0 ratings0% found this document useful (0 votes) 198 views25 pagesCohen Tannoudji Lecture
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content,
claim it here.
Available Formats
Download as PDF or read online on Scribd
166
MANIPULATING ATOMS WITH PHOTONS
Nobel Lecture, December 8, 1997
by
Ciaupe N. Conen-Taxnounyt
Collége de France et Laboratoire Kastler Brossel* de I’Ecole Normale
Supérieure, 24 rue Lhomond, 75231 Paris Cedex 05, France
Electromagnetic interactions play a central role in low energy physics. They
are responsible for the cohesion of atoms and molecules and they are at the
origin of the emission and absorption of light by such systems. This light is
not only a source of information on the structure of atoms. It can also be
used to act on atoms, to manipulate them, to control their various degrees of
freedom. With the development of laser sources, this research field has con-
siderably expanded during the last few years. Methods have been developed
to trap atoms and to cool them to very low temperatures. This has opened the
way to a wealth of new investigations and applications.
Two types of degrees of freedom can be considered for an atom: the in-
ternal degrees of freedom, such as the electronic configuration or the spin
polarization, in the center of mass system; the external degrees of freedom,
which are essentially the position and the momentum of the center of mass.
The manipulation of internal degrees of freedom goes back to optical pump-
ing [1], which uses resonant exchanges of angular momentum between
atoms and circularly polarized light for polarizing the spins of these atoms.
These experiments predate the use of lasers in atomic physics. The manipu-
lation of external degrees of freedom uses the concept of radiative forces re-
sulting from the exchanges of linear momentum between atoms and light.
Radiative forces exerted by the light coming from the sun were already in-
voked by J. Kepler to explain the tails of the comets. Although they are very
small when one uses ordinary light sources, these forces were also investiga-
ted experimentally in the beginning of this century by P. Lebedev, E. F.
Nichols and G. F. Hull, R. Frisch. For a historical survey of this research field,
we refer the reader to review papers [2, 3, 4, 5) which also include a discus-
sion of early theoretical work dealing with these problems by the groups of
A. P. Kazantsey, V. S. Letokhoy, in Russia, A. Ashkin at Bell Labs, S. Stenholm
in Helsinki.
It turns out that there is a strong interplay between the dynamics of inter-
nal and external degrees of freedom. This is at the origin of efficient laser
cooling mechanisms, such as “Sisyphus cooling” or “Velocity Selective Cohe-
* Laboratoire Kastler Brossel is a Laboratoire associé au CNRS et a l'Université Pierre et Marie
Curie,167
rent Population Trapping”, which were discovered at the end of the 80's (for
a historical survey of these developments, see for example [6]). These me-
chanisms have allowed laser cooling to overcome important fundamental
limits, such as the Doppler limit and the single photon recoil limit, and to
reach the microKelvin, and even the nanoKelvin range. We devote a large
part of this paper (sections 2 and 3) to the discussion of these mechanisms
and to the description of a few applications investigated by our group in Paris
(section 4). There is a certain continuity between these recent developments
in the field of laser cooling and trapping and early theoretical and experi-
mental work performed in the 60’s and the 70's, dealing with internal de-
grees of freedom. We illustrate this continuity by presenting in section 1 a
brief survey of various physical processes, and by interpreting them in terms
of two parameters, the radiative broadening and the light shift of the atomic
ground state.
1, BRIEF REVIEW OF PHYSICAL PROCESSES:
To classify the basic physical processes which are used for manipulating atoms
by light, it is useful to distinguish two large categories of effects: dissipative
(or absorptive) effects on the one hand, reactive (or dispersive) effects on the
other hand. This partition is relevant for both internal and external degrees
of freedom.
1.1 EXISTENCE OF TWO TYPES OF EFFECTS IN ATOM-PHOTON INTERAC-
TIONS
Consider first a light beam with frequency @, propagating through a medium
consisting of atoms with resonance frequency @,. The index of refraction de-
scribing this propagation has an imaginary part and a real part which are as-
sociated with two types of physical processes. The incident photons can be ab-
sorbed, more precisely scattered in all directions. The corresponding
attenuation of the light beam is maximum at resonance. It is described by the
imaginary part of the index of refraction which varies with @,-@, as a Loren
absorption curve. We will call such an effect a dissipative (or absorptive) ef
fect. The speed of propagation of light is also modified. The corresponding
dispersion is described by the real part n of the index of refraction whose dif-
ference from 1, n-I, varies with @,-«, as a Lorentz dispersion curve. We will
call such an effect a reactive (or dispersive) effect.
