coat-ing for the hydrogen. The hydrogen gas was compressedusing a piston-in-cylinder arrangement (Bell et al., 1986)or inside a helium bubble (Sprik et al., 1985). These at-tempts failed, however, because when the cell was madevery cold the hydrogen stuck to the helium surface andrecombined. When one tried to avoid that problem bywarming the cell sufficiently to prevent sticking, the den-sity required to reach BEC was correspondingly in-creased, which led to another problem. The requisitedensities could not be reached because the rate of three-body recombination of atoms into hydrogen moleculesgoes up rapidly with density and the resulting loss ofatoms limited the density (Hess, 1986).

Stymied by these problems, Harold Hess (Hess, 1986)from the MIT hydrogen group realized that magnetictrapping of atoms (Migdall et al., 1985; Bagnato et al.,1987) would be an improvement over a cell. Atoms in amagnetic trap have no contact with a physical surfaceand thus the surface-recombination problem could becircumvented. Moreover, thermally isolated atoms in amagnetic trap would allow cooling by evaporation to farlower temperatures than previously obtained. In a re-markable paper, Hess (1986) laid out most of the impor-tant concepts of evaporative cooling of trapped atomsfor the attainment of BEC. Let the highest-energy atomsescape from the trap, and the mean energy, and thus thetemperature, of the remaining atoms will decrease. For adilute gas in an inhomogeneous potential, decreasing thetemperature will decrease the occupied volume. Onecan thus actually increase the density of the remainingatoms by removing atoms from the sample. The all im-portant (for BEC) phase-space density is dramaticallyincreased as this happens because density is rising while

Rev. Mod. Phys., Vol. 74, No. 3, July 2002

temperature is decreasing. The Cornell University hy-drogen group also considered evaporative cooling(Lovelace et al., 1985). By 1988 the MIT group had dem-onstrated these virtues of evaporative cooling of mag-netically trapped spin-polarized hydrogen. By 1991 theyobtained, at a temperature of 100 °K, a density that wasonly a factor of 5 below BEC (Doyle, 1991a). Furtherprogress was limited by dipolar relaxation, but perhapsmore fundamentally by loss of signal-to-noise, and thedifficulty of measuring the characteristics of the coldestand smallest clouds (Doyle, 1991b). Evaporative workwas also performed by the Amsterdam group (Luitenet al., 1993).

At roughly the same time, but independent from thehydrogen work, an entirely different type of cold-atomphysics and technology was being developed. Lasercooling and trapping has been reviewed elsewhere (Ari-mondo et al., 1991; Chu, 1998; Cohen-Tannoudji, 1998;Phillips, 1998), but here we mention some of the high-lights most relevant to our work. The idea that laserlight could be used to cool atoms was suggested in earlypapers by Wineland and Dehmelt (1975), by Hansch andSchawlow (1975), and by Letokhov’s group (Letokhov,1968). Early optical force experiments were performedby Ashkin (Bjorkholm et al., 1978). Trapped ions werelaser-cooled at the University of Washington (Neu-hauser et al., 1978) and at the National Bureau of Stan-dards (now NIST) in Boulder (Wineland et al., 1978).Atomic beams were deflected and slowed in the early1980s (Andreev et al., 1981; Ertmer et al., 1985; Prodanet al., 1985). Optical molasses, where the atoms arecooled to very low temperatures by six perpendicularintersecting laser beams, was first studied at Bell Labs(Chu et al., 1985). Measured temperatures in the earlymolasses experiments were consistent with the so-calledDoppler limit, which amounts to a few hundred mi-crokelvin in most alkalis. Light was first used to hold(trap) atoms using the dipole force exerted by a stronglyfocused laser beam (Chu et al., 1986). In 1987 and 1988there were two major advances that became central fea-tures of the method of creating BEC. First, a practicalspontaneous-force trap, the magneto-optical trap(MOT) was demonstrated (Raab et al., 1987); and sec-ond, it was observed that under certain conditions, thetemperatures in optical molasses are in fact much colderthan the Doppler limit (Lett et al., 1988; Chu et al., 1989;Dalibard et al., 1989). The MOT had the essential ele-ments needed for a widely useful optical trap: it requiredrelatively modest amounts of laser power, it was muchdeeper than dipole traps, and it could capture and holdrelatively large numbers of atoms. These were headytimes in the laser-cooling business. With experimentyielding temperatures mysteriously far below whattheory would predict, it was clear that we all lived underthe authority of a munificent God.

During the mid 1980s one of us (Carl) began investi-gating how useful the technology of laser trapping andcooling could become for general use in atomic physics.Originally this took the form of just making it cheaperand simpler by replacing the expensive dye lasers with

Page 4: Nobel Lecture: Bose-Einstein condensation in a dilute gas ...duine102/summerschoolTP/BECNobel.pdf · experimental Bose-Einstein condensation. For a somewhat more scholarly treatment

878 E. A. Cornell and C. E. Wieman: BEC in a dilute gas

vastly cheaper semiconductor lasers, and then searchingfor ways to allow atom trapping with these low-cost butalso low-power lasers (Pritchard et al., 1986; Watts andWieman, 1986). With the demonstration of the MOTand sub-Doppler molasses Carl’s group began eagerlystudying what physics was limiting the coldness anddenseness of these trapped atoms, with the hope of ex-tending the limits further. They discovered that severalatomic processes were responsible for these limits.Light-assisted collisions were found to be the major lossprocess from the MOT as the density increased (Seskoet al., 1989). However, even before that became a seri-ous problem, the light pressure from reradiated photonslimited the density (Walker et al., 1990; Sesko et al.,1991). At about the same time, the sub-Doppler tem-peratures of molasses found by Phillips, Chu, andCohen-Tannoudji were shown to be due to a combina-tion of light-shifts and optical pumping that becameknown as Sysiphus cooling (Dalibard and Cohen-Tannoudji, 1989). Random momentum fluctuations fromthe scattered photons limit the ultimate temperature toabout a factor of 10 above the recoil limit. In largersamples, the minimum temperature was higher yet, be-cause of the multiple scattering of the photons. Whilecarrying out studies on the density limits of MOT’sCarl’s group also continued the effort in technology de-velopment. This resulted in the creation of a usefulMOT in a simple glass vapor cell (Monroe et al., 1990),thereby eliminating the substantial vacuum chamber re-quired for the slowed atomic beam loading that had pre-viously been used.

Seeking to take advantage of the large gains in phase-space density provided by the MOT while avoiding thelimitations imposed by the undesirable effects of pho-tons, Carl and his student Chris Monroe decided to tryloading the cold MOT atoms into a magnetic trap (Mon-roe et al., 1990; see Fig. 2). This worked remarkablywell. Because further cooling could be carried out as theatoms were transferred between optical and magnetictrap it was possible to get very cold samples, the coldestthat had been produced at that time. More importantly,these were not optical molasses samples that werequickly disappearing but rather magnetically trappedsamples that could be held and studied for extended pe-riods. These samples were about a hundred times colderthan any previous trapped atom samples, with a corre-spondingly increased phase-space density. This was asatisfying achievement, but as much as the result itself, itwas the relative simplicity of the apparatus required thatinspired us (including now Eric Cornell, who joined theproject as a postdoc in 1990) to see just how far we couldpush this marriage of laser cooling and trapping andmagnetic trapping.

Previous laser traps involved expensive massive lasersystems and large vacuum chambers for atomic beamprecooling. Previous magnetic traps for atoms were usu-ally (Bagnato et al., 1987; Doyle, 1991) extremely com-plex and bulky (often with superconducting coils) be-cause of the need to have sufficiently large depths andstrong confinement. Laser traps and magnetic traps were

Rev. Mod. Phys., Vol. 74, No. 3, July 2002

both somewhat heroic experiments individually, to beundertaken only by a select handful of well-equippedAMO laboratories. The prospect of trying to get bothtraps working, and working well, in the same room andon the same day, was daunting. However, in the firstJILA magnetic trap experiment our laser sources weresimple diode lasers, the vacuum system was a small glassvapor cell, and the magnetic trap was just a few turns ofwire wrapped around it. This magnetic field was ad-equate because of the low temperatures of the laser-cooled and trapped samples. Being able to produce suchcold and trapped samples in this manner encouragedone to fantasize wildly about possible things to do withsuch an atom sample. Inspired by the spin-polarized hy-drogen work, our fantasizing quickly turned to the ideaof evaporative cooling further to reach BEC. It wouldrequire us to increase the phase-space density by 5 or-ders of magnitude, but since we had just gained about 15orders of magnitude almost for free with the vapor cellMOT, this did not seem so daunting.

The JILA vapor-cell MOT (Fig. 3), with its superim-posed ion pump trap, introduced a number of ideas thatare now in common use in the hybrid trapping business(Monroe et al., 1990; Monroe, 1992): (i) Vapor-cell(rather than beam) loading, (ii) fused-glass rather thanwelded-steel architecture, (iii) extensive use of diode la-sers, (iv) magnetic coils located outside the chamber, (v)overall chamber volume measured in cubic centimetersrather than liters, (vi) temperatures measured by imag-ing an expanded cloud, (vii) magnetic-field curvaturescalibrated in situ by observing the frequency of dipoleand quadrupole (sloshing and pulsing) cloud motion,(viii) the basic approach of a MOT and a magnetic trapwhich are spatially superimposed (indeed, which oftenshare some magnetic coils) but temporally sequential,and (ix) optional use of additional molasses and opticalpumping sequences inserted in time between the MOTand magnetic trapping stages. It is instructive to notehow a modern, Ioffe-Pritchard-based BEC device (Fig.4) resembles its ancestor (Fig. 3).

As we began to think about applying the technique ofevaporative cooling with hydrogen to our very cold al-kali atoms we looked carefully at the hydrogen workand its lessons. When viewed from our 1990 perspectivethe previous decade of work on polarized hydrogen pro-vided a number of important insights. It was clear thatthe unique absence of any bound states for spin-polarized hydrogen was actually not an important issue(other than its being the catalyst for starting the entirefield, of course!). Bound states or not, a very coldsample of spin-polarized hydrogen, like every other gas,has a lower-energy state to which it can go, and its sur-vival depends on the preservation of metastability. Forhydrogen the lower-energy state is a solid, althoughfrom an experimental point of view the rate-limitingprocess is the formation of diatomic molecules (with ap-propriately reoriented spins). Given that all atomicgases are only metastable at the BEC transition point,the real experimental issue becomes: How well can one

Page 5: Nobel Lecture: Bose-Einstein condensation in a dilute gas ...duine102/summerschoolTP/BECNobel.pdf · experimental Bose-Einstein condensation. For a somewhat more scholarly treatment

879E. A. Cornell and C. E. Wieman: BEC in a dilute gas

preserve the requisite metastability while still coolingsufficiently far to reach BEC?

The realization that metastability was the key experi-mental challenge one should focus on was probably atleast as important to the attainment of BEC as any ofthe experimental techniques we subsequently developedto actually achieve it. The work on hydrogen providedan essential guide for evaluating and tackling this chal-lenge. It provided us with a potential cooling technique(evaporative cooling of magnetically trapped atoms) andmapped out many of the processes by which a magneti-cally trapped atom can be lost from its metastable state.

FIG. 2. Chris Monroe examines an early hybrid MOT-magnetic trap apparatus [Color].

FIG. 3. The glass vapor cell and magnetic coils used in earlyJILA efforts to hybridize laser cooling and magnetic trapping(see Monroe et al., 1990). The glass tubing is 2.5 cm in diam-eter. The Ioffe current bars have been omitted for clarity.

Rev. Mod. Phys., Vol. 74, No. 3, July 2002

The hydrogen work made it clear that it was all an issueof good versus bad collisions. The good collisions areelastic collisions that rethermalize the atoms duringevaporation. The more collisions there are, the morequickly and efficiently one can cool. The bad collisionsare the inelastic collisions that quench the metastability.Hydrogen had already shown that three-body recombi-nation collisions and dipole spin-flip collisions were themajor inelastic culprits. The fact that hydrogen research-ers were fairly close to reaching BEC was also a strongencouragement. It meant that the goal was not ridicu-lously distant and that one only had to do a little betterin the proportion of good to bad collisions to succeed.

The more we thought about this, the more we beganto suspect that our heavy alkali atoms would likely havemore favorable collision properties than hydrogen at-oms and thus have a good chance of success. Althoughknowledge of the relevant collision cross sections wastotally nonexistent at that time, we were able to come upwith arguments for how the cross sections might scalerelative to hydrogen. These are discussed in more detailbelow in the section discussing why collisional concernsmake it likely that BEC can be created in a large num-ber of different species. Here we will just give a briefsummary consistent with our views circa 1990. The di-pole spin-flip collisions that limited hydrogen involvespin-spin interactions and thus could be expected to besimilar for the alkalis and for hydrogen because themagnetic moments are all about the same. The goodcollisions needed for evaporative cooling, however,should be much larger for heavy alkalis with their fatfluffy electron clouds than for hydrogen. The other vil-lain of the hydrogen effort, three-body recombination,was a total mystery, but because it goes as density cubedwhile the good elastic collisions go as density squared, itseemed as if we should always be able to find a suffi-ciently low-density and low-temperature regime to avoidit (see Monroe, 1992).

As a minor historical note, we might point out thatduring these considerations we happily ignored the factthat the temperatures required to achieve BEC in a

FIG. 4. Modern MOT and magnetic trap apparatus, used byCornish et al., 2000 [Color].

Page 6: Nobel Lecture: Bose-Einstein condensation in a dilute gas ...duine102/summerschoolTP/BECNobel.pdf · experimental Bose-Einstein condensation. For a somewhat more scholarly treatment

880 E. A. Cornell and C. E. Wieman: BEC in a dilute gas

heavy alkali gas are far colder than those needed for thesame density of hydrogen. The critical temperature forideal-gas BEC is inversely proportional to the mass. Itwas clear that we would need to cool to well under amicrokelvin, and a large three-body recombination ratewould have required us to go to possibly far lower tem-peratures. To someone coming from a traditional cryo-genics background this would (and probably did) seemlike sheer folly. The hydrogen work had been pushinghard for some years at the state of the art in cryogenictechnology, and here we proposed to happily jump farbeyond that. Fortunately we were coming to this froman AMO background in a time when temperaturesachieved by laser cooling were dropping through thefloor. Optimism was in the air. In fact, we later discov-ered optimism can take one only so far: There were ac-tually considerable experimental difficulties, and furthercooling came at some considerable effort and a five-yeardelay. Nevertheless, it is remarkable that with evapora-tive cooling a magnetically trapped sample of atoms, sur-rounded on all sides by a 300-K glass cell, can be cooledto reach temperatures of only a few nanokelvin, andmoreover it looks quite feasible to reach even coldertemperatures.

General collisional considerations gave us some hopethat the evaporative cooling hybrid trap approach withalkali atoms would get us to BEC, or, if not, at leastreveal some interesting new physics that would preventit. Nonetheless, there were powerful arguments againstpursuing this. First, our 1990-era arguments in favor of itwere based on some very fuzzy intuition; there were nocollision data or theories to back it up and there werestrong voices in disagreement. Second, the hydrogen ex-periments seemed to be on the verge of reaching BEC,and in fact we thought it was likely that if BEC could beachieved they would succeed first. However, our beliefin the virtues of our technology really carried the day inconvincing us to proceed. With convenient lasers in thenear-IR, and with the good optical access of a room-temperature glass cell, detection sensitivity could ap-proach single-atom capability. We could take pictures ofonly a few thousand trapped atoms and immediatelyknow the energy and density distribution. If we wantedto modify our magnetic trap it only required a few hourswinding and installing a new coil of wires. This was adramatic contrast with the hydrogen experiments that,like all state-of-the-art cryogenics experiments, requiredan apparatus that was the better part of two stories, andthe time to modify it was measured in (large) fractionsof a year. Also, atomic hydrogen was much more diffi-cult to detect and so the diagnostics were far more lim-ited. This convinced us that although hydrogen wouldlikely succeed first, our hybrid trap approach with easilyobserved and manipulated alkali samples would be ableto carry out important science and so was well worthpursuing in its own right.

From the very beginning in 1990, our work on BECwas heavily involved with cold atomic collisions. Thiswas somewhat ironic since previously both of us hadactively avoided the large fraction of AMO work on the

Rev. Mod. Phys., Vol. 74, No. 3, July 2002

subject of atomic collisions. Atomic collisions at verycold temperatures is now a major branch of the disci-pline of AMO physics, but at the end of the 1980s therewere almost no experimental data, and what there wascame in fact from the spin-polarized hydrogen experi-ments (Gillaspy et al., 1989). There was theoretical workon hydrogen from Shlyapnikov and Kagan (Kagan et al.,1981, 1984), and from Silvera and Verhaar (Lagendijket al., 1986). An early paper by Pritchard (1986) includesestimates on low-temperature collisional properties foralkalis. His estimates were extrapolations from room-temperature results, but in retrospect, several were sur-prisingly accurate. As we began to work on evaporativecooling, much of our effort was devoted to determiningthe sizes of all the relevant good and bad collision crosssections. Our efforts were helped by the theoretical ef-forts of Boudewijn Verhaar, who was among the first totake our efforts seriously and attempt to calculate therates in question. Chris Greene also provided us withsome useful theoretical estimates.

Starting in 1990 we carried out a series of experimentsexploring various magnetic traps and measuring the rel-evant collision cross sections. As this work proceededwe developed a far better understanding of the condi-tions necessary for evaporative cooling and a muchclearer understanding of the relevant collisional issues(Monroe et al., 1993; Newbury et al., 1995). Our experi-mental concerns evolved accordingly. In the early ex-periments (Monroe et al., 1990, 1993; Cornell et al., 1991;Monroe, 1992) a number of issues came up that continueto confront all BEC experiments: the importance ofaligning the centers of the MOT and the magnetic trap,the density-reducing effects of mode-mismatch, the needto account carefully for the (previously ignored) force ofgravity, heating (and not merely loss) from backgroundgas collisions, the usefulness of being able to turn off themagnetic fields rapidly, the need to synchronize manychanges in laser status and magnetic fields together withimage acquisition, an appreciation for the many issuesthat can interfere with accurate determinations of den-sity and temperature by optical methods, either flores-cence or absorption imaging, and careful stabilization ofmagnetic fields. The mastery of these issues in theseearly days made it possible for us to proceed relativelyquickly to quantitative measurements with the BEConce we had it.

In 1992 we came to realize that dipolar relaxation inalkalis should in principle not be a limiting factor. Asexplained in the final section of this article, collisionalscaling with temperature and magnetic field is such that,except in pathological situations, the problem of goodand bad collisions in the evaporative cooling of alkalis isreduced to the ratio of the elastic collision rate to therate of loss due to imperfect vacuum; dipolar relaxationand three-body recombination can be finessed, particu-larly since our preliminary data showed they were notenormous. It was reassuring to move ahead on efforts toevaporate with the knowledge that, while we were es-sentially proceeding in the dark, there were not as manymonsters in the dark as we had originally imagined.

Page 7: Nobel Lecture: Bose-Einstein condensation in a dilute gas ...duine102/summerschoolTP/BECNobel.pdf · experimental Bose-Einstein condensation. For a somewhat more scholarly treatment

881E. A. Cornell and C. E. Wieman: BEC in a dilute gas

It rapidly became clear that the primary concernswould be having sufficient elastic collision rate in themagnetic trap and sufficiently low background pressureto have few background collisions that removed atomsfrom the trap. To accomplish this it was clear that weneeded higher densities in the magnetic trap than wewere getting from the MOT. Our first effort to increasethe density two years earlier was based on a multiple-loading scheme (Cornell et al., 1991). Multiple MOT-loads of atoms were launched in moving molasses, opti-cally pumped into an untrapped Zeeman level, focusedinto a magnetic trap, then optically repumped into atrapped level. The repumping represented the necessarydissipation, so that multiple loads of atoms could be in-serted in a continuously operating magnetic trap. Inpractice, each step of the process involved some losses,and the final result was disappointing. Later, however, asdiscussed below, we resurrected the idea of multipleloading from one MOT to another to good advantage(Gibble et al., 1995; Myatt et al., 1996). This is now atechnique currently in widespread practice.

In addition to building up the initial density we real-ized that the collision rate could be dramatically in-creased by, after loading into a magnetic trap, compress-ing the atoms by further increasing the curvature of theconfining magnetic fields. In a harmonic trap, the colli-sion rate after adiabatic compression scales as the finalconfining frequency squared (Monroe, 1992). Thismethod is discussed by Monroe (1992) and was imple-mented first in early ground-state collisional work (Mon-roe et al., 1993).

In fall of 1992, Eric’s postdoctoral appointment con-cluded, and, after a tour through the job market, he de-cided to take the equivalent of an assistant professorposition at JILA/NIST. He decided to use his startupmoney to build a new experimental apparatus thatwould be designed to put these ideas together to makesure evaporation worked as we expected. Meanwhile,we continued to pursue the possibility of enhanced col-lision cross sections in cesium using a Feshbach reso-nance. At that point our Monte Carlo simulations saidthat a ratio of about 150 elastic collisions per trap life-time was required to achieve runaway evaporation. Thisis the condition where the elastic collision rate wouldcontinue to increase as the temperature decreased, andhence evaporation would continue to improve as thetemperature was reduced. We also had reasonable deter-minations of the elastic collision cross sections.

So the plan was to build a simple quadrupole trap thatwould allow very strong squeezing to greatly enhancethe collision rate, combined with a good vacuum systemin order to make sure evaporative cooling worked asexpected. Clearly, there was much to be gained by build-ing a more tightly confining magnetic trap, but the re-quirement of adequate optical access for the MOT,along with engineering constraints on power dissipation,made the design problem complicated.

When constructing a trap for weak-field-seeking at-oms, with the aim of confining the atoms to a spatial sizemuch smaller than the size of the magnets, one would

Rev. Mod. Phys., Vol. 74, No. 3, July 2002

like to use linear gradients. In that case, however, one isconfronted with the problem of the minimum in themagnitude of the magnetic fields (and thus of the con-fining potential) occurring at a local zero in the magneticfield. This zero represents a hole in the trap, a site atwhich atoms can undergo Majorana transitions (Majo-rana, 1931) and thus escape from the trap. If one usesthe second-order gradients from the magnets to providethe confinement, there is a marked loss of confinementstrength. This scaling is discussed by Petrich et al. (1995).We knew that once the atoms became cold enough theywould leak out the hole in the bottom of the trap, butthe plan was to go ahead and get evaporation and worryabout the hole later. We also recognized that even withsuccessful evaporative cooling, and presuming we couldsolve the issue of the hole in the quadrupole trap, therewas still the question of the sign of scattering length,which must be positive to ensure the stability of a largecondensate.

In setting up the new apparatus Eric chose to use ru-bidium. Given the modulo arithmetic that goes into de-termining a scattering length, it seemed fair to treat thescattering lengths of different isotopes as statistically in-dependent events, and rubidium with its two stable iso-topes offered two rolls of the dice for the same lasersystem. Eric then purchased a set of diode lasers for therubidium wavelength, but of course we kept the originalcesium-tuned diode lasers. The wavelengths of cesiumand of the two rubidium isotopes are sufficiently similarthat in most cases one can use the same optics. Thus wepreserved the option of converting from one species toanother in a matter of weeks. The chances then of Na-ture’s conspiring to make the scattering length negative,for both hyperfine levels, for all three atoms, seemedvery small.

Progress in cold collisions, particularly the experimentand theory of photoassociative collisions, had movedforward so rapidly that by the time we had evaporativelycooled rubidium to close to BEC temperatures a coupleof years later there existed, at the 20% level, values forseveral of the elastic scattering lengths. In particular, weknew that it was positive for the 2,2 state of Rb-87(Thorsheim et al., 1987; Lett et al., 1993; Miller et al.,1993; Abraham et al., 1995; Gardner et al., 1995; McAl-exander et al., 1995).

Our original idea for the quadrupole trap experimentwas to pulse a burst of rubidium into our cell, where wewould catch a large sample in the MOT and then hold itas the residual rubidium was quickly pumped away, leav-ing a long trap lifetime. We, particularly Eric’s postdoc,Mike Anderson, spent many frustrating months discov-ering how difficult this seemingly simple idea was to ac-tually implement in practice. The manner in which ru-bidium interacted with glass and stainless-steel surfacesconspired to make this so difficult we finally gave up. Weended up going with a far-from-optimum situation ofworking with extremely low rubidium pressure and do-ing our best at maximizing the number of atoms cap-tured in the MOT from this feeble vapor and enhancingthe collision rate for those relatively few atoms as much

Page 8: Nobel Lecture: Bose-Einstein condensation in a dilute gas ...duine102/summerschoolTP/BECNobel.pdf · experimental Bose-Einstein condensation. For a somewhat more scholarly treatment

882 E. A. Cornell and C. E. Wieman: BEC in a dilute gas

as possible. We recognized that this was a major com-promise, but we had been trying to evaporate for sometime, and we were getting impatient! We had no stom-ach for building another apparatus just to see evapora-tion. Fortunately we were able to find two key elementsto enhance the MOT loading and density. First was theuse of a dark-spot MOT in which there is a hole in thecenter of the MOT beams so the atoms are not excited.This technique had been demonstrated by Ketterle(Ketterle et al., 1993) as a way to greatly enhance thedensity of atoms in a MOT under conditions of a veryhigh loading rate. The number of atoms we could load inour vapor cell MOT with very low rubidium vapor wasdetermined by the loading rate over the loss rate. In thiscase the loss rate was the photoassociative collisions wehad long before found to be important for losses fromMOT’s. The dark-spot geometry reduced this two-bodyphotoassociative loss in part because in our conditions itreduced the density of atoms in the MOT (Andersonet al., 1994).

Using this approach we were able to obtain 108 atomsin the MOT collected out of a very low vapor back-ground (so that magnetic trap lifetime was greater than100 s). The second key element was the invention of thecompressed MOT (CMOT), a technique for substan-tially enhancing the density of atoms in the MOT on atransient basis. For the CMOT, the MOT was filled andthen the field gradient and laser detuning were suddenlychanged to greatly suppress the multiple photon scatter-ing. This produced much higher densities and cloudswhose shape was a much better match to the desiredshape of the cloud in the magnetic trap. This was a verytransient effect because the losses from the MOT weremuch larger under these conditions, but that was notimportant; the atoms needed only to be held for themilliseconds required before they were transferred tothe magnetic trap (Petrich et al., 1994; see Fig. 5). Withthese improvements and a quadrupole trap that pro-vided substantial squeezing, we were able to finally dem-onstrate evaporative cooling in rubidium.

Cooling by evaporation is a process found throughoutNature. Whether the material being cooled is an atomicnucleus or the Atlantic Ocean, the rate of natural evapo-ration and the minimum temperature achievable arelimited by the particular fixed value of the work functionof the evaporating substance. In magnetically confinedatoms, no such limit exists, because the work function issimply the height of the lowest point in the rim of theconfining potential. Hess (1986) pointed out that, by per-turbing the confining magnetic fields, one could makethe work function of a trap arbitrarily low; as long asfavorable collisional conditions persist, there is no lowerlimit to the temperatures attainable in this forced evapo-ration.

Pritchard (Pritchard et al., 1989) pointed out thatevaporation could be performed more conveniently ifthe rim of the trap were defined by an rf-resonance con-dition, rather than simply by the topography of the mag-netic field; experimentally, his group made first use ofposition-dependent rf transitions to selectively transfer

Rev. Mod. Phys., Vol. 74, No. 3, July 2002

magnetically trapped sodium atoms between Zeemanlevels and thus characterized their temperature (Martinet al., 1988). In our experiment we used Pritchard’s tech-nique of an rf field to selectively evaporate.

It was a great relief to see evaporative cooling of laserprecooled, magnetically trapped atoms finally work, aswe had been anticipating it would for so many years.Unfortunately, it worked exactly as well, but no better,than we had anticipated. The atoms were cooled toabout 40 mK and then disappeared, at just the tempera-ture we had estimated they would be lost, through thehole in the bottom of the quadrupole trap. Eric came upwith an idea that solved this problem. It was a design fora new type of trap that required relatively little modifi-cation to the apparatus and so was quickly implemented.This was the Time Orbiting Potential (TOP) trap inwhich a small rotating magnetic field was added to thequadrupole field (Petrich et al., 1995). This moved thefield zero in an orbit faster than the atoms could follow.It was the perfect solution to our problem.

Mike Anderson, another postdoc, Wolfgang Petrich,and graduate student Jason Ensher quickly imple-mented this design. Their efforts were spurred on by therealization that there were several other groups who hadnow demonstrated or were known to be on the verge ofdemonstrating evaporative cooling in alkalis in the pur-suit of BEC. The TOP design worked well, and thesamples were cooled far colder, in fact too cold for us toreliably measure. We had been measuring temperaturesimply by looking at the spatial size of the cloud in themagnetic trap. As the temperature was reduced the sizedecreased, but we were now reaching temperatures solow that the size had reached the resolution limit of theoptical system. We saw dramatic changes in the shapesof the images as the clouds became very small, but weknew that a variety of diffraction and aberration effectscould greatly distort images when the sample size be-came only a few wavelengths in size, so our reaction tothese shapes was muted, and we knew we had to havebetter diagnostics before we could have meaningful re-sults. Here we were helped by our long experience instudying various trapped clouds over the years. We al-ready knew the value of turning the magnetic trap off tolet the cloud expand and then imaging the expandedcloud to get a measure of the momentum distribution inthe trap. Since the trap was harmonic, the momentumdistribution and the original density distribution werenearly interchangeable. Unfortunately, once the mag-netic field was off, the atoms not only expanded but alsosimply fell under the influence of gravity. We found thatthe atoms tended to fall out of the field of view of ourmicroscope before they had sufficiently expanded. Thefinal addition to the apparatus was a supplementarymagnetic coil, which provided sufficient field gradient tocancel the effects of gravity while minimizing any pertur-bation to the relative ballistic trajectories of the expand-ing atoms.

Anderson, Ensher, and a new graduate student, MikeMatthews (Fig. 6), worked through a weekend to installthe antigravity coil and, after an additional day or two of

Page 9: Nobel Lecture: Bose-Einstein condensation in a dilute gas ...duine102/summerschoolTP/BECNobel.pdf · experimental Bose-Einstein condensation. For a somewhat more scholarly treatment

883E. A. Cornell and C. E. Wieman: BEC in a dilute gas

trial and error, got the new field configuration shimmedup. By June 5, 1995 the new technology was workingwell and we began to look at the now greatly expandedclouds. To our delight, the long-awaited two-componentdistribution was almost immediately apparent (Fig. 7)when the samples were cooled to the regime where BECwas expected. The excitement was tempered by the con-

FIG. 5. Wolfgang Petrich working on CMOT [Color].

FIG. 6. From left, Mike Anderson, Debbie Jin, Mike Mat-thews, and Jason Ensher savor results of early BEC experi-ment [Color].

Rev. Mod. Phys., Vol. 74, No. 3, July 2002

cern that after so many years of anticipating two com-ponent clouds as a signature of BEC, we might be fool-ing ourselves.

Almost from the beginning of the search for BEC, itwas recognized (Lovelace and Tommila, 1987) that asthe sample started to condense, there would be a spikein the density and momentum distributions correspond-ing to the macroscopic population of the ground state.This would show up as a second component on top ofthe much broader normal thermal distribution of uncon-densed atoms. This was the signature we had been hop-ing to see from our first days of contemplating BEC. Thesize of the BEC component in these first observationsalso seemed almost too good to be true. In those days itwas known that in the much higher density of the con-densate, three-body recombination would be a moredominant effect than in the lower-density uncondensedgas. For hydrogen it was calculated that the condensedcomponent could never be more than a few percent ofthe sample. The three-body rate constants were totallyunknown for alkali atoms at that time, but because ofthe H results it still seemed reasonable to expect thecondensate component might only be a modest fractionof the total sample. But in our first samples we saw itcould be nearly 100%! In the light of the prevailingmyth of unattainability that had grown up around BECover the years, our observations seemed too good to betrue. We were experienced enough to know that whenresults in experimental physics seem too good to be true,they almost always are! We worried that in our enthusi-asm we might confuse the long-desired BEC with somespurious artifact of our imaging system.

However, our worries about the possibility of deludingourselves were quickly and almost entirely alleviated bythe anisotropy of the BEC cloud. This was a very dis-tinctive signature of BEC, the credibility of which wasgreatly enhanced to us by the fact that it first revealeditself in the experiment, and then we recognized its sig-nificance, rather than vice versa. It was a somewhat for-tuitous accident that the TOP trap provided a distinctlyanisotropic trapping potential, since we did not appreci-ate its benefits until we saw the BEC data. A normalthermal gas (in the collisionally thin limit) released froman anisotropic potential will spread out isotropically.This is required by the equipartition theorem. However,a Bose-Einstein condensate is a quantum wave and so itsexpansion is governed by a wave equation. The moretightly confined direction will expand the most rapidly, amanifestation of the uncertainty principle. Seeing theBEC component of our two-component distribution dis-play just this anisotropy, while the broader uncondensedportion of the sample observed at the same time, withthe same imaging system, remained perfectly isotropic(as shown in Fig. 8), provided the crucial piece of cor-roborating evidence that this was the long-awaited BEC.

By coincidence we were scheduled to present progressreports on our efforts to achieve BEC at three interna-tional conferences in the few weeks following these ob-servations (Anderson et al., 1996). Nearly all the expertsin the field were represented at one or more of these

Page 10: Nobel Lecture: Bose-Einstein condensation in a dilute gas ...duine102/summerschoolTP/BECNobel.pdf · experimental Bose-Einstein condensation. For a somewhat more scholarly treatment

884 E. A. Cornell and C. E. Wieman: BEC in a dilute gas

FIG. 7. Three density distributions of the expanded clouds of rubidium atoms at three different temperatures. The appearance ofthe condensate is apparent as the narrow feature in the middle image. On the far right, nearly all the atoms in the sample are inthe condensate. The original experimental data were two-dimensional black and white shadow images, but these images have beenconverted to three dimensions and given false color density contours [Color].

conferences, and the data were sufficient to convince themost skeptical of them that we had truly observed BEC.This consensus probably facilitated the rapid refereeingand publication of our results.

Rev. Mod. Phys., Vol. 74, No. 3, July 2002

In the original TOP-trap apparatus we were able toobtain so-called pure condensates of a few thousand at-oms. By pure condensates we meant that nearly all theatoms were in the condensed fraction of the sample.

FIG. 8. Looking down on the three images of Figure 7 (Anderson et al., 1995). The condensate in B and C is clearly elliptical inshape [Color].

Page 11: Nobel Lecture: Bose-Einstein condensation in a dilute gas ...duine102/summerschoolTP/BECNobel.pdf · experimental Bose-Einstein condensation. For a somewhat more scholarly treatment

885E. A. Cornell and C. E. Wieman: BEC in a dilute gas

Samples of this size were easily large enough to image.Over the few months immediately following the originalobservation, we undertook the process of a technologi-cal shoring up of the machine, until the machine reachedthe level of reliability necessary to crank out condensateafter reproducible condensate. This set the stage for thefirst generation of experiments characterizing the prop-erties of the condensate, most notably the condensateexcitation studies discussed below.

Although by 1995 and 1996 we were able to carry outa number of significant BEC experiments with the origi-nal TOP-trap machine, even by 1994, well before theoriginal condensates were observed, we had come to re-alize the limitations of the single-cell design. Our effortsto modulate the vapor pressure were not very successful,which forced us to operate at a steady-state rubidiumvapor pressure. Choosing the value of vapor pressure atwhich to operate represented a compromise betweenour need to fill the vapor-cell MOT with as many atomsas possible and our need to have the lifetime in the mag-netic trap as long as possible. The single-cell design alsocompelled us to make a second compromise, this timeover the size of the glass cell. The laser beams of theMOT enter the cell through the smooth, flat region ofthe cell; the larger the glass cell, the larger the MOTbeams, and the more atoms we could herd into the MOTfrom the room-temperature background vapor. On theother hand, the smaller the glass cell, the smaller theradii of the magnetic coils wound round the outside ofthe cell, and the stronger the confinement provided bythe magnetic trap. Hans Rohner in the JILA specialtyshop had learned how (Rohner, 1994) to create glasscells with the minimum possible wasted area. But evenwith the dead space between the inner diameter of themagnetic coils and the outer diameter of the clear glasswindows made as small as it could be, we were con-fronted with an unwelcome tradeoff.

Thus, in 1994, in parallel with our efforts to push ashard as we could toward BEC in our original, single-cellTOP trap, we began working on a new technology thatwould avoid this painful tradeoff. This approach was amodified version of our old multiple loading scheme inwhich many loads from a MOT were transferred to amagnetic trap in a differentially pumped vacuum cham-ber. That approach had been defeated by the difficulty intransferring atoms from MOT to magnetic trap withoutlosing phase-space density. There was no dissipation inthe magnetic trap to compensate for a slightly too hardor too soft push from one trap to the other. This made usrecognize the importance of having dissipation in thesecond trap, and so we went to a system in which atomswere captured in a large-cell MOT in a region of highrubidium pressure, and then transferred through a smalltube into a second, small-cell MOT in a low-pressureregion. This eliminated the previous disadvantages whilepreserving the advantages of multiple loading to getmuch larger numbers of trapped atoms in a low-vacuumregion. The approach worked well, particularly when wefound that simple strips of plastic refrigerator magnetmaterial around the outside of the transfer tube between

Rev. Mod. Phys., Vol. 74, No. 3, July 2002

the two traps provided an excellent guide to confine theatoms as they were pushed from one trap to the other(Myatt et al., 1996).

With this scheme we were still able to use inexpensivelow-power diode lasers to obtain about one hundredtimes more atoms in the magnetic trap than in our singleMOT-loaded TOP magnetic trap and with a far longerlifetime; we saw trap lifetimes up to 1000 s in the doubleMOT magnetic trap. This system started working in1996 and it marked a profound difference in the easewith which we could make BEC (Myatt et al., 1997). Inthe original BEC experiment everything had to be verywell optimized to achieve the conditions necessary forrunaway evaporative cooling and thereby BEC. In thedouble MOT system there were orders of magnitude tospare. Not only did this allow us to routinely obtainmillion-atom pure condensates, but it also meant that wecould dispense with the dark-spot optical configurationwith its troublesome alignment. We could be much lessprecise with many other aspects of the experiment aswell.

The first magnetic trap we used with the double-MOTBEC machine was not a TOP trap, but instead was ourold baseball-style Ioffe-Pritchard trap. The baseball coiltrap is rather complementary to the TOP trap in thateach has unique capabilities. For example, the geometryof the TOP trap potential can be changed over a widerange, although the range of dc fields is quite limited. Incontrast, the geometry of the baseball coil trap potentialcan be varied only by small amounts, but the dc biasfield can be easily varied over hundreds of gauss. Thus in1996, when we upgraded the original BEC machine toincorporate the double-MOT technology, we preservedthe TOP trap coil design. Each is well suited to certaintypes of experiments, as will be evident in the discus-sions below.

With the double-MOT setups we were able to rou-tinely make million-atom condensates in a highly reli-able manner in both TOP and baseball-type magnetictraps. These were used to carry out a large number ofexperiments with condensates over the period from 1996to the present. Some of our favorite experiments arebriefly discussed below.

FAVORITE EXPERIMENTS

Collective excitations

In this section, by excitations we mean coherent fluc-tuations in the density distribution. Excitation experi-ments in dilute-gas BEC have been motivated by twomain considerations. First, a Bose-Einstein condensateis expected to be a superfluid, and a superfluid is definedby its dynamical behavior. Studying excitations is an ob-vious initial step toward understanding dynamical be-havior. Second, in experimental physics a precision mea-surement is almost always a frequency measurement,and the easiest way to study an effect with precision is tofind an observable frequency that is sensitive to that ef-fect. In the case of dilute-gas BEC, the observed fre-

Page 12: Nobel Lecture: Bose-Einstein condensation in a dilute gas ...duine102/summerschoolTP/BECNobel.pdf · experimental Bose-Einstein condensation. For a somewhat more scholarly treatment

886 E. A. Cornell and C. E. Wieman: BEC in a dilute gas

quency of standing-wave excitations in a condensate is aprecise test of our understanding of the effect of inter-actions.

BEC excitations were first observed by Jason Ensher,Mike Matthews, and then-postdoc Debbie Jin, using de-structive imaging of expanded clouds (Jin et al., 1996).The nearly zero-temperature clouds were coherently ex-cited (see below), then allowed to evolve in the trap forsome particular dwell time, and then rapidly expandedand imaged via absorption imaging. By repeating theprocedure many times with varying dwell times, thetime-evolution of the condensate density profile can bemapped out. From these data, frequencies and dampingrates can be extracted. In axially symmetric traps, exci-tations can be characterized by their projection of angu-lar momentum on the axis. The perturbation on the den-sity distribution caused by the excitation of lowest-lyingm50 and m52 modes can be characterized as simpleoscillations in the condensate’s linear dimensions. Figure9 shows the widths of an oscillating condensate as afunction of dwell time.

A frequency-selective method for driving the excita-tions is to modulate the trapping potential at the fre-quency of the excitation to be excited (Jin et al., 1996).Experimentally this is accomplished by summing a smallac component onto the current in the trapping magnets.In a TOP trap, it is convenient enough to independentlymodulate the three second-order terms in the transversepotential. By controlling the relative phase of thesemodulations, one can impose m50, m52, or m522symmetry on the excitation drive.

There have been a very large number of theory paperspublished on excitations; much of this work is reviewedby Dalfovo et al. (1999). All the zero-temperature,small-amplitude excitation experiments published to

FIG. 9. Zero-temperature excitation data from Jin et al.(1996). A weak m50 modulation of the magnetic trappingpotential is applied to a 4500-atom condensate in a 132-Hz(radial) trap. Afterward, the freely evolving response of thecondensate shows radial oscillations. Also observed is a sym-pathetic response of the axial width, approximately 180° out ofphase. The frequency of the excitation is determined from asine wave fit to the freely oscillating cloud widths.

Rev. Mod. Phys., Vol. 74, No. 3, July 2002

date have been very successfully modeled theoretically.Quantitative agreement has been by and large verygood; small discrepancies can be accounted for by as-suming reasonable experimental imperfections with re-spect to the T50 and small-amplitude requirements oftheory.

The excitation measurements discussed above werethen revisited at nonzero temperature (Jin et al., 1997).The frequency of the condensate excitations was clearlyobserved to depend on the temperature, and the damp-ing rates showed a strong temperature dependence. Thiswork is important because it bears on the little-studiedfinite-temperature physics of interacting condensates.Connection with theory (Hutchinson et al., 1997; Doddet al., 1998; Fedichev and Shlyapnikov, 1998) remainssomewhat tentative. The damping rates, which are ob-served to be roughly linear in temperature, have beenexplained in the context of Landau damping (Liu, 1997;Fedichev et al., 1998). The frequency shifts are difficultto understand, in large part because the data so far havebeen collected in a theoretically awkward, intermediateregime: the cloud of noncondensate atoms is neither sothin as to have completely negligible effect on the con-densate, nor so thick as to be deeply in the hydrody-namic (HD) regime. In this context, hydrodynamic re-gime means that the classical mean free path in thethermal cloud is much shorter than any of its physicaldimensions. In the opposite limit, the collisionless re-gime, there are conceptual difficulties with describingthe observed density fluctuations as collective modes.Recent theoretical work suggests that good agreementwith experiment may hinge on correctly including therole of the excitation drive (Stoof, 2000; Jackson andZaremba, 2002).