Dissipative effects and reactive effects also appear for the atoms, as a result
of their interaction with photons. They correspond to a broadening and toa
shift of the atomic energy levels, respectively. Such effects already appear
when the atom interacts with the quantized radiation field in the vacuum
state. It is well known that atomic excited states get a natural width I, which is
also the rate at which a photon is spontaneously emitted from such states.
Atomic energy levels are also shifted as a result of virtual emissions and reab-
sorptions of photons by the atom. Such a radiative correction is simply the
Lamb shift [7].168 Physics 1997
Similar effects are associated with the interaction with an incident light
beam. Atomic ground states get a radiative broadening I’, which is also the
rate at which photons are absorbed by the atom, or more precisely scattered
from the incident beam. Atomic energy levels are also shifted as a result of
virtual absorptions and reemissions of the incident photons by the atom.
Such energy displacements iA’ are called light shifts, or ac Stark shifts [8, 9].
In view of their importance for the following discussions, we give now a
brief derivation of the expressions of I” and A’, using the so-called dressed-
atom approach to atom-photon interactions (see for example [10], chapter
VI). In the absence of coupling, the two dressed states |g, N) (atom in the
ground state gin the presence of N photons) and | ¢ N-J) (atom in the exci-
ted state ¢ in the presence of N-J photons) are separated by a splitting #4,
where 6 = @,-@, is the detuning between the light frequency @, and the
atomic frequency @,. The atom-light interaction Hamiltonian V,, couples
these two states because the atom in gcan absorb one photon and jump to ¢
The corresponding matrix element of V,, can be written as AQ/2, where the
so-called Rabi frequency Q is proportional to the transition dipole moment
and to VN. Under the effect of such a coupling, the two states repel each
other, and the state | g, N) is shifted by an amount fA’, which is the light shift
of g. The contamination of |g, N) by the unstable state | ¢ N-J) (having a
width P) also confers to the ground state a width I’, In the limit where QI’. Such a dependence of 7,,, on the Rabi frequen-Claude N. Cohen-Tannoudji 175
cy 2 and on the detuning 6 has been checked experimentally with cesium
atoms [40]. Fig.3 presents the variations of the measured temperature Twith
the dimensionless parameter Q2/T[6|. Measurements of T'versus intensity for
different values of 6 show that T depends linearly, for low enough intensities,
ona single parameter which is the light shift of the ground state Zeeman sub-
levels.
0 05 U
otair
Figure 8. Temperature measurements in cesium optical molasses (from reference [40]). The left
part of the figure shows the fluorescence light emitted by the molasses observed through a
‘window of the vacuum chamber. The horizontal bright line is the fluorescence light emitted by
the atomic beam which feeds the molasses and which is slowed down by a frequency chirped
laser beam. Right part of the figure : temperature of the atoms measured by a time of flight tech-
nique versus the dimensionless parameter 2/|5|T proportional to the light shift (Q is the op-
tical Rabi frequency, Sthe detuning and I” the natural width of the excited state)
2.2 THE LIMITS OF SISYPHUS COOLING
At low intensity, the light shift U, « #2/5is much smaller than #I. This ex-
plains why Sisyphus cooling leads to temperatures much lower than those
achievable with Doppler cooling. One cannot however decrease indefinitely
the laser intensity. The previous discussion ignores the recoil due to the
spontaneously emitted photons which increases the kinetic energy of the
atom by an amount on the order of Ey, where
Ep = hPk?/2M (3)
is the recoil energy of an atom absorbing or emitting a single photon. When
U, becomes on the order or smaller than Ep, the cooling due to Sisyphus
cooling becomes weaker than the heating due to the recoil, and Sisyphus
cooling no longer works. This shows that the lowest temperatures which can176 Physics 1997
be achieved with such a scheme are on the order of a few Ep/ky This result is
confirmed by a full quantum theory of Sisyphus cooling [41, 42] and is in
good agreement with experimental results. The minimum temperature in
Fig.3 is on the order of LOE p/kg
2.3 OPTICAL LATTICES
For the optimal conditions of Sisyphus cooling, atoms become so cold that
they get trapped in the quantum vibrational levels of a potential well (see
Fig. 4). More precisely, one must consider energy bands in this perodic struc
ture [43]. Experimental observation of such a quantization of atomic motion
Probe transmission
Figure 4. Probe absorption spectrum of a LD optical lattice (from reference (44]). The upper
‘part of the figure shows the two counterpropagating laser beams with frequency @ and orthogo-
nal linear polarizations forming the 1D-optical lattice, and the probe beam with frequency @,
whose absorption is measured by a detector. The lower part of the figure shows the probe trans-
mission versus @,-@. The two lateral resonances corresponding to amplification or absorption of
the probe are due to stimulated Raman processes between vibrational levels of the atoms trapped
in the light field (see the two insets). The central narrow structure is a Rayleigh line due to the
antiferromagnetic spatial order of the atoms.Claude N. Cohen-Tannoudji 177
in an optical potential was first achieved in one dimension [44] [45]. Atoms
are trapped in a spatial periodic array of potential wells, called a “1D-optical
lattice”, with an antiferromagnetic order, since two adjacent potential wells
correspond to opposite spin polarizations. 2D and 3D optical lattices
have been realized subsequently (see the review papers [46] [47]).