Two-component condensates

As mentioned above, the double-MOT system madeit possible to produce condensates even if one werequite sloppy with many of the experimental parameters.One such parameter was the spin state in which the at-oms are optically pumped before being loaded into themagnetic trap. As our student Chris Myatt was tinkeringaround setting up the evaporation one day, he noticed,to his surprise, that there seemed to be two differentclouds of condensate in the trap. They were roughly atthe locations expected for the 2,2 and 1,21 spin states tosit, but that seemed impossible to us because these twostates could undergo spin-exchange collisions that wouldcause them to be lost from the trap, and the spin-exchange collision cross sections were thought to beenormous. After extensive further studies to try andidentify what strange spurious effect must be responsiblefor the images of two condensate clouds we came torealize that they had to be those two spin states. By aremarkable coincidence, the triplet and singlet phaseshifts are identical and so at ultralow temperatures thespin-exchange collisions are suppressed in 87Rb by threeto four orders of magnitude! This suppression meantthat the different spin species could coexist and their

Page 13: Nobel Lecture: Bose-Einstein condensation in a dilute gas ...duine102/summerschoolTP/BECNobel.pdf · experimental Bose-Einstein condensation. For a somewhat more scholarly treatment

887E. A. Cornell and C. E. Wieman: BEC in a dilute gas

mixtures could be studied. In early work we showed thatone could carry out sympathetic cooling to make BECby evaporating only one species and using it as a coolingfluid to chill the second spin state (Myatt et al., 1997).We also were able to see how the two condensates inter-acted and pushed each other apart, excluding all but asmall overlap in spite of the fact that they were highlydilute gases.

These early observations stimulated an extensive pro-gram of research on two-component condensates. AfterMyatt’s original measurements (Myatt et al., 1997), ourwork in this field, led by postdoc David Hall, concen-trated on the 1,21 and 2,11 states (see Fig. 10) becausethey could be coherently interconverted using two-photon (microwave plus rf) transitions and they hadnearly identical magnetic moments and so saw nearlythe same trapping potentials (Matthews et al., 1998).When the two-photon radiation field is turned off, therate of spontaneous interconversion between the twospin species essentially vanishes, and moreover the opti-cal imaging process readily distinguishes one speciesfrom the other, as their difference in energy (6.8 GHz) isvery large compared to the excited-state linewidth. Inthis situation, one may model the condensate dynamicsas though there were two distinct quantum fluids in thetrap. Small differences in scattering length make the twofluids have a marginal tendency to separate spatially, atleast in an inhomogeneous potential, but the interspe-cies healing length is long so that in the equilibrium con-figuration there is considerable overlap between the twospecies (Hall et al., 1998a, 1998b). On the other hand,the presence of a near-resonant two-photon couplingdrive effectively brings the two energy levels quite closeto one another: on resonance, the corresponding dressedenergy levels are separated only by the effective Rabifrequency for the two-photon drive. In this limit, onemay in a certain sense think of the condensate as beingdescribed by a two-level, spinor field (Cornell et al.,1998; Matthews et al., 1999b).

We got a lot of mileage out of this system and con-tinue to explore its properties today. One of the more

FIG. 10. Energy-level diagram for ground electronic state of87Rb. The first condensates were created in the 2,2 state. Mix-tures containing the 2,2 and 1,21 state were found to coexist.In later studies we created condensates in the 1,21 state andthen excited it to the 2,1 state using a microwave plus rf two-phonon transition.

Rev. Mod. Phys., Vol. 74, No. 3, July 2002

dramatic experiments we did in the two-level conden-sate was the creation, via a sort of wave-function engi-neering, of a quantized vortex. In this experiment wemade use of both aspects of the two-level system—thedistinguishable fluids and the spinor gas. Starting with anear-spherical ball of atoms, all in the lower spin state,we applied the two-photon drive for about 100 ms. Atthe same time, we illuminated the atoms with an off-resonant laser beam whose intensity varied both in timeand in space. The laser beam was sufficiently far fromresonance that by itself it did not cause the condensateto transition from state to state, but the associated acStark shift was large enough to affect the resonant prop-erties of the two-photon drive. The overall scheme isdescribed by Matthews et al. (1999a) and Williams andHolland (1999). The net effect was to leave the atomsnear the center of the ball of atoms essentially unper-turbed, while converting the population in an equatorialbelt around the ball into the upper spin state. This con-version process also imposed a winding in the quantumphase, from 0 around to two pi, in such a way that by thetime the drive was turned off, the upper-spin-state atomswere in a vortex state, with a single quantum of circula-tion (see Fig. 11). The central atoms were nonrotatingand, like the pimento in a stuffed olive, served only tomark the location of the vortex core. The core atomscould in turn be selectively blasted away, leaving theupper-state atoms in a bare vortex configuration, whosedynamic properties were shown by postdoc BrianAnderson and grad student Paul Haljan to be essentiallythe same as those of the filled vortex (Anderson et al.,2000).

Coherence and condensate decay

One of our favorite BEC experiments was simply tolook at how a condensate goes away (Burt et al., 1997).The attraction of this experiment is its inherent simplic-ity combined with the far-reaching implications of theresults. Although it was well established that conden-sates lived for a finite period, fractions of a second tomany seconds depending on conditions, no one hadidentified the actual process by which atoms were beinglost from the condensate. To do this our co-workersChris Myatt, Rich Ghrist, and Eric Burt simply madecondensates and carefully watched the number of atomsand shape of the condensate as a function of time. Fromthese data we determined that the loss process variedwith the cube of the density, and hence must be three-body recombination. This was rather what we had ex-pected, but it was nice to have it confirmed. In the pro-cess of this measurement we also determined the three-body rate constant, and this was more interesting.Although three-body rate constants still cannot be accu-rately calculated, it was predicted long ago (Kagan et al.,1985) that they should depend on the coherence proper-ties of the wave function. In a normal thermal samplethere are fluctuations and the three-body recombinationpredominantly takes place at high-density fluctuations.If there is higher-order coherence, however, as one hasin macroscopically occupied quantum states such as a

Page 14: Nobel Lecture: Bose-Einstein condensation in a dilute gas ...duine102/summerschoolTP/BECNobel.pdf · experimental Bose-Einstein condensation. For a somewhat more scholarly treatment

888 E. A. Cornell and C. E. Wieman: BEC in a dilute gas

FIG. 11. Condensate images showing the first BEC vortex and the measurement of its phase as a function of azimuthal angle: (a)the density distribution of atoms in the upper hyperfine state after atoms have been put in that state in a way that forms a vortex;(b) the same state after a pi/2 pulse has been applied that mixes upper and lower hyperfine states to give an interferogramreflecting the phase distribution of the upper state; (c) residual condensate in the lower hyperfine state from which the vortex wasformed that interferes with a to give the image shown in (b); (d) a color map of the phase difference reflected in (b); (e) radialaverage at each angle around the ring in (d). The data are repeated after the azimuthal angle 2p to better show the continuityaround the ring. This shows that the cloud shown in (a) has the 2p phase winding expected for a quantum vortex with one unit ofangular momentum. From Matthews et al., 1999a [Color].

FIG. 12. Bosenova explosion from Roberts et al. (2001). Fromtop to bottom these images show the evolution of the cloudfrom 0.2 to 4.8 ms after the interaction was made negative,triggering a collapse. On the left the explosion products arevisible as a blue glow expanding out of the center, leaving asmall condensate remnant that is unchanged at subsequenttimes. On the right is the same image amplified by a factor of3 to better show the 200 nK explosion products [Color].

Rev. Mod. Phys., Vol. 74, No. 3, July 2002

Page 15: Nobel Lecture: Bose-Einstein condensation in a dilute gas ...duine102/summerschoolTP/BECNobel.pdf · experimental Bose-Einstein condensation. For a somewhat more scholarly treatment

889E. A. Cornell and C. E. Wieman: BEC in a dilute gas

single-mode laser, or as was predicted to exist in a dilutegas BEC, there should be no such density fluctuations.On this basis it was predicted that the three-body rateconstant in a Bose-Einstein condensate would be 3 fac-torial or 6 times lower than what it would be for thesame atoms in a thermal sample. It is amusing that sucha relatively mundane collision process can be used toprobe the quantum correlations and coherence in thisfashion. After measuring the three-body rate constant inthe condensate we then repeated the measurement in avery cold but uncondensed sample. The predicted factorof 6 (actually 7.462.6) was observed, thereby confirmingthe higher-order coherence of BEC (Burt et al., 1997).

Feshbach resonance physics

In 1992 Eric Cornell and Chris Monroe realized thatdipole collisions at ultralow temperatures might have in-teresting dependencies on magnetic field, as discussed inthe Appendix. With this in mind we approached Boud-wijn Verhaar about calculating the magnetic-field depen-dencies of collisions between atoms in the lower F spinstates. When he did this calculation he discovered (Ties-inga et al., 1993) that there were dramatic resonances inall the cross sections as a function of magnetic field thatare now known as Feshbach resonances because of theirsimilarity to scattering resonances described by HermanFeshbach in nuclear collisions. From the beginning Ver-haar appreciated that these resonances would allow oneto tune the s-wave scattering length of the atoms andthereby change both the elastic collision cross sectionsand the self-interaction in a condensate, although thiswas several years before condensates had been created.

In 1992 we hoped that these Feshbach resonanceswould give us a way to create enormous elastic collisioncross sections that would facilitate evaporative cooling.With this in mind we attempted to find Feshbach reso-nances in the elastic scattering of first cesium and then,with postdoc Nate Newbury, rubidium. These experi-ments did provide us with elastic scattering cross sec-tions (Monroe et al., 1993; Newbury et al., 1995), butwere unable to locate the few-gauss-wide Feshbach reso-nances in the thousand-gauss range spanned by thentheoretical uncertainty.

By 1997 the situation had dramatically changed, how-ever. A large amount of work on cold collisions, BECproperties, and theoretical advances provided accuratevalues for the interaction potentials, and so we werefairly confident that there was likely to be a reasonablywide Feshbach resonance in rubidium 85 that was within20 or 30 gauss of 150 G. This was a quite convenient biasfield at which to operate our baseball magnetic trap, sowe returned to the Feshbach resonance in the hope thatwe could now use it to make a Bose-Einstein condensatewith adjustable interactions.

The time was clearly ripe for Feshbach resonancephysics. Within a year Ketterle (Inouye et al., 1998) sawa resonance in sodium through enhanced loss of BEC,Dan Heinzen (Courteille et al., 1998) detected a Fesh-bach resonance in photoassociation in 85Rb, we (Rob-

Rev. Mod. Phys., Vol. 74, No. 3, July 2002

erts et al., 1998; notably students Jake Roberts and NeilClaussen) detected the same resonance in the elasticscattering cross section, and Chu (Vuletic et al., 1999)detected Feshbach resonances in cesium. Our expecta-tions that it would be as easy or easier to form BEC in85Rb as it was in 87Rb and then use this resonance tomanipulate the condensate were sadly naive, however.Due to enhancement of bad collisions by the Feshbachresonance, it was far more difficult and could only beaccomplished by following a complicated and precariousevaporation path. However, by finding the correct pathand cooling to 3 nK we were able to obtain pure 85Rbcondensates of 16 000 atoms (Roberts et al., 2001).

The scattering length of these condensates could thenbe readily adjusted by varying the magnetic field over afew gauss in the vicinity of the Feshbach resonance(Cornish et al., 2000). This has opened up a wide rangeof possible experiments, from studying the instability ofcondensates when the self-interaction is sufficiently at-tractive (negative a) to exploring the development ofcorrelations in the wave function as the interactions aremade large and repulsive. This regime provides one witha new way to probe such disparate subjects as molecularBose-Einstein condensates and the quantum behavior ofliquids, where there is a high degree of correlation. Thiswork represents some of the most recent BEC experi-ments, but almost everything we have explored with thissystem has shown dramatic and unexpected results.Thus it is clear that we are far from exhausting the fullrange of interesting experiments that are yet to be car-ried out with BEC.

In the first of these Feshbach resonance experimentsour students Jake Roberts, Neil Claussen, and postdocSimon Cornish suddenly changed the magnetic field tomake a negative. We observed that, as expected, thecondensate became unstable and collapsed, losing alarge number of atoms (Roberts et al., 2001). The dy-namics of the collapse process were quite remarkable.The condensate was observed to shrink slightly and thenundergo an explosion in which a substantial fraction ofthe atoms were blown off (Donley, 2001). A large frac-tion of the atoms also simply vanished, presumably turn-ing into undetectable molecules or very energetic atoms,and finally a small cold stable remnant was left behindafter the completion of the collapse. This process is il-lustrated in Fig. 12. Because of its resemblance (on avastly lower energy scale) to a core collapse supernova,we have named this the Bosenova. There is now consid-erable theoretical effort to model this process andprogress is being made. However, as yet there is no clearexplanation of the energy and anisotropy of the atoms inthe explosion, the fraction of vanished atoms, and thesize of the cold remnant. One of the more puzzling as-pects is that the cold remnant can be far larger than thecondensate stability condition that determines the col-lapse point would seem to allow (Donley, 2001).

Another very intriguing result of Feshbach resonancestudies in 85Rb was observed when our students NeilClaussen and Sarah Thompson and postdoc ElizabethDonley quickly jumped the magnetic field close to the

Page 16: Nobel Lecture: Bose-Einstein condensation in a dilute gas ...duine102/summerschoolTP/BECNobel.pdf · experimental Bose-Einstein condensation. For a somewhat more scholarly treatment

890 E. A. Cornell and C. E. Wieman: BEC in a dilute gas

resonance while keeping the scattering length positive.They found that they could observe the sample oscillateback and forth between being an atomic and a molecularcondensate as a function of time after the sudden per-turbation (Donley et al., 2002). This curious system of aquantum superposition of two chemically distinct spe-cies will no doubt be a subject of considerable futurestudy.

An optimistic appendix

Until a new technology comes along to replace evapo-rative cooling, the crucial issue in creating BEC with anew atom is collisions. In practice, this means that plan-ning a BEC experiment with a new atom requires learn-ing to cope with ignorance. It is easy to forget that es-sentially nothing is known about the ultralow-temperature collisional properties of any atomic ormolecular species that is not an atom in the first row ofthe Periodic Table. One cannot expect theorists to re-lieve one’s ignorance. Interatomic potentials derivedfrom room-temperature spectroscopy are generally notadequate to allow theoretical calculations of cold elasticand inelastic collision rates, even at the order-of-magnitude level. Although the cold collisional proper-ties of a new atom can be determined, this is a majorendeavor, and in most cases it is easier to discoverwhether evaporation will work by simply trying it.

Launching into such a major new project without anyassurances of success is a daunting prospect, but we be-lieve that, if one works hard enough, the probability thatany given species can be evaporatively cooled to thepoint of BEC is actually quite high. The scaling argu-ments presented below in support of this assertion arelargely the same as those that originally encouraged usto pursue BEC in alkalis, although with a bit more re-finement provided by age and experience.

Although there is an extensive literature now onevaporative cooling, the basic requirement is simply thatthere be on the order of 100 elastic collisions per atomper lifetime of the atoms in the trap. Since the lifetimeof the atoms in the trap is usually limited by collisions,the requirement can be restated: the rate of elastic col-lisions must be about two orders of magnitude higherthan the rate of bad collisions. As mentioned above,there are three bad collisional processes, and these eachhave different dependencies on atomic density in thetrap, n : background collisions (independent of n), two-body dipolar relaxation (an), and three-body recombi-nation (an2). The rate for elastic collisions is nsv ,where n is the mean density, s is the zero-energy s-wavecross section, and v is the mean relative velocity. Therequirement of 100 elastic-to-inelastic collisions mustnot only be satisfied immediately after the atoms areloaded into the trap, but also as evaporation proceedstoward larger n and smaller v . With respect to evapo-rating rubidium 87 or the lower hyperfine level of so-dium 23, Nature has been kind. One need only arrangefor the initial trapped cloud to have sufficiently large n ,and design a sufficiently low-pressure vacuum chamber,

Rev. Mod. Phys., Vol. 74, No. 3, July 2002

and evaporation works. The main point of this section,however, is that evaporation is likely to be possible evenwith less favorable collision properties.

Considering the trap loss processes in order, first ex-amine background loss. Trap lifetimes well in excess ofwhat are needed for 87Rb and Na have been achievedwith standard vacuum technology. For example, we nowhave magnetic trap lifetimes of nearly 1000 s. (This wasa requirement to achieve BEC in 85Rb with its less fa-vorable collisions.) If one is willing to accept the addedcomplications of a cryogenic vacuum system, essentiallyinfinite lifetimes are possible. If the background trap lossis low enough to allow evaporative cooling to begin, itwill never be a problem at later stages of evaporationbecause nv increases.

If dipolar relaxation is to be a problem, it will likelybe late in the evaporative process when the density ishigh and velocity low. There is no easy solution to alarge dipolar relaxation rate in terms of changing thespring constant of the trap or the pressure of the vacuumchamber. Fortunately, one is not required to accept thevalue of dipolar collisions that Nature provides. In fact,all one really has to do is operate the trap with a verylow magnetic bias field in a magnetic trap, or if one usesan optical trap very far off-resonance (such as CO2 la-ser), trap the atoms in the lowest spin state, for whichthere are no dipole collisions. The bias field dependencecomes about because below a field of roughly 5 G, thedipolar rate in the lower hyperfine level drops rapidly tozero. This behavior is simple to understand. At low tem-perature, the incoming collisional channel must bepurely s wave. Dipolar relaxation changes the projectionof spin angular momentum, so to conserve angular mo-mentum the outgoing collisional channel must be dwave or higher. The nonzero outgoing angular momen-tum means that there is an angular momentum barrier inthe effective molecular potential, a barrier of a few hun-dred microkelvin. If the atoms are trapped in the lowerhyperfine state (F51,mF521, in rubidium 87) the out-going energy from a dipolar collision is only the Zeemanenergy in the trapping fields, and for B less than about 5G this energy is insufficient to get the atoms back outover the angular momentum barrier. If relaxation is tooccur, it can happen only at interatomic radii larger thanthe outer turning point of the angular momentum bar-rier. For smaller and smaller fields, the barrier getspushed further out, with correspondingly lower transi-tion rates.

It is unlikely that the three-body recombination rateconstant could ever be so large that three-body recom-bination would be a problem when the atoms are firstloaded from a MOT into the evaporation trap. Asevaporation proceeds, however, just as for the dipolarcollisions, it becomes an increasingly serious concern.Because of its density dependence, however, it can al-ways be avoided by manipulating the trapping potential.Adiabatically reducing the trap confinement has no ef-fect on the phase-space density but it reduces both thedensity and the atom velocity. The ratio of three-body toelastic collisions scales as 1/nv . Therefore, as long as

Page 17: Nobel Lecture: Bose-Einstein condensation in a dilute gas ...duine102/summerschoolTP/BECNobel.pdf · experimental Bose-Einstein condensation. For a somewhat more scholarly treatment

891E. A. Cornell and C. E. Wieman: BEC in a dilute gas

one can continue to turn down the confining strength ofone s trap, one can ensure that three-body recombina-tion will not prevent evaporative cooling all the waydown to the BEC transition.

To summarize, given (i) a modestly flexible magnetictrap, (ii) an arbitrarily good vacuum, (iii) a true groundstate with FÞ0, and (iv) nonpathological collisionalproperties, almost any magnetically trappable speciescan be successfully evaporated to BEC. If one is using avery far off-resonance optical trap (such as a CO2 dipoletrap) one can extend these arguments to atoms that can-not be magnetically trapped. In that case, however, cur-rent technology makes it more difficult to optimize theevaporation conditions than in magnetic traps, and therequirement to turn the trap down sufficiently to avoid alarge three-body recombination rate can be more diffi-cult. Nevertheless, one can plausibly look forward toBEC in a wide variety of atoms and molecules in thefuture.

ACKNOWLEDGMENTS

We acknowledge support from the National ScienceFoundation, the Office of Naval Research, and the Na-tional Institute of Standards and Technology. We havebenefited enormously from the hard work and intellec-tual stimulation of our many students and postdocs.They include Brian Anderson, Mike Anderson, SteveBennett, Eric Burt, Neil Claussen, Ian Coddington,Kristan Corwin, Liz Donley, Peter Engels, Jason Ensher,David Hall, Debbie Jin, Tetsuo Kishimoto, Heather Le-wandowski, Mike Matthews, Jeff McGuirk, Chris Mon-roe, Chris Myatt, Nate Newbury, Scott Papp, CindyRegal, Mike Renn, Jake Roberts, Peter Schwindt, DavidSesko, Michelle Stephens, William Swann, SarahThompson, Thad Walker, Yingju Wang, Richard Watts,Chris Wood, and Josh Zirbel. We also have had helpfrom many other JILA faculty, including John Bohn,Chris Greene, and Murray Holland.

REFERENCES

Abraham, E. R. I., W. I. McAlexander, C. D. Sackett, and R.G. Hulet, 1995, Phys. Rev. Lett. 74, 1315.

Anderson, M. H., J. R. Ensher, M. R. Matthews, C. E. Wie-man, and E. A. Cornell, 1995, Science 269, 198.

Anderson, M. H., J. R. Ensher, M. R. Matthews, C. E. Wie-man, and E. A. Cornell, 1996, in Laser Spectroscopy, XII In-ternational Conference, edited by Massimo Inguscio, MariaAllegrini, and Antonio Sasso (World Scientific, Singapore),p. 3.

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

Anderson, M. H., W. Petrich, J. R. Ensher, and E. A. Cornell,1994, Phys. Rev. A 50, R3597.

Andreev, S. V., V. I. Balykin, V. S. Letokhov, and V. G. Mino-gin, 1981, Zh. Eksp. Teor. Fiz. 34, 463.

Arimondo, E., W. D. Phillips, and F. Strumia, 1991, Eds., Pro-ceedings of the International School of Physics ‘‘EnricoFermi,’’ Course CXVIII, Laser Manipulation of Atoms andIons (North-Holland, Amsterdam).

Rev. Mod. Phys., Vol. 74, No. 3, July 2002

Bagnato, V. S., G. P. Lafyatis, A. G. Martin, E. L. Raab, R. N.Ahmad-Bitar, and D. E. Pritchard, 1987, Phys. Rev. Lett. 58,2194.

Bell, D. A., F. H. Hess, G. P. Kochanski, S. Buchman, L. Pol-lack, Y. M. Xiao, D. Kleppner, and T. J. Greytak, 1986, Phys.Rev. B 34, 7670.

Bjorkholm, J. E., R. R. Freeman, A. Ashkin, and D. B. Pear-son, 1978, Phys. Rev. Lett. 41, 1361.

Bose, S., 1924, Z. Phys. 26, 178.Burt, E. A., R. W. Ghrist, C. J. Myatt, M. J. Holland, E. A.

Cornell, and C. E. Wieman, 1997, Phys. Rev. Lett. 79, 337.Chu, S., 1998, Rev. Mod. Phys. 70, 685.Chu, S., J. E. Bjorkholm, A. Ashkin, and A. Cable, 1986, Phys.

Rev. Lett. 57, 314.Chu, S., L. Holberg, J. E. Bjorkholm, A. Cable, and A. Ashkin,

1985, Phys. Rev. Lett. 55, 48.Chu, S., D. S. Weiss, Y. Shevy, and P. J. Ungar, 1989, Proceed-

ings 11th International Conference on Atomic Physics (WorldScientific, Singapore), pp. 636–638.

Cohen-Tannoudji, C. N., 1998, Rev. Mod. Phys. 70, 707.Cornell, E. A., J. R. Ensher, and C. E. Wieman, 1999, in Pro-

ceedings of the International School of Physics ‘‘EnricoFermi,’’ Course CXL, edited by M. Inguscio, S. Stringari, andC. E. Wieman (Italian Physical Society, Bologna), p. 15. Alsoavailable as cond-mat/9903109.

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

Cornell, E. A., C. Monroe, and C. E. Wieman, 1991, Phys. Rev.Lett. 67, 2439.

Cornish, S. L., N. R. Claussen, J. L. Roberts, E. A. Cornell,and C. E. Wieman, 2000, Phys. Rev. Lett. 85, 1795.

Courteille, P., R. S. Freeland, D. J. Heinzen, F. A. van-Abeelen, and B. J. Verhaar, 1998, Phys. Rev. Lett. 81, 69.

Dalfovo, F., S. Giorgini, L. P. Pitaevskii, and S. Stringari, 1999,Rev. Mod. Phys. 71, 463.

Dalibard, J., and C. Cohen-Tannoudji, 1989, J. Opt. Soc. Am.B 6, 2023.

Dalibard, J., C. Salomon, A. Aspect, E. Arimondo, R. Kaiser,N. Vansteenkiste, and C. Cohen-Tannoudji, 1989, Proceed-ings 11th International Conference on Atomic Physics (WorldScientific, Singapore), pp. 199–214.

Dodd, R. J., M. Edwards, C. W. Clark, and K. Burnett, 1998,Phys. Rev. A 57, R32.

Donley, E. A., 2001, Nature (London) 412, 295.Donley, E. A., N. R. Claussen, S. T. Thompson, and C. E.

Wieman, 2002, Nature (London) 417, 529.Doyle, J. M., 1991a, Phys. Rev. Lett. 67, 603.Doyle, J. M., 1991b, Ph.D. thesis (Massachusetts Institute of

Technology).Einstein, A., 1924, Sitzungsber. K. Preuss. Akad. Wiss., Phys.

Math. Kl. 261.Einstein, A., 1925, Sitzungsber. K. Preuss. Akad. Wiss., Phys.

Math. Kl. 3.Ertmer, W., R. Blatt, J. L. Hall, and M. Zhu, 1985, Phys. Rev.

Lett. 54, 996.Fedichev, P. O. and G. V. Shlyapnikov, 1998, Phys. Rev. A 58,

3146.Fedichev, P. O., G. V. Shlyapnikov, and J. T. M. Walraven,

1998, Phys. Rev. Lett. 80, 2269.Gardner, J. R., R. A. Cline, J. D. Miller, D. J. Heinzen, H. M.

J. M. Boesten, and B. J. Verhaar, 1995, Phys. Rev. Lett. 74,3764.

Page 18: Nobel Lecture: Bose-Einstein condensation in a dilute gas ...duine102/summerschoolTP/BECNobel.pdf · experimental Bose-Einstein condensation. For a somewhat more scholarly treatment

892 E. A. Cornell and C. E. Wieman: BEC in a dilute gas

Gibble, K., S. Chang, and R. Legere, 1995, Phys. Rev. Lett. 75,2666.

Gillaspy, J. D., I. F. Silvera, and J. S. Brooks, 1989, Phys. Rev.B 40, 210.

Hall, D. S., M. R. Matthews, J. R. Ensher, C. E. Wieman, andE. A. Cornell, 1998a, Phys. Rev. Lett. 81, 1539.

Hall, D. S., M. R. Matthews, C. E. Wieman, and E. A. Cornell,1998b, Phys. Rev. Lett. 81, 1543.

Hansch, T. W., and A. L. Schawlow, 1975, Opt. Commun. 13,68.

Hardy, W. N., M. Morrow, R. Jochemsen, and A. Berlinsky,1982, Physica B & C 110, 1964.

Hecht, C. E., 1959, Physica 25, 1159.Hess, H. F., 1986, Phys. Rev. B 34, 3476.Hess, H. F., D. A. Bell, G. P. Kochanski, R. W. Cline, D. Klep-

pner, and T. J. Greytak, 1983, Phys. Rev. Lett. 51, 483.Hutchinson, D. A. W., E. Zaremba, and A. Griffin, 1997, Phys.

Rev. Lett. 78, 1842.Inouye, S., M. R. Andrews, J. Stenger, H. J. Miesner, D. M.

Stamper-Kurn, and W. Ketterle, 1998, Nature (London) 392,151.

Jackson, B., and E. Zaremba, 2002, Phys. Rev. Lett. 88, 180402.Jin, D. S., J. R. Ensher, M. R. Matthews, C. E. Wieman, and E.

A. Cornell, 1996, Phys. Rev. Lett. 77, 420.Jin, D. S., M. R. Matthews, J. R. Ensher, C. E. Wieman, and E.

A. Cornell, 1997, Phys. Rev. Lett. 78, 764.Johnson, B. R., J. S. Denker, N. Bigelow, L. P. Levy, J. H.

Freed, and D. M. Lee, 1984, Phys. Rev. Lett. 52, 1508.Kagan, Y., G. V. Shlyapnikov, and I. A. Vartanyants, 1984,

Phys. Lett. A 101, 27.Kagan, Y., B. V. Svistunov, and G. V. Shlyapnikov, 1985, JETP

Lett. 42, 209.Kagan, Y., I. A. Vartanyants, and G. V. Shlyapnikov, 1981, Sov.

Phys. JETP 54, 590.Ketterle, W., K. B. Davis, M. A. Joffe, A. Martin, and D. E.

Pritchard, 1993, Phys. Rev. Lett. 70, 2253.Ketterle, W., D. S. Durfee, and D. M. Stamper-Kurn, 1999, in

Proceedings of the International School of Physics ‘‘EnricoFermi,’’ Course CXL, edited by M. Inguscio, S. Stringari, andC. E. Wieman (Italian Physical Socity, Bologna), p. 176.

Lagendijk, A., I. F. Silvera, and B. J. Verhaar, 1986, Phys. Rev.B 33, 626.

Letokhov, V., 1968, Zh. Eksp. Teor. Fiz. Pis’ma Red 7, 348.Lett, P. D., K. Helmerson, W. D. Phillips, L. P. Ratliff, S. L.

Rolston, and M. E. Wagshul, 1993, Phys. Rev. Lett. 71, 2200.Lett, P. D., R. N. Watts, C. I. Westbrook, W. D. Phillips, P. L.

Gould, and H. J. Metcalf, 1988, Phys. Rev. Lett. 61, 169.Liu, W. V., 1997, Phys. Rev. Lett. 79, 4056.London, F., 1938, Nature (London) 141, 643.Lovelace, R. V. E., C. Mehanian, T. J. Tomilla, and D. M. Lee,

1985, Nature (London) 318, 30.Lovelace, R. V. E., and T. J. Tommila, 1987, Phys. Rev. A 35,

3597.Luiten, O. J., H. G. C. Werij, I. D. Setija, M. W. Reynolds, T.

W. Hijmans, and J. T. M. Walraven, 1993, Phys. Rev. Lett. 70,544.

Majorana, E., 1931, Nuovo Cimento 8, 107.Martin, A. G., K. Helmerson, V. S. Bagnato, G. P. Lafyatis, and

D. E. Pritchard, 1988, Phys. Rev. Lett. 61, 2431.Matthews, M. R., B. P. Anderson, P. C. Haljan, D. S. Hall, M.

J. Holland, J. E. Williams, C. E. Wieman, and E. A. Cornell,1999b, Phys. Rev. Lett. 83, 3358.

Rev. Mod. Phys., Vol. 74, No. 3, July 2002

Matthews, M. R., B. P. Anderson, P. C. Haljan, D. S. Hall, C.E. Wieman, and E. A. Cornell, 1999a, Phys. Rev. Lett. 83,2498.

Matthews, M. R., D. S. Hall, D. S. Jin, J. R. Ensher, C. E.Wieman, E. A. Cornell, F. Dalfovo, C. Minniti, and S. Strin-gari, 1998, Phys. Rev. Lett. 81, 243.

McAlexander, W. I., E. R. I. Abraham, N. W. M. Ritchie, C. J.Williams, H. T. C. Stoof, and R. G. Hulet, 1995, Phys. Rev. A51, R871.

Migdall, A. L., J. V. Prodan, W. D. Phillips, T. H. Bergeman,and H. J. Metcalf, 1985, Phys. Rev. Lett. 54, 2596.

Miller, J. D., R. A. Cline, and D. J. Heinzen, 1993, Phys. Rev.Lett. 71, 2204.

Monroe, C., 1992, Ph.D. thesis (University of Colorado, Boul-der).

Monroe, C., W. Swann, H. Robinson, and C. Wieman, 1990,Phys. Rev. Lett. 65, 1571.

Monroe, C. R., E. A. Cornell, C. A. Sackett, C. J. Myatt, andC. E. Wieman, 1993, Phys. Rev. Lett. 70, 414.

Myatt, C. J., E. A. Burt, R. W. Ghrist, E. A. Cornell, and C. E.Wieman, 1997, Phys. Rev. Lett. 78, 586.

Myatt, C. J., N. R. Newberry, R. W. Ghrist, S. Loutzenhiser,and C. E. Wieman, 1996, Opt. Lett. 21, 290.

Neuhauser, W., M. Hohenstatt, P. Toschek, and H. Dehmelt,1978, Phys. Rev. Lett. 41, 233.

Newbury, N. R., C. J. Myatt, and C. E. Wieman, 1995, Phys.Rev. A 51, R2680.

Pais, A., 1982, Subtle Is the Lord . . . (Oxford University Press,Oxford), English translation of Einstein’s quotes and histori-cal interpretation.

Petrich, W., M. H. Anderson, J. R. Ensher, and E. A. Cornell,1994, J. Opt. Soc. Am. B 11, 1332.

Petrich, W., M. H. Anderson, J. R. Ensher, and E. A. Cornell,1995, Phys. Rev. Lett. 74, 3352.

Phillips, W. D., 1998, Rev. Mod. Phys. 70, 707.Pritchard, D. E., 1986, in Electronic and Atomic Collisions,

edited by D. C. Lorents, W. Meyerhof, and J. R. Peterson(North-Holland, Amsterdam), p. 593.

Pritchard, D. E., K. Helmerson, and A. G. Martin, 1989, Pro-ceedings 11th International Conference on Atomic Physics(World Scientific, Singapore), pp. 179–197.

Pritchard, D. E., E. L. Raab, V. Bagnato, C. E. Wieman, and R.N. Watts, 1986, Phys. Rev. Lett. 57, 310.

Prodan, J., A. Migdall, W. D. Phillips, I. So, H. Metcalf, and J.Dalibard, 1985, Phys. Rev. Lett. 54, 992.

Raab, E. L., M. Prentiss, A. Cable, S. Chu, and D. E. Prit-chard, 1987, Phys. Rev. Lett. 59, 2631.

Roberts, J. L., N. R. Claussen, J. P. Burke, Jr., C. H. Greene, E.A. Cornell, and C. E. Wieman, 1998, Phys. Rev. Lett. 81,5109.

Roberts, J. L., N. R. Claussen, S. L. Cornish, E. A. Donley, E.A. Cornell, and C. E. Wieman, 2001, Phys. Rev. Lett. 86,4211.

Rohner, H., 1994, Proceedings of the 39th Symposium on theArt of Glassblowing (American Scientific Glassblowing Soci-ety, Wilmington, Delaware), p. 57.

Sesko, D., T. Walker, C. Monroe, A. Gallagher, and C. Wie-man, 1989, Phys. Rev. Lett. 63, 961.

Sesko, D., T. G. Walker, and C. E. Wieman, 1991, J. Opt. Soc.Am. B 8, 946.

Silvera, I. F., and J. T. M. Walraven, 1980, Phys. Rev. Lett. 44,164.

Page 19: Nobel Lecture: Bose-Einstein condensation in a dilute gas ...duine102/summerschoolTP/BECNobel.pdf · experimental Bose-Einstein condensation. For a somewhat more scholarly treatment

893E. A. Cornell and C. E. Wieman: BEC in a dilute gas

Sprik, R., J. T. M. Walraven, and I. F. Silvera, 1985, Phys. Rev.B 32, 5668.

Stoof, H. T. C., 2000, private communication.Stwalley, W. C., and L. H. Nosanow, 1976, Phys. Rev. Lett. 36,

910.Thorsheim, H. R., J. Weiner, and P. S. Julienne, 1987, Phys.

Rev. Lett. 58, 2420.Tiesinga, E., B. J. Verhaar, and H. T. C. Stoof, 1993, Phys. Rev.

A 47, 4114.Tisza, L., 1938, Nature (London) 141, 913.Vuletic, V., A. J. Kerman, C. Chin, and S. Chu, 1999, Phys.

Rev. Mod. Phys., Vol. 74, No. 3, July 2002

Rev. Lett. 82, 1406.Walker, T., D. Sesko, and C. Wieman, 1990, Phys. Rev. Lett. 64,

408.Watts, R. N., and C. E. Wieman, 1986, Opt. Lett. 11, 291.Williams, J. E., and M. J. Holland, 1999, Nature (London) 401,

568.Wineland, D. J., and H. Dehmelt, 1975, Bull. Am. Phys. Soc.

20, 637.Wineland, D. J., R. E. Drullinger, and F. L. Walls, 1978, Phys.

Rev. Lett. 40, 1639.

Page 20: Nobel Lecture: Bose-Einstein condensation in a dilute gas ...duine102/summerschoolTP/BECNobel.pdf · experimental Bose-Einstein condensation. For a somewhat more scholarly treatment

REVIEWS OF MODERN PHYSICS, VOLUME 74, OCTOBER 2002

Nobel lecture: When atoms behave as waves:Bose-Einstein condensation and the atom laser*

Wolfgang Ketterle†

Department of Physics, MIT-Harvard Center for Ultracold Atoms,and Research Laboratory of Electronics, Massachusetts Institute of Technology,Cambridge, Massachusetts 02139

(Published 20 November 2002)

I. INTRODUCTION

The lure of lower temperatures has attracted physi-cists for the past century, and with each advance towardsabsolute zero, new and rich physics has emerged. Lay-people may wonder why ‘‘freezing cold’’ is not coldenough. But imagine how many aspects of nature wewould miss if we lived on the surface of the sun. Withoutinventing refrigerators, we would only know gaseousmatter and never observe liquids or solids, and miss thebeauty of snowflakes. Cooling to normal earthly tem-peratures reveals these dramatically different states ofmatter, but this is only the beginning: many more statesappear with further cooling. The approach into thekelvin range was rewarded with the discovery of super-conductivity in 1911 and of superfluidity in helium-4 in1938. Cooling into the millikelvin regime revealed thesuperfluidity of helium-3 in 1972. The advent of lasercooling in the 1980s opened up a new approach toultralow-temperature physics. Microkelvin samples ofdilute atom clouds were generated and used for preci-sion measurements and studies of ultracold collisions.Nanokelvin temperatures were necessary to explorequantum-degenerate gases, such as Bose-Einstein con-densates first realized in 1995. Each of these achieve-ments in cooling has been a major advance, and recog-nized with a Nobel prize.

This paper describes the discovery and study of Bose-Einstein condensates (BEC’s) in atomic gases from mypersonal perspective. Since 1995, this field has grownexplosively, drawing researchers from the communitiesof atomic physics, quantum optics, and condensed-matter physics. The trapped ultracold vapor hasemerged as a new quantum system that is unique in theprecision and flexibility with which it can be controlledand manipulated. At least 30 groups have now createdcondensates, and the publication rate on Bose-Einsteincondensation has soared following the discovery of thegaseous condensates in 1995 (see Fig. 1).

*The 2001 Nobel Prize in Physics was shared by E. A. Cor-nell, Wolfgang Ketterle, and E. Wieman. This lecture is thetext of Professor Ketterle’s address on the occasion of theaward. The lecture of Professors Cornell and Wieman is re-printed in the July 2002 issue of Reviews of Modern Physics.

†URL: http://cua.mit.edu/ketterle–group/

0034-6861/2002/74(4)/1131(21)/$35.00 1131

The phenomenon of Bose-Einstein condensation waspredicted long ago, in a 1925 paper by Albert Einstein(Einstein, 1925b) using a method introduced by Satyen-dra Nath Bose to derive the black-body spectrum (Bose,1924). When a gas of bosonic atoms is cooled below acritical temperature Tc , a large fraction of the atomscondenses in the lowest quantum state. Atoms at tem-perature T and with mass m can be regarded asquantum-mechanical wave packets that have a spatialextent on the order of a thermal de Broglie wavelengthldB5(2p\2/mkBT)1/2. The value of ldB is the positionuncertainty associated with the thermal momentum dis-tribution and increases with decreasing temperature.When atoms are cooled to the point where ldB is com-parable to the interatomic separation, the atomic wavepackets ‘‘overlap’’ and the gas starts to become a ‘‘quan-tum soup’’ of indistinguishable particles. Bosonic atomsundergo a quantum-mechanical phase transition andform a Bose-Einstein condensate (Fig. 2), a cloud of at-oms all occupying the same quantum-mechanical state ata precise temperature (which, for an ideal gas, is relatedto the peak atomic density n by nldB

3 52.612). If theatoms are fermions, cooling gradually brings the gascloser to being a ‘‘Fermi sea’’ in which exactly one atomoccupies each low-energy state.

Creating a BEC is thus simple in principle: make a gasextremely cold until the atomic wave packets start tooverlap! However, in most cases quantum degeneracywould simply be preempted by the more familiar transi-tions to a liquid or solid. This more conventional con-densation into a liquid and solid can only be avoided atextremely low densities, about a hundred-thousandth

FIG. 1. Annual number of published papers which have thewords ‘‘Bose’’ and ‘‘Einstein’’ in their title, abstracts, or key-words. The data were obtained by searching the ISI (Institutefor Scientific Information) database.

© The Nobel Foundation 2001

Page 21: Nobel Lecture: Bose-Einstein condensation in a dilute gas ...duine102/summerschoolTP/BECNobel.pdf · experimental Bose-Einstein condensation. For a somewhat more scholarly treatment

1132 Wolfgang Ketterle: When atoms behave as waves

the density of normal air. Under those conditions, theformation time of molecules or clusters by three-bodycollisions (which is proportional to the inverse densitysquared) is stretched to seconds or minutes. Since therate of binary elastic collisions drops only proportionalto the density, these collisions are much more frequent.Therefore thermal equilibrium of the translational de-gree of freedom of the atomic gas is reached much fasterthan chemical equilibrium, and quantum degeneracy canbe achieved in an effectively metastable gas phase. How-ever, such ultralow density lowers the temperature re-quirement for quantum degeneracy into the nanokelvinto microkelvin range.

The achievement of Bose-Einstein condensation re-quired first the identification of an atomic system whichwould stay gaseous all the way to the BEC transition,and second, the development of cooling and trappingtechniques to reach the required regime of temperatureand density. Even around 1990, it was not certain thatNature would provide us with such a system. Indeed,many people doubted that BEC could ever be achieved,and it was regarded as an elusive goal. Many believedthat pursuing BEC would result in new and interestingphysics, but whenever one would come close, some newphenomenon or technical limitation would show up. Anews article in 1994 quoted Steve Chu: ‘‘I am betting onNature to hide Bose condensation from us. The last 15years she’s been doing a great job’’ (Taubes, 1994).

In brief, the conditions for BEC in alkali gases arereached by combining two cooling methods. Laser cool-

FIG. 2. Criterion for Bose-Einstein condensation. At hightemperatures, a weakly interacting gas can be treated as a sys-tem of ‘‘billiard balls.’’ In a simplified quantum description, theatoms can be regarded as wave packets with an extension oftheir de Broglie wavelength ldB . At the BEC transition tem-perature, ldB becomes comparable to the distance betweenatoms, and a Bose condensate forms. As the temperature ap-proaches zero, the thermal cloud disappears, leaving a pureBose condensate.

Rev. Mod. Phys., Vol. 74, No. 4, October 2002

ing is used to precool the gas. The principle of lasercooling is that scattered photons are on average blue-shifted with respect to the incident laser beam. As aresult, the scattered light carries away more energy thanhas been absorbed by the atoms, resulting in net cooling.Blueshifts are caused by Doppler shifts or ac Starkshifts. The different laser cooling schemes are describedin the 1997 Nobel lectures in physics (Chu, 1998; Cohen-Tannoudji, 1998; Phillips, 1998). After the precooling,the atoms are cold enough to be confined in a magnetictrap. Wall-free confinement is necessary, otherwise theatoms would stick to the surface of the container. It isnoteworthy that similar magnetic confinement is alsoused for plasmas which are too hot for any material con-tainer. After magnetically trapping the atoms, forcedevaporative cooling is applied as the second coolingstage (Masuhara et al., 1988; Ketterle and van Druten,1996; Walraven, 1996). In this scheme, the trap depth isreduced, allowing the most energetic atoms to escapewhile the remainder rethermalize at steadily lower tem-peratures. Most BEC experiments reach quantum de-generacy between 500 nK and 2 mK, at densities be-tween 1014 and 1015 cm23. The largest condensates areof 100 million atoms for sodium, and a billion for hydro-gen; the smallest are just a few hundred atoms. Depend-ing on the magnetic trap, the shape of the condensate iseither approximately round, with a diameter of 10–50mm, or cigar-shaped with about 15 mm in diameter and300 mm in length. The full cooling cycle that produces acondensate may take from a few seconds to as long asseveral minutes.