3. SUBRECOIL LASER COOLING
3.1 THE SINGLE PHOTON RECOIL LIMIT. HOW TO CIRCUMVENT IT
In most laser cooling schemes, fluorescence cycles never cease. Since the ran-
dom recoil ik communicated to the atom by the spontaneously emitted pho-
tons cannot be controlled, it seems impossible to reduce the atomic momen-
tum spread &p below a value corresponding to the photon momentum fik.
Condition 5p = ik defines the “single photon recoil limit”. It is usual in laser
cooling to define an effective temperature Tin terms of the half-width 6p (at
1/V¢) of the momentum distribution by k,7/2= 6p2/2M. In the temperature
scale, condition 5p = fk defines a “recoil temperature” Ty by:
kpT; PK?
“y= on = ER “
The value of Tp ranges from a few hundred nanoKelvin for alkalis to a few
microKelvin for helium.
0 v
Figure 5. Subrecoil laser cooling. The random walk of the atom in velocity space is supposed to
be characterized by a jump rate Rwhich vanishes for v= 0 (a). As a result of this inhomogeneous
random walk, atoms which fall in a small interval around v = 0 remain trapped there for a long
time, on the order of (R(v)}~!, and accumulate (b).
It is in fact possible to circumvent this limit and to reach temperatures T
lower than T,, a regime which is called “subrecoil” laser cooling. The basic
idea is to create a situation where the photon absorption rate I’, which is
also the jump rate Rof the atomic random walk in velocity space, depends on
the atomic velocity v = p/M and vanishes for v= 0 (Fig.5a). Consider then an
atom with v = 0. For such an atom, the absorption of light is quenched.
Consequently, there is no spontaneous reemission and no associated random
recoil. One protects in this way ultracold atoms (with v ~0) from the “bad” ef-
fects of the light. On the other hand, atoms with v# 0 can absorb and remit
light. In such absorption-spontaneous emission cycles, their velocities change
in a random way and the corresponding random walk in vspace can transfer178 Physics 1997
atoms from the v# 0 absorbing states into the v~ 0 dark states where they re-
main trapped and accumulate (see Fig. 5b). This reminds us of what hap-
pens in a Kundt’s tube where sand grains vibrate in an acoustic standing
wave and accumulate at the nodes of this wave where they no longer move.
Note however that the random walk takes place in velocity space for the situ-
ation considered in Fig. 5b, whereas it takes place in position space in a
Kundt’s tube.
Up to now, two subrecoil cooling schemes have been proposed and de-
monstrated. In the first one, called “Velocity Selective Coherent Population
Trapping” (VSCPT), the vanishing of R(v) for v = 0 is achieved by using
destructive quantum interference between different absorption amplitudes
[48]. The second one, called Raman cooling, uses appropriate sequences of
stimulated Raman and optical pumping pulses for tailoring the appropriate
shape of R(v) [49].
3.2 BRIEF SURVEY OF VSCPT
We first recall the principle of the quenching of absorption by “coherent po-
pulation trapping”, an effect which was discovered and studied in 1976 [50,
51]. Consider the 34evel system of Fig.6, with two ground state sublevels g,
and g and one excited sublevel ¢,, driven by two laser fields with frequencies
),, and ®,», exciting the transitions g, © 4 and g, ¢ é, respectively. Let iA
be the detuning from resonance for the stimulated Raman process consisting
of the absorption of one @,, photon and the stimulated emission of one @»
photon, the atom going from g, to g,. One observes that the fluorescence
2]
Figure 6. Coherent population trapping. A three-level atom g;, f, is driven by owo laser fields
with frequences @,, and 0, exciting the transitions g, ¢ and g, + 4 respectively. hd is the de~
tuning from resonance for the stimulated Raman process induced between g, and g, by the two
laser fields «@,, and @,, When 4= 0, atoms are optically pumped in a linear superposition of g,
and g, which no longer absorbs light because of a destructive interference between the two ab-
sorption amplitudes g, —> 4 and g~> é,Claude N. Cohen-Tannoudji 179
rate R vanishes for A = 0. Plotted versus A, the variations of R are similar to
those of Fig.5a with v replaced by A. The interpretation of this effect is that
atoms are optically pumped into a linear superposition of g, and g, which is
not coupled to @ because of a destructive interference between the two ab-
sorption amplitudes g, —> ¢ and g, — @.