After this short overview, I want to provide the his-torical context for the search for BEC and then describethe developments which led to the observation of BECin sodium at MIT. Finally, some examples will illustratethe novel physics which has been explored using Bose-Einstein condensates. A more detailed account of thework of my group has been presented in four compre-hensive review papers (Ketterle and van Druten, 1996;Ketterle et al., 1999; Ketterle and Inouye, 2001;Stamper-Kurn and Ketterle, 2001).

II. BEC AND CONDENSED-MATTER PHYSICS

Bose-Einstein condensation is one of the most intrigu-ing phenomena predicted by quantum statistical me-chanics. The history of the theory of BEC is very inter-esting, and is nicely described in the biographies ofEinstein (Pais, 1982) and London (Gavroglu, 1995) andreviewed by Griffin (1999). For instance, Einstein madehis predictions before quantum theory had been fullydeveloped, and before the differences between bosonsand fermions had been revealed (Einstein, 1925a). AfterEinstein, important contributions were made by, mostnotably, London, Landau, Tisza, Bogoliubov, Penrose,Onsager, Feynman, Lee, Yang, Huang, Beliaev, and Pi-taevskii. An important issue has always been the rela-tionship between BEC and superfluidity in liquid he-lium, an issue that was highly controversial betweenLondon and Landau (see Gavroglu, 1995). Works byBogoliubov, Beliaev, Griffin, and others showed that

Page 22: Nobel Lecture: Bose-Einstein condensation in a dilute gas ...duine102/summerschoolTP/BECNobel.pdf · experimental Bose-Einstein condensation. For a somewhat more scholarly treatment

1133Wolfgang Ketterle: When atoms behave as waves

Bose-Einstein condensation gives the microscopic pic-ture behind Landau’s ‘‘quantum hydrodynamics.’’ BECis closely related to superconductivity, which can be de-scribed as being due to Bose-Einstein condensation ofCooper pairs. Thus Bose-Einstein condensation is at theheart of several macroscopic quantum phenomena.

BEC is unique in that it is a purely quantum-statisticalphase transition, i.e., it occurs even in the absence ofinteractions. Einstein (1925a) described the transition ascondensation ‘‘without attractive forces.’’ This makesBEC an important paradigm of statistical mechanics,which has been discussed in a variety of contexts incondensed-matter, nuclear, particle, and astrophysics(Griffin et al., 1995). On the other hand, real-life par-ticles will always interact, and even the weakly interact-ing Bose gas behaves qualitatively differently from theideal Bose gas (Huang, 1987). It was believed for quitesome time that interactions would always lead to ‘‘ordi-nary’’ condensation (into a solid) before Bose-Einsteincondensation would happen. Liquid helium was the onlycounterexample, where the light mass and concomitantlarge zero-point kinetic energy prevents solidificationeven at zero kelvin. Erwin Schrodinger wrote in 1952 ina textbook on thermodynamics about BEC: ‘‘The densi-ties are so high and the temperatures so low—those re-quired to exhibit a noticeable departure [from classicalstatistics]—that the van der Waals corrections are boundto coalesce with the possible effects of degeneration, andthere is little prospect of ever being able to separate thetwo kinds of effect’’ (Schrodinger, 1952). What he didn’tconsider were dilute systems in a metastable gaseousphase!

The quest to realize BEC in a dilute weakly interact-ing gas was pursued in at least three different directions:liquid helium, excitons, and atomic gases. Experimental(Crooker et al., 1983; Reppy, 1984) and theoretical work(Rasolt et al., 1984) showed that the onset of superfluid-ity for liquid helium in Vycor has features of dilute-gasBose-Einstein condensation. At sufficiently low cover-age, the helium adsorbed on the porous spongelike glassbehaved like a dilute three-dimensional gas. However,the interpretation of these results is not unambiguous(Cho and Williams, 1995).

Excitons, which consist of weakly bound electron-holepairs, are composite bosons. The physics of excitons insemiconductors is very rich and includes the formationof an electron-hole liquid and biexcitons. As nicely dis-cussed by Wolfe et al. (1995) and Fortin et al. (1995),there are systems where excitons form a weakly interact-ing gas. However, the initial evidence for Bose-Einsteincondensation in Cu2O (Lin and Wolfe, 1993) was re-tracted (O’Hara et al., 1999). Recent work in coupledquantum-well structures is very promising (Butov et al.,2002). When excitons strongly interact with light in acavity, they form polaritons. In such polariton systems,stimulated scattering and nonequilibrium condensateshave been observed recently (Yamamoto, 2000; Sabaet al., 2001; Baumberg, 2002).

Rev. Mod. Phys., Vol. 74, No. 4, October 2002

III. SPIN-POLARIZED HYDROGEN

Dilute atomic gases are distinguished from thecondensed-matter systems discussed above by the ab-sence of strong interactions. Interactions at the densityof a liquid or a solid considerably modify and complicatethe nature of the phase transition. Hecht (1959) andStwalley and Nosanow (1976) used the quantum theoryof corresponding states to conclude that spin-polarizedhydrogen would remain gaseous down to zero tempera-ture and should be a good candidate to realize Bose-Einstein condensation in a dilute atomic gas. These sug-gestions triggered several experimental efforts, mostnotably by Silvera and Walraven in Amsterdam, byGreytak and Kleppner at MIT, and by others at Moscow,Turku, British Columbia, Cornell, Harvard, and Kyoto.The stabilization of a spin-polarized hydrogen gas (Clineet al., 1980; Silvera and Walraven, 1980) created greatexcitement about the prospects of exploring quantum-degenerate gases. Experiments were first done by fillingcryogenic cells with the spin-polarized gas, and by com-pressing it, and since 1985, by magnetic trapping andevaporative cooling. BEC was finally accomplished in1998 by Kleppner, Greytak and collaborators (Friedet al., 1998). See Greytak and Kleppner (1984), Silveraand Walraven (1986), Greytak (1995), and Walraven(1996) and in particular Kleppner et al. (1999) for a fullaccount of the pursuit of Bose-Einstein condensation inatomic hydrogen. Evidence for a phase transition in twodimensions was reported in 1998 (Safonov et al., 1998).

The work in alkali atoms is based on the work in spin-polarized hydrogen in several respects:

• Studies of spin-polarized hydrogen showed that sys-tems can remain in a metastable gaseous state close toBEC conditions. The challenge was then to find thewindow in density and temperature where this meta-stability is sufficient to realize BEC.

• Many aspects of BEC in an inhomogeneous potential(Goldman et al., 1981; Huse and Siggia, 1982; Oliva,1989), and the theory of cold collision processes (see,for example, Stoof et al., 1988) developed in the 1980sfor hydrogen could be applied directly to the alkalisystems.

• The technique of evaporative cooling was developedfirst for hydrogen (Hess, 1986; Masuhara et al., 1988)and then used for alkali atoms.

IV. LASER COOLING

Laser cooling opened a new route to ultralow tem-perature physics. Laser cooling experiments, with room-temperature vacuum chambers and easy optical access,look very different from cryogenic cells with multilayerthermal shielding around them. Also, the number ofatomic species that can be studied at ultralow tempera-tures was greatly extended from helium and hydrogen toall of the alkali atoms, metastable rare gases, severalearth-alkali atoms, and others (the list of laser-cooled

Page 23: Nobel Lecture: Bose-Einstein condensation in a dilute gas ...duine102/summerschoolTP/BECNobel.pdf · experimental Bose-Einstein condensation. For a somewhat more scholarly treatment

1134 Wolfgang Ketterle: When atoms behave as waves

atomic species is still growing). A full account of therelevant laser cooling techniques and their developmentis given by Arimondo et al. (1992), Metcalf and van derStraten, (1994), and Adams and Riis (1997) and in the1997 Nobel lectures of Chu (1998), Cohen-Tannoudji(1998), and Phillips (1998).

Some papers and proposals written in the early andmid 1980s, before and during the developments of thebasic cooling and trapping techniques, listed quantumdegeneracy in a gas as a visionary goal for this newemerging field (Letokhov and Minogin, 1980; Chu et al.,1985; Pritchard, 1986). However, major limitations of la-ser cooling and trapping were soon identified. Althoughthere is no fundamental low-temperature limit, the finaltemperature provided by polarization gradient cooling—about ten times the recoil energy—was regarded as apractical limit. Subrecoil laser cooling techniques, espe-cially in three dimensions, were harder to implement,and required long cooling times. The number and den-sity of atoms were limited by inelastic, light-induced col-lisions (leading to trap loss; see Walker and Feng, 1994;and Weiner, 1995) and by absorption of scattered laserlight (Walker et al., 1990), which results in an outwardradiation pressure (weakening the trapping potentialand limiting the density). Furthermore, since the lowesttemperatures could not be achieved at the highest den-sities (Drewsen et al., 1994; Townsend et al., 1995; 1996),most trapping and cooling techniques reached a maxi-mum phase-space density nldB

3 51025; a value of 2.612is needed for BEC. This was the situation when the au-thor joined the field of cold atoms in 1990. It was onlymore recently that major increases in phase-space den-sity were achieved by laser cooling (DePue et al., 1999;Ido et al., 2000; Kerman et al., 2000) but so far lasercooling by itself has not been able to reach BEC.

V. THE EFFORT AT MIT 1990–1996

A. Improving laser cooling

When I teamed up with Dave Pritchard at MIT in1990 as a postdoc, the initial goal was to build an intensesource of cold atoms to study cold collisions and purelong-range molecules. However, Dave and I frequentlytalked about the limitations in density and temperatureof the current techniques and tried to develop ideas onhow to get around them. One limitation of magnetictraps is that they can hold atoms only in weak-field-seeking hyperfine states. Therefore a collision betweentwo trapped atoms can lead to a spin flip, and the Zee-man energy is converted into kinetic energy (dipolar re-laxation). This process has been a major limitation tothe experiments in atomic hydrogen.

First, we asked ourselves if the inclusion of electricand gravitational fields would allow the stable confine-ment of atoms in their lowest hyperfine states—but theanswer was negative (Ketterle and Pritchard, 1992a).One loophole was time-dependent magnetic fields, andbuilding on an earlier proposal (Lovelace et al., 1985), Idesigned an experiment to confine sodium atoms with ac

Rev. Mod. Phys., Vol. 74, No. 4, October 2002

magnetic fields which looked feasible. However, welearned that Eric Cornell at Boulder had developed asimilar idea and experimentally implemented it (Cornellet al., 1991)—so we left the idea on the drawing board. Itwasn’t the last time that Eric and I would develop simi-lar ideas independently and almost simultaneously!

Trapping atoms in the lowest hyperfine state was notnecessary to accomplish BEC. Already in 1986, Prit-chard correctly estimated the rate constants of elasticand inelastic collisions for alkali atoms (Pritchard, 1986).From these estimates one could easily predict that foralkali atoms, in contrast to hydrogen, the so-called goodcollisions (elastic collisions necessary for the evapora-tion process) would clearly dominate over the so-calledbad collisions (inelastic two- and three-body collisions);therefore evaporative cooling in alkalis would probablynot be limited by intrinsic loss and heating processes.However, there was pessimism (Vigue, 1986) and skep-ticism, and the above-mentioned experimental (Cornellet al., 1991) and theoretical (Ketterle and Pritchard,1992a) work on traps for strong-field-seeking atoms hasto be seen in this context.

In those years, there were some suggestions that time-dependent potentials could lead to substantial cooling,but we showed that this was not possible (Ketterle andPritchard, 1992b). Real cooling needs an open systemwhich allows entropy to be removed from thesystem—in laser cooling in the form of scattered pho-tons, in evaporative cooling in the form of discarded at-oms. Dave and I brainstormed about novel laser coolingschemes. In 1991, at the Varenna summer school, Davepresented a new three-level cooling scheme (Pritchardand Ketterle, 1992). Inspired by these ideas, I developeda scheme using Raman transitions. Replacing the six la-ser beams in optical molasses by counterpropagatingbeams driving the Doppler-sensitive Raman transition,we hoped to realize Doppler molasses with a linewidththat was proportional to the optical pumping rate andtherefore adjustable. We had started setting up radiofrequency (rf) electronics and magnetic shields for Ra-man cooling when we heard that Mark Kasevich andSteve Chu were working on Raman cooling using laserpulses (Kasevich and Chu, 1992). For this reason, andalso because around the same time we had developedthe idea for the Dark SPOT (spontaneous force opticaltrap; see later in this section), we stopped our work onRaman cooling.

Our experimental work in those years focused first ongenerating a large flux of slow atoms. In my first monthsat MIT, when I overlapped with Kris Helmerson andMin Xiao, we built a sodium vapor cell magneto-opticaltrap (MOT). The idea was inspired by the Boulder ex-periment (Monroe et al., 1990), and our hope was tovastly increase the loading rate by additional frequenciesor frequency chirps added to the red side of the D2 reso-nance line. The idea failed—we first suspected thatnearby hyperfine levels of sodium may have adverselyinterfered, but it was later shown that it didn’t work forcesium either (Lindquist et al., 1992) because of the un-favorable duty cycle of the chirp. Still, except for a cryo-genic setup which was soon abandoned, it was the first

Page 24: Nobel Lecture: Bose-Einstein condensation in a dilute gas ...duine102/summerschoolTP/BECNobel.pdf · experimental Bose-Einstein condensation. For a somewhat more scholarly treatment

1135Wolfgang Ketterle: When atoms behave as waves

magneto-optical trap built at MIT (Dave Pritchard’s ear-lier work on magneto-optical trapping was carried out atBell Labs in collaboration with Steve Chu’s group). We(Michael Joffe, Alex Martin, Dave Pritchard and my-self) then put our efforts on beam slowing, and got dis-tracted from pursuing Zeeman slowing by the idea ofisotropic light slowing (Ketterle, Martin et al., 1992). Inthis scheme, atoms are sent through a cavity with dif-fusely reflecting walls and exposed to an isotropic lightfield. For red-detuned light the atoms preferentially ab-sorb light from a forward direction and are slowed. Theexperiment worked very well and it was a lot of fun todo. However, the requirements for laser power and thevelocity capture range of this method were inferior toZeeman slowing, so we decided to build an optimizedZeeman slower.

We adopted the new design by Greg Lafyatis in whichthe magnetic field increases rather than decreases as in aconventional Zeeman slower (Barrett et al., 1991). Werealized that at the magnetic-field maximum it would bepossible to apply some additional transverse laser cool-ing to collimate the slow beam. Michael Joffe, a gradu-ate student, wound a solenoid which had radial accessfor four extra laser beams. The collimation worked(Joffe et al., 1993), but not as well as we had hoped, andwe felt that the small gain was not worth the added com-plexity. Still, even without collimation, our Zeemanslower provided one of the largest slow-atom fluxes re-ported until then, and soon after we had a magneto-optical trap with a large cloud of sodium atoms. In hind-sight, I am amazed at how many different schemes weconsidered and tried out, but this may have been neces-sary to distill the best approach.

The 1991 Varenna summer school on laser coolingwas memorable to me for several reasons. I had joinedthe field of cold atoms just a year earlier, and there I metmany colleagues for the first time and established long-lasting relationships. I still have vivid memories of onelong afternoon when Dave Pritchard and I sat outsidethe meeting place, which offered a spectacular view ofLake Como, and brainstormed about the big goals ofour field and how to approach them. Dave’s encourage-ment was crucial to me and helped to increase my self-confidence in my new field of research. We consideredoptions and strategies on how to combine laser coolingand evaporative cooling, something which had been onour mind for some time.

Following the example of the spin-polarized hydrogenexperiment at MIT (Masuhara et al., 1988), evaporationcould be done in a magnetic trap using rf-induced spinflips, as suggested by Pritchard and collaborators (Prit-chard, Helmerson, and Martin, 1989). Magnetic trapsand laser cooling had already been used simultaneouslyin the first experiments on magnetic trapping at NIST(Migdall et al., 1985) and MIT (Bagnato et al., 1987),and on Doppler cooling of magnetically trapped atomsat MIT (Prichard et al., 1989; Helmerson et al., 1992). In1990, a magnetic trap was loaded from a magneto-optical trap and optical molasses in Boulder (Monroeet al., 1990). The laser cooling route to BEC was sum-

Rev. Mod. Phys., Vol. 74, No. 4, October 2002

marized by Monroe, Cornell, and Wieman (1992). Somost of the pieces to get to BEC were known in 1990,but there was doubt about whether they would fit to-gether.

Laser cooling works best at low densities where lightabsorption and light-induced collisions are avoided,whereas evaporative cooling requires a high collisionrate and high density. The problem is the much highercross section for light scattering of ;1029 cm2, while thecross section for elastic scattering of atoms is a thousandtimes smaller. In hindsight, it would have been sufficientto provide tight magnetic compression after laser cool-ing and an extremely good vacuum to obtain a lifetimeof the sample that is much longer than the time betweencollisions, as demonstrated at Rice University (Bradleyet al., 1995). However, our assessment was that one ma-jor improvement had to be done to laser cooling tobridge the gap in density between the two coolingschemes. Dave and I discussed possibilities on how tocircumvent the density-limiting processes in magneto-optical traps. We considered coherent population trap-ping schemes in which atoms are put into a coherentsuperposition state which does not absorb the light. Wedeveloped some ideas on how atoms near the center ofthe trap would be pumped into such a dark state, but thenumbers were not too promising. A few months later, asimple idea emerged. If the so-called repumping beamof the magneto-optical trap would have a shadow in thecenter, atoms would stay there in the lower hyperfinestate and not absorb the trapping light, which is nearresonant for atoms in the upper hyperfine state. In aMOT, the density is limited by losses due to excited-state collisions and by multiple scattering of light, whichresults in an effective repulsive force between atoms.When atoms are kept in the dark, the trapping forcedecreases by a factor which is proportional to the prob-ability of the atoms to be in the resonant hyperfine state.However, the repulsive force requires both atoms to beresonant with the light and decreases with the square ofthis factor. Therefore there is net gain in confinement bykeeping atoms in the dark. Of course, there is a limit tohow far you can push this concept, which is reachedwhen the size of the cloud is no longer determined bythe balance of trapping and repulsive forces, but by thefinite temperature of the cloud.

The gain in density of this scheme, called Dark SPOT,over the standard MOT is bigger when the number oftrapped atoms is large. So in 1992, we tweaked up theMOT to a huge size before we implemented the idea. Itworked almost immediately, and we got very excitedabout the dark shadows cast by the trapped atoms whenthey were illuminated by a probe beam. We inferred thatthe probe light had been attenuated by a factor of morethan e2100 (Ketterle et al., 1993a). This implied that wehad created a cloud of cold atoms with an unprec-edented combination of number and density.

B. Combining laser cooling and evaporative cooling

The following weeks and months were quite dramatic.What should we do next? Dave Pritchard had planned

Page 25: Nobel Lecture: Bose-Einstein condensation in a dilute gas ...duine102/summerschoolTP/BECNobel.pdf · experimental Bose-Einstein condensation. For a somewhat more scholarly treatment

1136 Wolfgang Ketterle: When atoms behave as waves

to use this trap as an excellent starting point for thestudy of cold collisions and photoassociation—and in-deed other groups had major successes along these lines(Heinzen, 1999; Weiner et al., 1999). But there was alsothe exciting prospect of combining laser cooling withevaporative cooling. We estimated the elastic collisionrate in the Dark SPOT trap to be around 100 Hz (Ket-terle et al., 1993a) which appeared to be more than suf-ficient to start runaway evaporation in a magnetic trap.After some discussions, the whole group decided to gofor the more ambitious and speculative goal of evapora-tive cooling. It was one of those rare moments wheresuddenly the whole group’s effort gets refocused. Evenbefore we wrote the paper on the Dark SPOT trap, weplaced orders for essential components to upgrade ourexperiment to ultrahigh vacuum and to magnetic trap-ping. All resources of the lab were now directed towardsthe evaporative cooling of sodium. The Dark SPOT trapwas a huge improvement towards combining high atomnumber and high density in laser cooling. It turned outto be crucial to the BEC work both at Boulder (Ander-son et al., 1995) and at MIT (Davis, Mewes, Andrews,et al., 1995) and seems to be still necessary in all currentBEC experiments with sodium, but not for rubidium.

The next step was the design of a tightly confiningmagnetic trap. We decided to use the spherical quadru-pole trap, which simply consists of two opposing coils—this design was used in the first demonstration of mag-netic trapping (Migdall et al., 1985). We knew that thistrap would ultimately be limited by Majorana flops inthe center of the trap where the magnetic field is zero.Near zero magnetic field, the atomic spin doesn’t precessfast enough to follow the changing direction of the mag-netic field—the result is a transition to another Zeemansublevel which is untrapped, leading to trap loss. Weestimated the Majorana flop rate, but there was someuncertainty about the numerical prefactor. Still, itseemed that Majorana flops would only become criticalafter the cloud had shrunk due to evaporative cooling,so they shouldn’t get in the way of demonstrating thecombination of laser cooling and evaporative cooling.After Michael Joffe presented our approach with thequadrupole trap at the QELS meeting in 1993, Eric Cor-nell informed me that he had independently arrived atthe same conclusion. In 1993, my group reported at theOSA meeting in Toronto the transfer of atoms from theDark SPOT trap into a magnetic trap, and the effects oftruncation of the cloud using rf induced spinflips (Ket-terle et al., 1993b).

At about this time, I joined the MIT faculty as assis-tant professor. Dave Pritchard made the unprecedentedoffer that if I stayed at MIT he would hand over to methe existing lab, including two grants. To make sure thatI would receive the full credit for the work towardsBEC, he decided not to stay involved in a field he hadpioneered and gave me full responsibility and indepen-dence. Dave told me that he wanted to focus on hisother two experiments, the single-ion mass measure-ment and the atom interferometry, although what hegave up was his ‘‘hottest’’ research activity. Even now, I

Rev. Mod. Phys., Vol. 74, No. 4, October 2002

am moved by his generosity and unusual mentorship.The two graduate students on the project, Ken Davisand Marc-Oliver Mewes, who had started their Ph.D.’sin 1991 and 1992, respectively, deliberated whether theyshould stay with Dave Pritchard and work on one of hisother experiments, or continue their work on BEC in anewly formed group headed by a largely unknown assis-tant professor. They both opted for the latter and wecould pursue our efforts without delay, along withMichael Andrews, who joined the group in the summerof 1993.

For a few months we got distracted from our goal ofevaporative cooling. Our optical molasses temperatureswere higher than those reported by the NIST group(Lett et al., 1989), and we felt that we had to learn thestate of the art before we could advance to even lowertemperatures. We suspected that the higher density ofatoms played a role, but we had to improve our tech-nique of temperature measurements. Our goal was tocharacterize the interplay of parameters in ‘‘dark’’ mo-lasses where most of the atoms are pumped into thedark hyperfine state. It was also a good project for thegraduate students to hone their skills and develop inde-pendence. After a few months we had made someprogress, but I became concerned about the delay andthe competition from Boulder. We decided to drop theproject and resume our work on evaporative cooling. Upto the present day, we have never implemented accuratediagnostics for the temperature obtained in lasercooling—it was just not important.

In the spring of 1994, we saw first evidence for anincrease in phase-space density by evaporative cooling.We reported these results at an invited talk at the Inter-national Quantum Electronics Conference (IQEC) inMay 1994. At the same meeting, the Boulder group re-ported similar results and the limitations due to the Ma-jorana flops as the temperature was reduced. It was clearthat the next step was an improvement of the magnetictrap, to trap atoms at a finite bias field which wouldsuppress the Majorana flops. During the meeting, I cameup with the idea of plugging the hole with a focusedlaser beam: a blue-detuned laser beam focused onto thezero-magnetic-field point would exert repulsive dipoleforces onto the atoms and keep them away from thisregion (Fig. 3). This idea seemed so obvious to me that Iexpected the Boulder group to come up with somethingsimilar. It was only at the next conference (ICAP 1994)in Boulder (Davis et al., 1994), when I presented ourapproach, that I learned about Eric Cornell’s idea ofsuppressing Majorana flops with a rapidly rotating mag-netic field—the so-called TOP trap (Petrich et al., 1995).However, we didn’t implement the optical plug immedi-ately. We wanted first to document our observation ofevaporative cooling. We realized that our fluorescencediagnostics were inadequate and implemented absorp-tion imaging which is now the standard technique forobserving Bose-Einstein condensation. In those days, wefocused on direct imaging of the trapped cloud (withoutballistic expansion), and Michael Andrews and Marc-Oliver Mewes developed a sophisticated computer codeto simulate absorption images in inhomogeneous mag-

Page 26: Nobel Lecture: Bose-Einstein condensation in a dilute gas ...duine102/summerschoolTP/BECNobel.pdf · experimental Bose-Einstein condensation. For a somewhat more scholarly treatment

1137Wolfgang Ketterle: When atoms behave as waves

netic fields. We thought that this would be a useful tool,but we rapidly advanced to much lower temperatureswhere the inhomogeneous Zeeman shifts were smallerthan the linewidth, and never needed the code againafter our first paper on evaporative cooling (Davis,Mewes, Joffe, et al., 1995).

In late 1994, we had a ‘‘core meltdown.’’ The magnetictrap was switched on without cooling water, and the sil-ver solder joints of the coils melted. Since in those daysthe magnetic coils were mounted inside the vacuumchamber, we had a catastrophic loss of vacuum and ma-jor parts of our setup had to be disassembled. I willnever forget the sight of coils dripping with water behinda UHV viewport. This happened just a few hours beforeMIT’s president, Charles Vest, visited our lab to get first-hand information on some of the research done on cam-pus. He still remembers this event. We had lost weeks ormonths of work in a very competitive situation. I wasdespondent and suggested to the group that we go outfor a beer and then figure out what to do, but the stu-dents immediately pulled out the wrenches and startedthe repair. I was moved to see their dedication andstrength, even at this difficult time. We replaced themagnetic trap by a much sturdier one. This turned out tobe crucial for the implementation of the plugged trapwhere the precise alignment of a laser beam relative tothe magnetic field center was important. So in hindsightthe disaster may not have caused a major delay.

FIG. 3. Experimental setup for cooling atoms to Bose-Einsteincondensation. Sodium atoms are trapped by a strong magneticfield, generated by two coils. In the center, the magnetic fieldvanishes, which allows the atoms to spin flip and escape.Therefore the atoms are kept away from the center of the trapby a strong (3.5-W) argon ion laser beam (‘‘optical plug’’),which exerts a repulsive force on the atoms. Evaporative cool-ing is controlled by radio-frequency radiation from an antenna.The rf selectively flips the spins of the most energetic atoms.The remaining atoms rethermalize (at a lower temperature) bycollisions among themselves. Evaporative cooling is forced bylowering the rf frequency.

Rev. Mod. Phys., Vol. 74, No. 4, October 2002

In early 1995, I had to tell my three graduate studentsthat we were rapidly using up startup money and ur-gently needed one of our two pending proposals ap-proved. Otherwise we would not be able to continuespending money in the way we had done until then andwould slow down. Fortunately, in April 1995, the NSFinformed me that my proposal was funded. It is interest-ing to look at some of the reviewers comments now,seven years later: ‘‘It seems that vast improvements arerequired [in order to reach BEC] . . . the current tech-niques are so far from striking range for BEC that it isnot yet possible to make . . . an assessment . . . ’’; ‘‘The sci-entific payoffs, other than the importance of producing aBEC itself, are unclear.’’ And a third reviewer: ‘‘ . . . therehave been few specific (or realistic) proposals of inter-esting experiments that could be done with a conden-sate.’’ Despite the skepticism, all reviewers concludedthat the proposed ‘‘experiments are valuable and worthpursuing.’’ After we received the funding decision, thewhole group celebrated with dinner, and a fourth gradu-ate student (Dallin Durfee), who had expressed his in-terest already months earlier, could finally be supported.

In late December 1994, our paper on evaporativecooling was submitted, and we were free to focus onplugging the hole. We had to learn how to align a pow-erful argon ion laser beam and image it through manyattenuators without major distortions. When the plugwas aligned, the result was spectacular (Fig. 4). We couldimmediately cool down to lower temperatures and keepmany more atoms. During evaporation, the cloud be-came so cold and small that we couldn’t resolve it anymore. The highest phase-space density measured was afactor of 30 below BEC, but we may have been evencloser. We had only a few runs of the experiment beforewe ran into severe vacuum problems. We focused ini-tially on spatial imaging and became limited by resolu-tion, whereas ballistic expansion and time-of-flight imag-ing would not have suffered from this limitation. We alsothought that BEC would be accomplished at lower den-sities and in larger clouds, so we worked on adiabaticdecompression and ran into problems with the zero ofthe magnetic field moving away from the plug.

In those months, we were plagued by vacuum prob-lems. The coils inside the vacuum showed some strangeoutgassing behavior and the vacuum slowly deterio-

FIG. 4. Absorption images of atom clouds trapped in the op-tically plugged trap. Cloud (a) is already colder than was at-tainable without the ‘‘plug’’ (Ar ion laser beam). Cloud (b)shows the breakup of the cloud into two ‘‘pockets’’ in the twominima of the potential. The size of cloud (c) reached theoptical resolution of the imaging system (<10 mm) still ab-sorbing 90% of the probe light. This sets an upper bound ontemperature (<10 mK) and a lower bound on density (531012 cm23).

Page 27: Nobel Lecture: Bose-Einstein condensation in a dilute gas ...duine102/summerschoolTP/BECNobel.pdf · experimental Bose-Einstein condensation. For a somewhat more scholarly treatment

1138 Wolfgang Ketterle: When atoms behave as waves

rated. We went through several bakeouts of theultrahigh-vacuum chamber in the spring and summer of1995. Furthermore, Ken Davis had to write his Ph.D.thesis and stopped working in the lab. It is interesting torecall my assessment of the field in those months; Ididn’t realize that BEC was just around the corner. InTom Greytak’s and Dan Kleppner’s group the BEC tran-sition was approached to within a factor of 3.5 in tem-perature in 1991 (Doyle et al., 1991), but it took severalmore years to advance further. So I prepared for a longhaul to cover the last order of magnitude to BEC.

By this time, the group was reinforced by Dan Kurn(now Dan Stamper-Kurn), a graduate student, andKlaasjan van Druten, my first postdoc. After months ofworking on vacuum and other problems, we were justready to run the machine again when we heard aboutthe breakthrough in Boulder in June of 1995 (Andersonet al., 1995). We feverishly made several attempts withtraps plugged by focused laser beams and light sheets,and tried different strategies of evaporation without suc-cess. The clouds disappeared when they were very cold.We conjectured that some jitter of the laser beam wasresponsible, and when accelerometers indicated vibra-tions of our vacuum chambers, we immediately decidedto eliminate all turbo and mechanical pumps. Unfortu-nately, when we were exchanging the turbo pump on ouroven chamber against an ion pump, we caused a leak inthe ultrahigh-vacuum part and had to go through an-other long bakeout. We also implemented a pointing sta-bilization for the optical plug beam. But when we finallyobtained BEC, we realized that it didn’t improve thecooling.

These were difficult months for me. The Rice grouphad cooled lithium to quantum degeneracy (Bradleyet al., 1995). A new subfield of atomic physics was open-ing up, and I was afraid that our approach with sodiumand the plugged trap would not be successful and wewould miss the excitement. I considered various strate-gies. Several people suggested that I adopt the successfulTOP trap used at Boulder. But I had already started tostudy several possible configurations for magnetic con-finement. I realized that a highly elongated Ioffe-Pritchard trap with adjustable bias field could provide agood confinement that was equivalent or superior to theTOP trap. Around August 1995, Dan Kurn worked outan optimized configuration, which was the cloverleafwinding pattern (Mewes et al., 1996a). I considered hav-ing the whole group work on this new approach, butseveral in my group wanted to give the plugged trap afew more attempts and at least characterize how far wecould approach BEC with our original approach. Fortu-nately, we followed that suggestion—it is always a goodidea to listen to your collaborators.

C. BEC in sodium

This was the situation on September 29, 1995, whenwe observed BEC in sodium for the first time. The goalof the run was to measure the lifetime of the trappedatoms and characterize possible heating processes. For

Rev. Mod. Phys., Vol. 74, No. 4, October 2002

our ultrahigh-vacuum pressure, rather slow evaporationshould have been most efficient, but we found out thatfaster evaporation worked much better. This was a clearsign for some other loss or heating process, e.g., due tofluctuations in the position of the plug. Around 11:30p.m., an entry in the lab book states that the lifetimemeasurements were not reliable, but they indicated life-times around ten seconds, enough to continue evapora-tion. Fifteen minutes later we saw some dark spots intime-of-flight absorption images, but they were quitedistorted since the optical plug beam, which we couldn’tswitch off, pushed atoms apart during the ballistic ex-pansion (Fig. 5). Still, the sudden appearance of darkspots meant groups of atoms with very small relativevelocity. For the next few hours, we characterized theappearance of those spots, but then decided that furtherprogress required an acousto-optical modulator toswitch off the optical plug. Between 4:00 and 5:30 in theearly morning, we installed optics and rf electronics andwere finally able to switch off the argon ion laser beamduring ballistic expansion. A few minutes later, we ob-served the bimodal distributions that are now the hall-mark of BEC. The lab book of this night captured theexcitement of the moment (Fig. 6).

Those first measurements were done by imaging theatoms in the lower hyperfine (F51) state. For the nextrun, which took place a few days later, we prepared op-tical pumping and imaging on the cycling F52 transi-tion, and obtained a much better signal-to-noise ratio inour images. The occurrence of BEC was very dramatic

FIG. 5. Time-of-flight absorption images of some of the firstcondensates produced at MIT in the night of September 29,1995. After the magnetic quadrupole trap was switched off, theatom cloud expanded ballistically. However, since the opticalplug (indicated by black circles) could not be turned off at thesame time, it distorted the expanding cloud. Still, as the tem-perature was lowered from top to bottom, a distinctly sharpshadow appeared marking the presence of a condensate.

Page 28: Nobel Lecture: Bose-Einstein condensation in a dilute gas ...duine102/summerschoolTP/BECNobel.pdf · experimental Bose-Einstein condensation. For a somewhat more scholarly treatment

1139Wolfgang Ketterle: When atoms behave as waves

(Fig. 7). Our animated rendering of the data obtained inthat run (done by Dallin Durfee) became well known(see Durfee and Ketterle, 1998). We had obtained con-densates with 500 000 atoms, 200 times more than inBoulder, with a cooling cycle (of only nine seconds) 40times shorter. Our paper was quickly written and sub-mitted only two weeks after the experiment (Davis,Meewes, Andrews, et al., 1995).

In my wildest dreams I had not assumed that the stepfrom evaporative cooling to BEC would be so fast. Fig-ure 8 shows how dramatic the progress was after laserand evaporative cooling were combined. Within lessthan two years, the number of alkali atoms in a singlequantum state was increased by about 12 orders ofmagnitude—a true singularity demonstrating that aphase transition was achieved!

FIG. 6. One page of the lab book during the night of Septem-ber 29, 1995, when BEC was first observed at MIT. The hand-writing is by Klaasjan van Druten. At 5:50 a.m., we had in-stalled a new acousto-optical modulator to switch off theoptical plug (Ar ion laser beam). Fifteen minutes later, we hadthe first definitive evidence for BEC in sodium.

Rev. Mod. Phys., Vol. 74, No. 4, October 2002

MIT with its long tradition in atomic physics was aspecial place to pursue the BEC work. The essential stepwas the combination of laser cooling and evaporativecooling. My next-door neighbors in Building 26 at MIThave been Dave Pritchard, a pioneer in laser coolingwho conceived the magneto-optical trap, and DanKleppner, who together with Harald Hess and TomGreytak conceived and realized evaporative cooling (seeFig. 9). I feel privileged for the opportunity to combinetheir work and take it to the next level. It is hard tooverestimate the roles which Dave Pritchard and DanKleppner have played for modern atomic physics. Thefamily tree of atomic physicists (Fig. 10) shows some ofthe remarkable physicists who were trained and inspiredby them.

Looking back, it seems that many techniques such asthe Dark SPOT, compressed MOT (Petrich et al., 1994),the TOP trap and the optically plugged trap were criticalfor first demonstrating BEC, but by no means indispens-able. This is best illustrated by the experiment at Rice,which used only Doppler cooling to load the magnetictrap—a technique which had been developed in the1980s. The collision rate was slow, but an excellentvacuum made a very slow evaporation process possible(Bradley et al., 1995). So in hindsight, BEC in alkaligases did not require major innovations in cooling andtrapping. It merely required enough optimism to risk afew years in the attempt to combine laser and evapora-tive cooling. Such an attempt needed a few years of veryfocused work as it involved the integration of severaltechnologies that were not standard in the field, includ-ing ultrahigh vacuum, sensitive CCD cameras and imageprocessing, high-current power supplies for magnetictraps, and flexible computer control of a multistep cool-ing and detection process. Figure 11 compares a state-of-the-art laser cooling experiment in 1993 to a BECexperiment in 2001 using the same vacuum apparatus inthe same laboratory at MIT. A lot of components havebeen added, and I continue to be impressed by my col-laborators, who now handle experiments far more com-plex than I did some five years ago.

D. The cloverleaf trap

After our first observation of BEC, we made the rightdecision for the wrong reason. We expected many othergroups to quickly upgrade their laser cooling experi-ments to magnetic trapping and evaporative cooling,and to join in during the next few months. Nobody ex-pected that it would take almost two years before thenext groups succeeded in reaching BEC (the groups ofDan Heinzen, Lene Hau, Mark Kasevich, and GerhardRempe followed in 1997). I was concerned that ourplugged trap would put us at a disadvantage since thetrapping potential strongly depended on the shape andalignment of the laser focus. So we decided to install thecloverleaf trap instead and discontinue our plugged trapafter only two experimental BEC ‘‘runs.’’

Since we didn’t want to break the vacuum, we in-stalled the new trap in an unfavorable geometry. The

Page 29: Nobel Lecture: Bose-Einstein condensation in a dilute gas ...duine102/summerschoolTP/BECNobel.pdf · experimental Bose-Einstein condensation. For a somewhat more scholarly treatment

1140 Wolfgang Ketterle: When atoms behave as waves

FIG. 7. Observation of Bose-Einstein condensation by absorption imaging. Shown is absorption vs two spatial dimensions. TheBose-Einstein condensate is characterized by its slow expansion observed after 6 ms time of flight. The left picture shows anexpanding cloud cooled to just above the transition point; middle: just after the condensate appeared; right: after further evapo-rative cooling has left an almost pure condensate. The total number of atoms at the phase transition is about 73105, thetemperature at the transition point is 2 mK [Color].

magnet coils for the plugged trap were oriented verti-cally in reentrant flanges, and when we replaced themwith cloverleaf coils, the weakly confining axis of theIoffe-Pritchard trap was vertical. In such a geometry, thegravitational sag would reduce the efficiency of rf-induced evaporation since atoms would only evaporate

FIG. 8. Progress in evaporative cooling of alkali atoms up to1996. The number of atoms in the lowest quantum state isproportional to the phase-space density and has to exceed acritical number of 2.612 to achieve Bose-Einstein condensa-tion. For N0,1023, the increase in phase-space density due toevaporation is plotted. For the Rice result of July 1995 seeBradley et al. (1995) and the erratum (Bradley et al., 1997).

Rev. Mod. Phys., Vol. 74, No. 4, October 2002

at the bottom of the cloud (Ketterle and van Druten,1996; Surkov et al., 1996). But before breaking thevacuum and reorienting the coils, we wanted to see thelimitation. Around December 1995, when we were juststarting to look at the efficiency of evaporation, we lostthe vacuum once again due to a cracked ceramic part inan electric feedthrough and decided to reorient thewhole experiment, with the weakly confining axis of thetrap now aligned horizontally. Since that time, now morethan six years, the machine has been under vacuum. Thisis in sharp contrast to the conditions in 1995, when wehad to open the chamber, pump down, and bake outevery couple of months. Finally, we had learned fromour previous mistakes and developed a very systematicprocedure for pumpdowns and bakeouts.

I still remember the night of March 13, 1996, when theexperiment was up and running, and Klaasjan vanDruten and I had fine-tuned the bias field of the mag-netic trap, so that the switchover to the new magnetictrap was finally completed. It was already after mid-night, too late to start some serious work, when Klaasjanasked half jokingly why don’t we just try to get BEC.Without knowing what our temperatures and densitieswere, without having ever measured the trap frequen-cies, we played around with the rf sweep that determinesthe cooling trajectory, and a condensate showed uparound 2:10 a.m. We were relieved since we hadn’t pro-duced condensates for almost half a year, but also theease at which we got the condensate in a new trap toldus our setup was robust and that we were ready to

Page 30: Nobel Lecture: Bose-Einstein condensation in a dilute gas ...duine102/summerschoolTP/BECNobel.pdf · experimental Bose-Einstein condensation. For a somewhat more scholarly treatment

1141Wolfgang Ketterle: When atoms behave as waves

FIG. 9. MIT faculty in ultralow-temperature atomic physics. Dan Kleppner, W.K., Tom Greytak, and Dave Pritchard look at thelatest sodium BEC apparatus [Color].

switch from engineering cooling schemes and traps tothe study of the condensate. The cloverleaf trap andother winding patterns for the Ioffe-Pritchard configura-tion are now used by almost all BEC experiments. Fig-ure 12 shows the experimental setup during those days.

Why hadn’t we considered this trap earlier andavoided the detours with the quadrupole trap, Majorana

FIG. 10. Family tree of atomic physicists. People with namesin italics are Nobel laureates.

Rev. Mod. Phys., Vol. 74, No. 4, October 2002

flops, and plugging the hole? First, the quadrupole trapwas simpler to build, and it allowed us to pursue evapo-rative cooling faster. Second, we initially favored thequadrupole trap based on an analysis which shows thatconfinement by a linear potential is much stronger thanby the quadratic potential of the Ioffe-Pritchard configu-ration (Ketterle, Durfee, and Stamper-Kurn, 1999).However, a very elongated Ioffe-Pritchard trap provideseffectively linear confinement in the two radial direc-tions, and it was only in 1995 that I realized that it wouldbe easy to adiabatically deform the round laser-cooledcloud to such an elongated shape.

The next weeks were exciting and dramatic: we imple-mented dispersive imaging and saw for the first time thecondensate in the trap. We could take images nonde-structively and recorded two sequential images of thesame condensate. After year-long concerns of how frag-ile and sensitive the condensate would be once created,it was an overwhelming experience to observe the con-densate without destroying it. Figure 13 shows a spatialimage of a condensate; it was taken in nondestructivedispersive imaging. We first implemented dispersive im-aging using the dark-ground technique (Andrews et al.,1996), but soon upgraded to phase-contrast imaging,which was the technique used to record the figure.

Page 31: Nobel Lecture: Bose-Einstein condensation in a dilute gas ...duine102/summerschoolTP/BECNobel.pdf · experimental Bose-Einstein condensation. For a somewhat more scholarly treatment

1142 Wolfgang Ketterle: When atoms behave as waves

FIG. 11. Comparison of a laser cooling and BEC experiment. The first photograph shows the author in 1993 working on the DarkSPOT trap. In the following years, this laser cooling experiment was upgraded to a BEC experiment. The second photographshows the same apparatus in 2001 after many additional components have been added [Color].

In the first week of April 1996, there was a workshopon ‘‘Collective effects in ultracold atomic gases’’ in LesHouches, France, where most of the leading groupswere represented. It was the first such meeting after thesummer of 1995, and it was not without strong emotionsthat I reported our results. Since no other experimentalgroup had made major progress in BEC over the lastfew months, it was our work which provided optimismfor further rapid developments.