The basic idea of VSCPT is to use the Doppler effect for making the detuning
A of the stimulated Raman process of Fig.6 proportional to the atomic velo-
city v. The quenching of absorption by coherent population trapping is thus
made velocity dependent and one achieves the situation of Fig. 5.a. This is ob-
tained by taking the two laser waves @,, and @;. counterpropagating along
the zaxis and by choosing their frequencies in such a way that A = 0 for an
atom at rest. Then, for an atom moving with a velocity v along the zaxis, the
opposite Doppler shifts of the two laser waves result in a Raman detuning A=
(k, +k.) v proportional to v.
A more quantitative analysis of the cooling process [52] shows that the dark
state, for which R = 0, is a linear superposition of two states which differ not
only by the internal state (g, or g,) but also by the momentum along the =
kbp) = e1 |g, Tiki) + ¢2 192, +k) 6)
This is due to the fact that g, and gy must be associated with different mo-
menta, ~hk, and +iky in order to be coupled to the same excited state
| @ p= 0) by absorption of photons with different momenta +/k, and ~fiky.
Furthermore, when A= 0, the state (5) is a stationary state of the total atom +
laser photons system. As a result of the cooling by VSCPT, the atomic mo-
mentum distribution thus exhibits two sharp peaks, centered at —ik, and
600;
400:
200
°%30°30°40 0 10 2030
v (cm/s)
Figure 7, One-dimensional VSCPT experiment. The left part of the figure shows the experimen-
tal scheme. The cloud of precooled trapped atoms is released while the two counterpropagating
VSCPT beams with orthogonal circular polarizations are applied during a time 6 = Ims. The
atoms then fall freely and their positions are detected 6.5 cm below on a microchannel plate.
‘The double band pattern is a signature of the 1D cooling process which accumulates the atoms
in a state which is a linear superposition of two different momenta. The right part of the figure
gives the velocity distribution of the atoms detected by the microchannel plate. The width dv of
the two peaks is clearly smaller than their separation 2v, where vq ~ 9.2 cm/s is the
recoil velocity. This is a clear signature of subrecoil cooling,180 Physics 1997
+fiky, with a width 6p which tends to zero when the interaction time @ tends
to infinity.
The first VSCPT experiment [48] was performed on the 28S, metastable
state of helium atoms. The two lower states g, and g, were the M=~-1 and
M = +1 Zeeman sublevels of the 2°, metastable state, ¢ was the M = 0
Zeeman sublevel of the excited 25P, state. The two counterpropagating laser
waves had the same frequency @, , = @,)= @, and opposite circular polariza-
tions. The two peaks of the momentum distribution were centered at 2hk,
with a width corresponding to T = Tp /2. The interaction time was then in-
creased by starting from a cloud of trapped precooled helium atoms instead
of using an atomic beam as in the first experiment [53]. This led to much
lower temperatures (see Fig. 8). Very recently, temperatures as low as Tp/800
have been observed [54].
10)
os.
os.
04.
02.
00+
° ‘yy eer) ‘860 1000” 1800 2000 2800” 3000
t(us) ets)
Figure 8. Measurement of the spatial correlation function of atoms cooled by VSCPT (from re-
ference [54]). After a cooling period of duration @, the two VSCPT beams are switched off dur
ing a dark period of duration ,, The two coherent wave packets into which atoms are pumped fly
apart with a relative velocity 2v, and get separated by a distance a = 2vgf,. Reapplying the two
VSCPT beams during a short probe pulse, one measures a signal S which can be shown to be
equal to [1 + G(a)]/2 where G(«) is the spatial overlap between the two identical wave packets se~
arated by a. From G(a), which is the spatial correlation function of each wave packet, one de-
termines the atomic momentum distribution which is the Fourier transform of G(a). The left
part of the figure gives Sversus a. The right part of the figure gives T,/T versus the cooling time
8, where T, is the recoil temperature and T the temperature of the cooled atoms determined
from the width of G(a). The straight line is a linear fit in agreement with the theoretical predic-
tions of Lévy statistics. The lowest temperature, on the order of T,/800, is equal to 5 nK.