E. Interference between two condensates

After we got BEC in the cloverleaf trap, both the ma-chine and the group were in overdrive. After years ofbuilding and improving, frequent failures and frustra-tion, it was like a phase transition to a situation wherealmost everything worked. Within three months aftergetting a condensate in the cloverleaf trap we had writ-ten three papers on the new trap and the phase transi-tion (Mewes et al., 1996a), on nondestructive imaging(Andrews et al., 1996), and on collective excitations(Mewes et al., 1996b). Klaasjan van Druten left thegroup, shortly after Christopher Townsend had joined usas a postdoc. As the next major goal, we decided tostudy the coherence of the condensate. With our opticalplug, we had already developed the tool to split a con-

Rev. Mod. Phys., Vol. 74, No. 4, October 2002

densate into two halves and hoped to observe their in-terference, which would be a clear signature of the long-range spatial coherence.

Around the same time, the idea came up to extractatoms from the condensate using rf induced spin flips—the rf output coupler. Some theorists regarded an outputcoupler as an open question in the context of the atomlaser. I suggested to my group that we could simply pulseon the radio-frequency source that was already usedduring evaporation, and couple atoms out of the con-densate by flipping their spin to a nontrapped state (Fig.14). The experiment worked the first time we tried it(but the quantitative work took awhile; Mewes et al.,1997). I have never regarded the output coupler as oneof our major accomplishments because it was so simple,but it had impact on the community and nobody hasever since regarded outcoupling as a problem!

In July 1996, we had the first results on the rf outputcoupler, and also saw the first fringes when two conden-sates were separated with a sheet of green light andoverlapped in ballistic expansion. I was in Australia forvacation and for the IQEC conference in Sydney. Bye-mail and telephone I discussed with my group the newresults. The fringes were most pronounced when thecondensates were accelerated into each other by remov-ing the light sheet shortly before switching off the mag-

Page 32: Nobel Lecture: Bose-Einstein condensation in a dilute gas ...duine102/summerschoolTP/BECNobel.pdf · experimental Bose-Einstein condensation. For a somewhat more scholarly treatment

1143Wolfgang Ketterle: When atoms behave as waves

FIG. 11. (Continued.)

netic trap. We concluded that some of the fringes mightbe related to sound and other collective effects that oc-cur when two condensates at fairly high density ‘‘touch’’each other. I presented those results at the Sydney meet-ing only to illustrate we were able to do experimentswith two condensates, but now we had to sort out whatwas happening.

It took us four more months until we observed cleaninterference between two condensates. When two con-densates that were initially separated by a distance dinterfere and the interference pattern is recorded after atime t of ballistic expansion, then the fringe spacing isthe de Broglie wavelength h/mv associated with therelative velocity v5d/t . For our geometry with two con-densates about 100 mm in length, we estimated that wewould need at least 60 ms of time of flight to observefringes with a 10-mm period, close to the resolution ofour imaging system. Unfortunately, due to gravity, theatoms dropped out of the field of view of our windowsafter 40 ms. So we tried to gain a longer expansion timein a fountain geometry where we magnetically launchedthe atoms and observed them when they fell backthrough the observation region after more than 100 ms(Townsend et al., 1997), but the clouds were distorted.We also tried to compensate gravity by a verticalmagnetic-field gradient. Some time later I learned about

Rev. Mod. Phys., Vol. 74, No. 4, October 2002

new calculations by the theory group at the Max PlanckInstitute in Garching, showing that the effective separa-tion of two elongated condensates is smaller than theircenter-of-mass separation (Rohrl et al., 1997). Thismeant that we could observe interference fringes afteronly 40 ms, just before the atoms fell out of the obser-vation region. We immediately had a discussion in thegroup and decided to stop working on fountains and‘‘antigravity’’ and simply let the atoms fall by 8 mm dur-ing 40 ms.

We made some ambiguous observations where we sawlow-contrast fringes together with some optical interfer-ence patterns of the probe light, but the breakthroughcame on November 21, 1996, when we observed strikinginterference patterns (Fig. 15). I still remember the situ-ation late that night when we wondered how could weprove beyond all doubt that these were matter-wave in-terference patterns and not some form of self-diffractionof a condensate confined by a light sheet and then re-leased. We came up with the idea of eliminating one ofthe condensates in the last moment by focusing resonantyellow light on it. Whimsically, this laser beam wasdubbed the ‘‘flame thrower.’’ If the fringes were self-diffraction due to the sharp edge in the confinement,they would remain; if they were true interference theywould vanish. This was like a double slit experiment in

Page 33: Nobel Lecture: Bose-Einstein condensation in a dilute gas ...duine102/summerschoolTP/BECNobel.pdf · experimental Bose-Einstein condensation. For a somewhat more scholarly treatment

1144 Wolfgang Ketterle: When atoms behave as waves

optics where you cover one of the slits. It took a fewhours to align the new laser beam, and we verified inphase-contrast imaging that we were able to selectivelyeliminate one of the two condensates.

We had a switch in our control panel which toggledbetween condensate elimination on and off. Then wewent back and aligned the setup for the observation ofinterference. When we toggled the switch we had to waitfor about half a minute until a new condensate was pro-duced. This was the moment of truth. If the fringes ap-peared without a second condensate, then Nature wouldhave fooled us for the whole night—but they disap-peared and an enormous tension disappeared, as well. Itwas already early the next morning, with people arrivingto work. I walked to Dan Kleppner’s office and told himthere was something he should see. So he shared themoment with us where we toggled the switch on alter-nating cooling cycles and correspondingly, the interfer-

FIG. 12. Experimental setup for cooling sodium atoms toBose-Einstein condensation around 1996. The atoms aretrapped and cooled in the center of the ultrahigh-vacuum(UHV) chamber. The atomic beam oven and the Zeemanslower are to the left (outside the photo). The cloverleaf mag-netic trap was mounted horizontally in reentrant flanges. Onlythe leads for the current and water cooling are visible. Thediagonal flange above accommodated a BNC feedthrough forradio-frequency fields which were used to control the evapora-tive cooling. The lens and the mirror above the chamber wereused to observe the condensate by dispersive or absorptionimaging [Color].

Rev. Mod. Phys., Vol. 74, No. 4, October 2002

ence pattern disappeared and reappeared. Interferencebetween two light beams is quite a sight, but with atomsit is more dramatic. Destructive interference means thatatoms plus atoms add up to vacuum!

The evidence for interference was so compelling thatwe submitted our paper based solely on the data of oneexperimental run (Andrews et al., 1997). This run ismemorable to me for another reason: it was to be thelast time I played a major role in preparing and runningan experiment. During the night, I had put in the opticsfor the ‘‘flame thrower.’’ Up to then, I was familiar withevery piece of equipment in the lab and never thoughtthis could change quickly, but it was like another phasetransition. Hans-Joachim Miesner had just arrived, thefirst postdoc who stayed for more than a year, and hesoon took over much responsibility for organizing thelab. There were more demands on my time to write pa-pers and give talks, the group grew with the addition oftwo more graduate students (Shin Inouye and ChrisKuklewicz), and we had intensified our efforts to build asecond BEC experiment. All this coincided in a fewmonths. After earning my Ph.D. in 1986, I had spenteleven more years in the lab during three postdoc posi-tions and as an assistant professor, but now began toplay an advisory role.

FIG. 13. Phase contrast images of trapped Bose gases acrossthe BEC phase transition. At high temperature, above theBEC transition temperature, the density profile of the gas issmooth. As the temperature drops below the BEC phase tran-sition, a high-density core of atoms appears in the center of thedistribution. This is the Bose-Einstein condensate. Loweringthe temperature further, the condensate number grows and thethermal wings of the distribution become shorter. Finally, thetemperature drops to the point where a pure condensate withno discernible thermal fraction remains. Each image shows anequilibrated gas obtained in one complete trapping and cool-ing cycle. The axial and radial frequencies are about 17 and230 Hz, respectively.

Page 34: Nobel Lecture: Bose-Einstein condensation in a dilute gas ...duine102/summerschoolTP/BECNobel.pdf · experimental Bose-Einstein condensation. For a somewhat more scholarly treatment

1145Wolfgang Ketterle: When atoms behave as waves

The papers on the rf output coupler (Mewes et al.,1997) and the interference (Andrews et al., 1997) of twocondensates appeared in the same week in January 1997.Together they demonstrated the ability to create mul-tiple pulses of coherent atoms, and have been regardedas the realization of an atom laser. The period startingwith the early dreams of pursuing BEC and ending withthe observation of the coherence of the condensate wasremarkable. It was full of speculation, dreams, unknownphysics, failures and successes, passion, excitement, andfrustration. This period fused together a team of verydifferent people who had one common denominator: the

FIG. 14. The MIT atom laser operating at 200 Hz. Pulses ofcoherent sodium atoms are coupled out from a Bose-Einsteincondensate confined in a magnetic trap (field of view 2.535.0 mm2). Every 5 ms, a short rf pulse transferred a fractionof these atoms into an unconfined quantum state. These atomswere accelerated downward by gravity and spread out due torepulsive interactions. The atom pulses were observed by ab-sorption imaging. Each pulse contained between 105 and 106

atoms.

FIG. 15. Interference pattern of two expanding condensatesobserved after 40 ms time of flight. The width of the absorp-tion image is 1.1 mm. The interference fringes have a spacingof 15 mm and are strong evidence for the long-range coherenceof Bose-Einstein condensates.

Rev. Mod. Phys., Vol. 74, No. 4, October 2002

passion for experimental physics. It was a unique expe-rience for me to work with these outstanding people(Fig. 16).

VI. THE MAGIC OF MATTER WAVES

Many studies of BEC’s have been performed over thelast several years. The progress until 1998 is nicely sum-marized in the Varenna summer school proceedings (In-guscio et al., 1999). The studies that were most excitingfor me displayed macroscopic quantum mechanics, thewavelike properties of matter on a macroscopic scale.These were also phenomena that no ordinary gas wouldshow and illustrated dramatically that a new form ofmatter had been created. The interference of two con-densates presented above (Fig. 15) is one such example.In the following, I want to discuss the amplificationof atoms and the observation of lattices of quantizedvortices.

These two examples are representative of the two ar-eas into which research on gaseous BEC can be divided:in the first (which could be labeled ‘‘the atomic conden-sate as a coherent gas’’ or ‘‘atom lasers’’), one would liketo have as little interaction as possible—almost like thephotons in a laser. The experiments are preferably doneat low densities. The Bose-Einstein condensate serves asan intense source of ultracold coherent atoms for experi-ments in atom optics, in precision studies or for explo-rations of basic aspects of quantum mechanics. The sec-ond area could be labeled ‘‘BEC as a new quantumfluid’’ or ‘‘BEC as a many-body system.’’ The focus hereis on the interactions between the atoms that are mostpronounced at high densities. The coherent amplifica-tion of atoms is an example of atom optics with conden-sates, and the study of vortices addresses the superfluidproperties of the gas.

FIG. 16. Team photo. This photo was taken in early 1996 infront of the MIT dome. The bottle of champagne was emptiedto celebrate BEC in the cloverleaf trap. Names and photos ofother collaborators during the period 1992–1996 have beenadded.

Page 35: Nobel Lecture: Bose-Einstein condensation in a dilute gas ...duine102/summerschoolTP/BECNobel.pdf · experimental Bose-Einstein condensation. For a somewhat more scholarly treatment

1146 Wolfgang Ketterle: When atoms behave as waves

A. Amplification of atoms in a Bose-Einstein condensate

Since atoms are de Broglie waves, there are manyanalogies between atoms and light, which consists ofelectromagnetic waves. This is exploited in the field ofatom optics where atoms are reflected, diffracted, andinterfere using various atom-optical elements (Adamset al., 1994). One important question was whether theseanalogies can be extended to the optical laser, which isbased on the amplification of light. When our groupdemonstrated a rudimentary atom laser in 1997 we hadsolved the problem of outcoupling (or extracting) atomsfrom the BEC and of verifying their coherence. Theatomic amplification process happened during the for-mation of the Bose-Einstein condensate (Miesner et al.,1998) which is quite different from the way light is am-plified in passing through an active medium. It was onlyin 1999 that our group managed to observe the amplifi-cation of atoms passing through another cloud of atomsserving as the active medium [Inouye, Pfau, et al., 1999(simultaneously with the group in Tokyo; Kozuma et al.,1999)].

Amplifying atoms is more subtle than amplifying elec-tromagnetic waves because atoms can only change theirquantum state and cannot be created. Therefore, even ifone could amplify gold atoms, one would not realize thedreams of medieval alchemy. An atom amplifier con-verts atoms from the active medium into an atomic wavethat is exactly in the same quantum state as the inputwave (Fig. 17).

The atom amplifier requires a reservoir, or an activemedium, of ultracold atoms that have a very narrowspread of velocities and can be transferred to the atomicbeam. A natural choice for the reservoir was a Bose-Einstein condensate. One also needs a coupling mecha-nism that transfers atoms from the reservoir at rest to aninput mode while conserving energy and momentum.This transfer of atoms was accomplished by scatteringlaser light. The recoil of the scattering process acceler-

FIG. 17. Amplification of light and atoms: In the optical laser,light is amplified by passing it through an excited inverted me-dium. In the MIT atom amplifier, an input matter wave is sentthrough a Bose-Einstein condensate illuminated by laser light.Bosonic stimulation by the input atoms causes light to be scat-tered by the condensate exactly at the angle at which a recoil-ing condensate atom joins the input matter wave and augmentsit.

Rev. Mod. Phys., Vol. 74, No. 4, October 2002

ated some atoms to exactly match the velocity of theinput atoms (Fig. 18). Not only were the atoms ampli-fied, but they were in exactly the same motional state asthe input atoms, i.e., they had the same quantum-mechanical phase. This was verified by interfering theamplified output with a copy of the input wave and ob-serving phase coherence.

This direct observation of atom amplification in thesummer of 1999 was preceded by a surprising occur-rence late one night in October 1998 when we discov-ered a new form of superradiance (Inouye, Chikkatur,et al., 1999). We were studying Bragg spectroscopy(Stenger et al., 1999) and illuminated a BEC with twolaser beams. I had no role in the running of the experi-ment and was working in my office, when around mid-night the students came from the lab and told me thatthey saw atoms shooting out from the condensate with avelocity component perpendicular to the direction of thelaser beams. We expected atoms to receive recoil mo-mentum only along the laser beams, and all motion per-pendicular to it to be diffuse due to the random direc-tion of spontaneous Rayleigh scattering.

The whole lab started to discuss what was going on.With a running machine, everything could be tried outimmediately. The first ideas were mundane: let’s illumi-nate the condensate with only one laser beam and seewhat happens (the directional beams remained). Wescrutinized the experimental setup for bouncing laserbeams or beams which had not been completelyswitched off, but we found nothing. Increasingly, we con-sidered that the observed phenomenon was genuine andnot due to some experimental artifact. Knowing that thecondensate was pencil shaped, the idea of laser emissionalong the long condensate axis came up, and this wasalready very close. We decided to stop the general dis-cussion and continue taking data; the machine was run-ning well and we wanted to take advantage of it. Sosome students, including Shin Inouye and AnanthChikkatur, characterized the phenomenon, while Dan

FIG. 18. Observation of atom amplification. Atom amplifica-tion is probed by sending an input beam through the atomamplifier, which is a Bose-Einstein condensate (BEC) illumi-nated with laser light. On the left side, the input beam haspassed through the condensate without amplification. Some 20ms later, a shadow picture is taken of the condensate and theinput atoms. When the amplification process was activated byilluminating the condensate with laser light, the output pulsecontained many more atoms than the input pulse—typical am-plification factors were between 10 and 100. The field of view is1.932.6 mm2.

Page 36: Nobel Lecture: Bose-Einstein condensation in a dilute gas ...duine102/summerschoolTP/BECNobel.pdf · experimental Bose-Einstein condensation. For a somewhat more scholarly treatment

1147Wolfgang Ketterle: When atoms behave as waves

Stamper-Kurn and I went to a blackboard and tried tofigure out what was going on. Within the next hour, wedeveloped the correct semiclassical description of super-radiance in a condensate. In the lab, the predicted strongdependence on laser polarization was verified. A fewmonths later we realized how we could use the super-radiant amplification mechanism to build a phase-coherent atom amplifier. However, the labs were under-going complete renovation at this point and we had towait until the machine was running again before thephase-coherent amplification was implemented.

The demonstration of an atom amplifier added a newelement to atom optics. In addition to passive elementslike beam splitters, lenses, and mirrors, there is now anactive atom-optical element. Coherent matter wave am-plifiers may improve the performance of atom interfer-ometers by making up for losses inside the device or byamplifying the output signal. Atom interferometers arealready used as precise gravity and rotation sensors.

B. Observation of vortex lattices in Bose-Einsteincondensates

Quantum mechanics and the wave nature of matterhave subtle manifestations when particles have angularmomentum, or more generally, when quantum systemsare rotating. When a quantum-mechanical particlemoves in a circle the circumference of the orbit has to bean integer multiple of the de Broglie wavelength. Thisquantization rule leads to the Bohr model and the dis-crete energy levels of the hydrogen atom. For a rotatingsuperfluid, it leads to quantized vortices (Nozieres andPines, 1990). If one spins a normal liquid in a bucket, thefluid will finally rotate as a rigid body where the velocitysmoothly increases from the center to the edge (Fig. 19,left). However, this smooth variation is impossible forparticles in a single quantum state. To fulfill the above-mentioned quantization rule, the flow field has to de-velop singular regions where the number of de Brogliewavelengths on a closed path jumps up by one. Onepossibility would be a radially symmetric flow field withconcentric rings. Between adjacent rings, the number ofde Broglie wavelengths on a circumference wouldchange by one.

FIG. 19. Comparison of the flow fields of rotating normal liq-uids and superfluids. A normal fluid undergoes rigid body ro-tation, whereas a superfluid develops an array of quantizedvortices.

Rev. Mod. Phys., Vol. 74, No. 4, October 2002

However, the energetically most favorable configura-tion is achieved when the singularities in the velocityfield are not distributed on cylindrical shells, but onlines. This corresponds to an array of vortices. In con-trast to classical vortices like those in tornados or in aflushing toilet, the vortices in a Bose-Einstein conden-sate are quantized: when an atom goes around the vor-tex core, its quantum-mechanical phase changes by ex-actly 2p. Such quantized vortices play a key role insuperfluidity and superconductivity. In superconductors,magnetic flux lines arrange themselves in regular latticesthat have been directly imaged. In superfluids, previousdirect observations of vortices had been limited to smallarrays (up to 11 vortices), both in liquid 4He (Yarmchuket al., 1979) and in rotating gaseous Bose-Einstein con-densates (BEC’s) by a group in Paris (Madison et al.,2000).

In 2001, our group observed the formation of highly-ordered vortex lattices in a rotating Bose-condensed gas(Abo-Shaeer et al., 2001). They were produced by spin-ning laser beams around the condensate, thus setting itinto rotation. The condensate then exhibited a remark-able manifestation of quantum mechanics at a macro-scopic level. The rotating gas cloud was riddled withmore than 100 vortices. Since the vortex cores weresmaller than the optical resolution, the gas was allowedto ballistically expand after the magnetic trap wasswitched off. This magnified the spatial structures 20-fold. A shadow picture of these clouds showed littlebright spots where the light penetrated through theempty vortex cores as if through tunnels (Fig. 20 shows anegative image).

A striking feature of the observed vortex lattices is theextreme regularity, free of any major distortions, even

FIG. 20. Observation of vortex lattices in rotating Bose-Einstein condensates. The examples shown contain (A) 16, (B)32, (C) 80, and (D) 130 vortices as the speed of rotation wasincreased. The vortices have ‘‘crystallized’’ in a triangular pat-tern. The diameter of the cloud in (D) was 1 mm after ballisticexpansion, which represents a magnification of 20. (Reprintedwith permission from Abo-Shaeer et al., 2001. Copyright 2001American Association for the Advancement of Science.)

Page 37: Nobel Lecture: Bose-Einstein condensation in a dilute gas ...duine102/summerschoolTP/BECNobel.pdf · experimental Bose-Einstein condensation. For a somewhat more scholarly treatment

1148 Wolfgang Ketterle: When atoms behave as waves

near the boundary. Such ‘‘Abrikosov’’ lattices were firstpredicted for quantized magnetic flux lines in type-II su-perconductors. However, Nature is not always perfect:some of the images showed distortions or defects of thevortex lattices; two examples are shown in Fig. 21. Thephysics of vortices is very rich. Subsequent work by mygroup and others has started to address the dynamicsand nonequilibrium properties of vortex structures. Howare vortices formed? How do they decay? Are the vor-tices straight or bent? Such experiments can be directlycompared with first-principles calculations, which arepossible for such a dilute system. This interplay betweentheory and experiment may lead to a better understand-ing of superfluidity and macroscopic quantum phenom-ena.

VII. OUTLOOK

The rapid pace of developments in atomic BEC dur-ing the last few years has taken the community by sur-prise. After decades of searching for an elusive goal,nobody expected that condensates would be so robustand relatively easy to manipulate. Further, nobody imag-ined that such a simple system would pose so many chal-lenges, not only to experimentalists, but also to our fun-damental understanding of physics. The list of futurechallenges, both for theorists and for experimentalists, islong and includes the exploration of superfluidity andsecond sound in Bose gases, the physics of correlationsand nonclassical wave functions (phenomena beyondthe Gross-Pitaevskii equation), the study of quantum-degenerate molecules and Fermi gases, the developmentof practical ‘‘high-power’’ atom lasers, and their applica-tion in atom optics and precision measurements. Thesescientific goals are closely interwoven with technologicaladvances to produce new single- or multi-speciesquantum-degenerate systems and novel ways of manipu-lation, e.g., using microtraps and atom chips. There isevery indication for more excitement to come!

Work on BEC at MIT has been a tremendous teameffort, and I am grateful to the past and present collabo-rators who have shared both the excitement and thehard work: J. R. Abo-Shaeer, M. R. Andrews, M. Boyd,G. Campbell, A. P. Chikkatur, J.-K. Chin, K. B. Davis,

FIG. 21. Vortex lattices with defects. In the left image, thelattice has a dislocation near the center of the condensate. Inthe right one, there is a defect reminiscent of a grain boundary.(Reprinted with permission from Abo-Shaeer et al., 2001.Copyright 2001 American Association for the Advancement ofScience.)

Rev. Mod. Phys., Vol. 74, No. 4, October 2002

K. Dieckmann, D. S. Durfee, A. Gorlitz, S. Gupta, T. L.Gustavson, Z. Hadzibabic, S. Inouye, M. A. Joffe, D.Kielpinski, M. Kohl, C. E. Kuklewicz, A. E. Leanhardt,R. F. Low, A. Martin, M.-O. Mewes, H.-J. Miesner, R.Onofrio, T. Pfau, D. E. Pritchard, C. Raman, D.Schneble, C. Schunck, Y.-I. Shin, D. M. Stamper-Kurn,C. A. Stan, J. Stenger, E. Streed, Y. Torii, C. G.Townsend, N. J. van Druten, J. M. Vogels, K. Xu, M. W.Zwierlein, and many MIT undergraduate students.Exemplary administrative support has been provided byCarol Costa for more than 12 years. Figure 22 showsthe team in November 2001. Special thanks go to DanKleppner and Tom Greytak for inspiration and constantencouragement. The author also acknowledges the fruit-ful interactions with colleagues all over the world who

FIG. 22. The author with his team in November 2001. Frontrow, from left to right: Z. Hadzibabic, K. Xu, S. Gupta, E.Tsikata, Y.-I. Shin. Middle row: A.P. Chikkatur, J.-K. Chin,D.E. Pritchard, W. K., G. Campbell, A.E. Leanhardt, M. Boyd.Back row: J.R. Abo-Shaeer, D. Schneble, J.M. Vogels, K.Dieckmann, C.A. Stan, Y. Torii, E. Streed.

FIG. 23. Lecturers, seminar speakers, and directors at thesummer school on ‘‘Bose-Einstein Condensation in AtomicGases’’ in Varenna, July 7–17, 1998. Front row: Jean Dalibard,Guglielmo Tino, Fernando Sols, Kris Helmerson. Back row:Sandro Stringari, Carl Wieman, Alexander Fetter, Tilman Es-slinger, Massimo Inguscio, William Phillips, Daniel Heinzen,Peter Fedichev, Lev Pitaevskii, W. K., Allan Griffin, Keith Bur-nett, Daniel Kleppner, Alain Aspect, Ennio Arimondo, The-odor Hansch, Eric Cornell.

Page 38: Nobel Lecture: Bose-Einstein condensation in a dilute gas ...duine102/summerschoolTP/BECNobel.pdf · experimental Bose-Einstein condensation. For a somewhat more scholarly treatment

1149Wolfgang Ketterle: When atoms behave as waves

have contributed to this rich and exciting field. Some ofthese colleagues are depicted in Fig. 23, which is a groupphoto of the lecturers at the Varenna summer school onBEC in 1998. In particular, the yearlong competitionwith the group at Boulder led by Eric Cornell and CarlWieman inspired the best from me and my team, anddespite tight competition, there has been genuine colle-giality and friendship. I want to thank the Office of Na-val Research, the National Science Foundation, theArmy Research Office, the Joint Services ElectronicsProgram, NASA, and the David and Lucile PackardFoundation for their encouragement and financial sup-port of this work.

REFERENCES

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

Adams, C. S., and E. Riis, 1997, Prog. Quantum Electron. 21,1.

Adams, C. S., M. Sigel, and J. Mlynek, 1994, Phys. Rep. 240,143.

Anderson, M. H., J. R. Ensher, M. R. Matthews, C. E. Wie-man, and E. A. Cornell, 1995, Science 269, 198.

Andrews, M. R., M.-O. Mewes, N. J. van Druten, D. S. Durfee,D. M. Kurn, and W. Ketterle, 1996, Science 273, 84.

Andrews, M. R., C. G. Townsend, H.-J. Miesner, D. S. Durfee,D. M. Kurn, and W. Ketterle, 1997, Science 275, 637.

Arimondo, E., W. D. Phillips, and F. Strumia, 1992, Eds., LaserManipulation of Atoms and Ions, Proceedings of the Interna-tional School of Physics Enrico Fermi, Course CXVIII(North-Holland, Amsterdam).

Bagnato, V. S., G. P. Lafyatis, A. G. Martin, E. L. Raab, R. N.Ahmad-Bitar, and D. E. Pritchard, 1987, Phys. Rev. Lett. 58,2194.

Barrett, T. E., S. W. Dapore-Schwartz, M. D. Ray, and G. P.Lafyatis, 1991, Phys. Rev. Lett. 67, 3483.

Baumberg, J. J., 2002, Phys. World March, 37.Bose, S. N., 1924, Z. Phys. 26, 178.Bradley, C. C., C. A. Sackett, J. J. Tollet, and R. G. Hulet,

1995, Phys. Rev. Lett. 75, 1687.Bradley, C. C., C. A. Sackett, J. J. Tollet, and R. G. Hulet,

1997, Phys. Rev. Lett. 79, 1170.Butov, L. V., C. W. Lai, A. L. Ivanov, A. C. Gossard, and D. S.

Chemla, 2002, Nature (London) 417, 47.Cho, H. and G. A. Williams, 1995, Phys. Rev. Lett. 75, 1562.Chu, S., 1998, Rev. Mod. Phys. 70, 685.Chu, S., L. Hollberg, J. E. Bjorkholm, A. Cable, and A. Ash-

kin, 1985, Phys. Rev. Lett. 55, 48.Cline, R. W., D. A. Smith, T. J. Greytak, and D. Kleppner,

1980, Phys. Rev. Lett. 45, 2117.Cohen-Tannoudji, C. N., 1998, Rev. Mod. Phys. 70, 707.Cornell, E. A., C. Monroe, and C. E. Wieman, 1991, Phys. Rev.

Lett. 67, 2439.Crooker, B. C., B. Hebral, E. N. Smith, Y. Takano, and J. D.

Reppy, 1983, Phys. Rev. Lett. 51, 666.Davis, K. B., M.-O. Mewes, M. R. Andrews, N. J. van Druten,

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

Davis, K. B., M.-O. Mewes, M. A. Joffe, M. R. Andrews, andW. Ketterle, 1995, Phys. Rev. Lett. 74, 5202.

Rev. Mod. Phys., Vol. 74, No. 4, October 2002

Davis, K. B., M. O. Mewes, M. A. Joffe, and W. Ketterle, 1994,in Fourteenth International Conference on Atomic Physics,Boulder, Colorado, 1994, Book of Abstracts, 1-M3 (Univer-sity of Colorado, Boulder, Colorado).

DePue, M. T., C. McCormick, S. L. Winoto, S. Oliver, and D.S. Weiss, 1999, Phys. Rev. Lett. 82, 2262.

Doyle, J. M., J. C. Sandberg, I. A. Yu, C. L. Cesar, D. Klepp-ner, and T. J. Greytak, 1991, Phys. Rev. Lett. 67, 603.

Drewsen, M., P. Laurent, A. Nadir, G. Santarelli, A. Clairon, Y.Castin, D. Grison, and C. Salomon, 1994, Appl. Phys. B: La-sers Opt. B59, 283.

Durfee, D. S., and W. Ketterle, 1998, Opt. Express 2, 299.Einstein, A., 1925a, Sitzungsber. Preuss. Akad. Wiss., Phys.

Math. Kl. Bericht 1, 3.Einstein, A., 1925b, Sitzungsber. Preuss. Akad. Wiss., Phys.

Math. Kl. Bericht 3, 18.Fortin, E., E. Benson, and A. Mysyrowicz, 1995, in Bose-

Einstein Condensation, edited by A. Griffin, D. W. Snoke,and S. Stringari (Cambridge University Press, Cambridge)pp. 519–523.

Fried, D. G., T. C. Killian, L. Willmann, D. Landhuis, S. C.Moss, D. Kleppner, and T. J. Greytak, 1998, Phys. Rev. Lett.81, 3811.

Gavroglu, K., 1995, Fritz London: A Scientific Biography(Cambridge University Press, Cambridge).

Goldman, V. V., I. F. Silvera, and A. J. Leggett, 1981, Phys.Rev. B 24, 2870.

Greytak, T. J., 1995, in Bose-Einstein Condensation, edited byA. Griffin, D. W. Snoke, and S. Stringari (Cambridge Univer-sity Press, Cambridge), pp. 131–159.

Greytak, T. J., and D. Kleppner, 1984, in New Trends in AtomicPhysics, Les Houches Summer School 1982, edited by G.Grynberg and R. Stora (North-Holland, Amsterdam), p.1125.

Griffin, A., 1999, in Bose-Einstein Condensation in AtomicGases, Proceedings of the International School of Physics En-rico Fermi, Course CXL, edited by M. Inguscio, S. Stringari,and C. E. Wieman (IOS Press, Amsterdam), p. 1.

Griffin, A., D. W. Snoke, and S. Stringari, 1995, Eds., Bose-Einstein Condensation (Cambridge University Press, Cam-bridge).

Hecht, C. E., 1959, Physica (Amsterdam) 25, 1159.Heinzen, D. J., 1999, in Bose-Einstein Condensation in Atomic

Gases, Proceedings of the International School of Physics En-rico Fermi, Course CXL, edited by M. Inguscio, S. Stringari,and C. E. Wieman (IOS, Amsterdam), pp. 351–390.

Helmerson, K., A. Martin, and D. E. Pritchard, 1992, J. Opt.Soc. Am. B 9, 1988.

Hess, H. F., 1986, Phys. Rev. B 34, 3476.Huang, K., 1987, Statistical Mechanics (Wiley, New York).Huse, D. A., and E. Siggia, 1982, J. Low Temp. Phys. 46, 137.Ido, T., Y. Isoya, and H. Katori, 2000, Phys. Rev. A 61, 061403.Inguscio, M., S. Stringari, and C. E. Wieman, 1999, Eds., Bose-

Einstein Condensation in Atomic Gases, Proceedings of theInternational School of Physics Enrico Fermi, Course CXL(IOS, Amsterdam).

Inouye, S., A. P. Chikkatur, D. M. Stamper-Kurn, J. Stenger, D.E. Pritchard, and W. Ketterle, 1999, Science 285, 571.

Inouye, S., T. Pfau, S. Gupta, A. P. Chikkatur, A. Gorlitz, D. E.Pritchard, and W. Ketterle, 1999, Nature (London) 402, 641.

Joffe, M. A., W. Ketterle, A. Martin, and D. E. Pritchard, 1993,J. Opt. Soc. Am. B 10, 2257.

Kasevich, M., and S. Chu, 1992, Phys. Rev. Lett. 69, 1741.

Page 39: Nobel Lecture: Bose-Einstein condensation in a dilute gas ...duine102/summerschoolTP/BECNobel.pdf · experimental Bose-Einstein condensation. For a somewhat more scholarly treatment

1150 Wolfgang Ketterle: When atoms behave as waves

Kerman, A. J., V. Vuletic, C. Chin, and S. Chu, 2000, Phys.Rev. Lett. 84, 439.

Ketterle, W., K. B. Davis, M. A. Joffe, A. Martin, and D. E.Pritchard, 1993a, Phys. Rev. Lett. 70, 2253.

Ketterle, W., K. B. Davis, M. A. Joffe, A. Martin, and D. E.Pritchard, 1993b, talk given at OSA Annual Meeting, Tor-onto, Canada, October 3–8.

Ketterle, W., D. S. Durfee, and D. M. Stamper-Kurn, 1999, inBose-Einstein Condensation in Atomic Gases, Proceedings ofthe International School of Physics Enrico Fermi, CourseCXL, edited by M. Inguscio, S. Stringari, and C. E. Wieman(IOS Press, Amsterdam), pp. 67–176.

Ketterle, W., and S. Inouye, 2001, C. R. Acad. Sci., Ser IV:Phys., Astrophys. 2, 339.

Ketterle, W., A. Martin, M. A. Joffe, and D. E. Pritchard, 1992,Phys. Rev. Lett. 69, 2483.

Ketterle, W., and D. E. Pritchard, 1992a, Appl. Phys. B: Pho-tophys. Laser Chem. B54, 403.

Ketterle, W., and D. E. Pritchard, 1992b, Phys. Rev. A 46,4051.

Ketterle, W., and J. J. van Druten, 1996, in Advances inAtomic, Molecular, and Optical Physics, edited by B. Beder-son and H. Walther (Academic, San Diego), Vol. 37, pp. 181–236.

Kleppner, D., T. J. Greytak, T. C. Killian, D. G. Fried, L. Will-mann, D. Landhuis, and S. C. Moss, 1999, in Bose-EinsteinCondensation in Atomic Gases, Proceedings of the Interna-tional School of Physics Enrico Fermi, Course CXL, editedby M. Inguscio, S. Stringari, and C. E. Wieman (IOS Press,Amsterdam), pp. 177–1999.

Kozuma, M., Y. Suzuki, Y. Torii, T. Sugiura, T. Kuga, E. W.Hagley, and L. Deng, 1999, Science 286, 2309.

Letokhov, V. S., and V. G. Minogin, 1980, Opt. Commun. 35,199.

Lett, P. D., W. D. Phillips, S. L. Rolston, C. E. Tanner, R. N.Watts, and C. I. Westbrook, 1989, J. Opt. Soc. Am. B 6, 2084.

Lin, J. L., and J. P. Wolfe, 1993, Phys. Rev. Lett. 71, 1222.Lindquist, K., M. Stephens, and C. Wieman, 1992, Phys. Rev.

A 46, 4082.Lovelace, R. V. E., C. Mahanian, T. J. Tommila, and D. M.

Lee, 1985, Nature (London) 318, 30.Madison, K. W., F. Chevy, W. Wohlleben, and J. Dalibard,

2000, J. Mod. Opt. 47, 2715.Masuhara, N., J. M. Doyle, J. C. Sandbert, D. Kleppner, T. J.

Greytak, H. F. Hess, and G. P. Kochanski, 1988, Phys. Rev.Lett. 61, 935.

Metcalf, H., and P. van der Straten, 1994, Phys. Rep. 244, 203.Mewes, M.-O., M. R. Andrews, D. M. Kurn, D. S. Durfee, C.

G. Townsend, and W. Ketterle, 1997, Phys. Rev. Lett. 78, 582.Mewes, M.-O., M. R. Andrews, N. J. van Druten, D. M. Kurn,

D. S. Durfee, and W. Ketterle, 1996a, Phys. Rev. Lett. 77, 416.Mewes, M.-O., M. R. Andrews, N. J. van Druten, D. M. Kurn,

D. S. Durfee, C. G. Townsend, and W. Ketterle, 1996b, Phys.Rev. Lett. 77, 988.

Miesner, H.-J., D. M. Stamper-Kurn, M. R. Andrews, D. S.Durfee, S. Inouye, and W. Ketterle, 1998, Science 279, 1005.

Migdall, A. L., J. V. Prodan, W. D. Phillips, T. H. Bergeman,and H. J. Metcalf, 1985, Phys. Rev. Lett. 54, 2596.

Monroe, C., E. Cornell, and C. Wieman, 1992, in Laser Ma-nipulation of Atoms and Ions, Proceedings of the Interna-tional School of Physics Enrico Fermi, Course CXVIII, editedby E. Arimondo, W. D. Phillips, and F. Strumia (North-Holland, Amsterdam), pp. 361–377.

Rev. Mod. Phys., Vol. 74, No. 4, October 2002

Monroe, C., W. Swann, H. Robinson, and C. Wieman, 1990,Phys. Rev. Lett. 65, 1571.

Nozieres, P., and D. Pines, 1990, The Theory of Quantum Liq-uids (Addison-Wesley, Redwood City, CA).

O’Hara, K. E., L. O. Suilleabhain, and J. P. Wolfe, 1999, Phys.Rev. B 60, 10565.

Oliva, J., 1989, Phys. Rev. B 39, 4197.Pais, A., 1982, Subtle is the Lord, The Science and the Life of

Albert Einstein (Clarendon, Oxford).Petrich, W., M. H. Anderson, J. R. Ensher, and E. A. Cornell,

1994, J. Opt. Soc. Am. B 11, 1332.Petrich, W., M. H. Anderson, J. R. Ensher, and E. A. Cornell,

1995, Phys. Rev. Lett. 74, 3352.Phillips, W. D., 1998, Rev. Mod. Phys. 70, 721.Pritchard, D. E., 1986, in Electronic and Atomic Collisions,

Invited Papers of the XIV International Conference on thePhysics of Electronic and Atomic Collisions, Palo Alto, Cali-fornia, July, 1985, edited by D. C. Lorents, W. E. Meyerhof,and J. R. Peterson (Elsevier, New York), pp. 593–604.

Pritchard, D. E., K. Helmerson, and A. G. Martin, 1989, inAtomic Physics 11, edited by S. Haroche, J. C. Gay, and G.Grynberg (World Scientific, Singapore), p. 179.

Pritchard, D. E., and W. Ketterle, 1992, in Laser Manipulationof Atoms and Ions, Proceedings of the International School ofPhysics Enrico Fermi, Course CXVIII, edited by E. Ari-mondo, W. D. Phillips, and F. Strumia (North-Holland, Am-sterdam), pp. 473–496.

Rasolt, M., M. H. Stephen, M. E. Fisher, and P. B. Weichman,1984, Phys. Rev. Lett. 53, 798.

Reppy, J. D., 1984, Physica B 126, 335.Rohrl, A., M. Naraschewski, A. Schenzle, and H. Wallis, 1997,

Phys. Rev. Lett. 78, 4143.Saba, M., C. Ciuti, J. Bloch, V. Thierry-Mieg, R. Andre, L. S.

Dang, S. Kundermann, A. Mura, G. Bongiovanni, J. Sl. Stae-hli, and B. Deveaud, 2001, Nature (London) 414, 731.

Safonov, A. I., S. A. Vasilyev, L. S. Yasnikov, I. I. Lukashevich,and S. Jaakola, 1998, Phys. Rev. Lett. 81, 4545.

Schrodinger, E., 1952, Statistical Thermodynamics (CambridgeUniversity Press, Cambridge); reprinted by Dover Publica-tions (New York, 1989).

Silvera, I. F., and J. T. M. Walraven, 1980, Phys. Rev. Lett. 44,164.

Silvera, I. F., and J. T. M. Walraven, 1986, in Progress in LowTemperature Physics, edited by D. F. Brewer (Elsevier, Am-sterdam) Vol. X, p. 139.

Stamper-Kurn, D., and W. Ketterle, 2001, in Coherent AtomicMatter Waves, Proceedings of the Les Houches SummerSchool, Course LXXII, 1999, edited by R. Kaiser, C. West-brook, and F. David (Springer, New York).

Stenger, J., S. Inouye, A. P. Chikkatur, D. M. Stamper-Kurn, D.E. Pritchard, and W. Ketterle, 1999, Phys. Rev. Lett. 82, 4569.

Stoof, H. T. C., J. M. V. A. Koelman, and B. J. Verhaar, 1988,Phys. Rev. B 38, 4688.

Stwalley, W. C., and L. H. Nosanow, 1976, Phys. Rev. Lett. 36,910.

Surkov, E. L., J. T. M. Walraven, and G. V. Shlyapnikov, 1996,Phys. Rev. A 53, 3403.

Taubes, G., 1994, Science 265, 184.Townsend, C. G., N. H. Edwards, C. J. Cooper, K. P. Zetie, C.

J. Foot, A. M. Steane, P. Szriftgiser, H. Perrin, and J. Dali-bard, 1995, Phys. Rev. A 52, 1423.

Townsend, C. G., N. H. Edwards, K. P. Zetie, C. J. Cooper, J.Rink, and C. J. Foot, 1996, Phys. Rev. A 53, 1702.

Page 40: Nobel Lecture: Bose-Einstein condensation in a dilute gas ...duine102/summerschoolTP/BECNobel.pdf · experimental Bose-Einstein condensation. For a somewhat more scholarly treatment

1151Wolfgang Ketterle: When atoms behave as waves

Townsend, C. G., N. J. van Druten, M. R. Andrews, D. S.Durfee, D. M. Kurn, M.-O. Mewes, and W. Ketterle, 1997, inAtomic Physics 15, Fifteenth International Conference onAtomic Physics, Amsterdam, August 1996, edited by H. B.van Linden van den Heuvell, J. T. M. Walraven, and M. W.Reynolds (World Scientific, Singapore), pp. 192–211.

Vigue, J., 1986, Phys. Rev. A 34, 4476.Walker, T., and P. Feng, 1994, in Advances in Atomic, Molecu-

lar, and Optical Physics, edited by B. Bederson and H.Walther (Academic, San Diego), Vol. 34, pp. 125–170.

Walker, T., D. Sesko, and C. Wieman, 1990, Phys. Rev. Lett. 64,408.

Walraven, J. T. M., 1996, in Quantum Dynamics of Simple Sys-

Rev. Mod. Phys., Vol. 74, No. 4, October 2002

tems, edited by G. L. Oppo, S. M. Barnett, E. Riis, and M.Wilkinson (Institute of Physics, London), pp. 315–352.

Weiner, J., 1995, in Advances in Atomic, Molecular, and Opti-cal Physics, edited by B. Bederson and H. Walther (Aca-demic, San Diego), Vol. 35, pp. 45–78.

Weiner, J., V. S. Bagnato, S. Zilio, and P. S. Julienne, 1999, Rev.Mod. Phys. 71, 1.

Wolfe, J. P., J. L. Lin, and D. W. Snoke, 1995, in Bose-EinsteinCondensation, edited by A. Griffin, D. W. Snoke, and S.Stringari (Cambridge University Press, Cambridge), pp. 281–329.

Yamamoto, Y., 2000, Nature (London) 405, 629.Yarmchuk, E. J., M. J. V. Gordon, and R. E. Packard, 1979,

Phys. Rev. Lett. 43, 214.