In fact, it is not easy to measure such low temperatures by the usual time of
flight techniques, because the resolution is then limited by the spatial extent
of the initial atomic cloud. A new method has been developed [54] which
consists of measuring directly the spatial correlation function of the atoms
Gla) = Jt%_ dzp*(z + @)9(z), where 9(z) is the wave function of the atomic
wave packet. This correlation function, which describes the degree of spatial
coherence between two points separated by a distance a, is simply the Fourier
transform of the momentum distribution | 9(f)|?. This method is analogous
to Fourier spectroscopy in optics, where a narrow spectral line (#) is more
easily inferred from the correlation function of the emitted electric field G(t)Claude N. Cohen-Tannoudji 181
‘2 duh¥(t+ 7).B(2), which is the Fourier transform of 1(q). Experimentally,
the measurement of G(a) is achieved by letting the two coherent VSCPT
wave packets fly apart with a relative velocity 2up = 2hk/m during a dark peri-
od t,, during which the VSCPT beams are switched off. During this dark pe-
tiod, the two wave packets get separated by a distance a = 2up/p, and one then
measures with a probe pulse a signal proportional to their overlap. Fig.8a
shows the variations with 4, of such a signal S (which is in fact equal to [1 +
G(a)]/2). From such a curve, one deduces a temperature T ~ Tg/625, cor-
responding to 5p =hk/25. Fig.8b shows the variations of T,,/Twith the VSCPT
interaction time @, As predicted by theory (see next subsection), Ty/T varies
linearly with @ and can reach values as large as 800.
LPS FS
se (a) 300
15
Ov (emis) “15
1s
v, (cms) 15 v, (emis)
1s
Figure 9. Two-dimensional VSCPT experiment (from reference [58]). The experimemtal scheme
is the same as in Fig, 7, but one uses now four VSCPT beams in a horizontal plane and atoms are
pumped into a linear superposition of four different momentum states giving rise to four peaks
in the two-dimensional velocity distribution (a). When three of the four VSCPT laser beams are
adiabatically switched off, the whole atomic population is transferred into a single wave packet
(b).
VSCPT has been extended to two [55] and three [56] dimensions. Fora J, =
1+ J,= 1 transition, it has been shown [57] that there is a dark state which is
described by the same vector field as the laser field. More precisely, if the la-
ser field is formed by a linear superposition of N plane waves with wave vec-
torsk, (i= 1,2,...N) having the same modulus k, one finds that atoms are cool-
ed in a coherent superposition of N wave packets with mean momenta fik,
and with a momentum spread 6p which becomes smaller and smaller as the
imteraction time @ increases. Furthermore, because of the isomorphism
between the de Broglie dark state and the laser field, one can adiabatically
change the laser configuration and transfer the whole atomic population in-
toa single wave packet or two coherent wave packets chosen at will [58]. Fig.
9 shows an example of such a coherent manipulation of atomic wave packets
in two dimensions. In Fig.9a, one sees the transverse velocity distribution182 Physics 1997
associated with the four wave packets obtained with two pairs of counterpro-
pagating laser waves along the x and y-axis in a horizontal plane; Fig.9b shows
the single wave packet into which the whole atomic population is transferred
by switching off adiabatically three of the four VSCPT beams. Similar results
have been obtained in three dimensions.
3.3 SUBRECOIL LASER COOLING AND LEVY STATISTICS
Quantum Monte Garlo simulations using the delay function [59, 60] have
provided new physical insight into subrecoil laser cooling [61]. Fig.10 shows
for example the random evolution of the momentum of an atom in a 1D-
VSCPT experiment. Each vertical discontinuity corresponds to a spontaneous
emission process during which p changes in a random way. Between two suc-
cessive jumps, p remains constant. It clearly appears that the random walk of
the atom in velocity space is anomalous and dominated by a few rare events
whose duration is a significant fraction of the total interaction time. A simple
analysis shows that the distribution P(t) of the trapping times 7 in a small
trapping zone near v = 0 is a broad distribution which falls as a power-law in
the wings. These wings decrease so slowly that the average value (7) of 7 (or
the variance) can diverge. In such cases, the central limit theorem (CLT) can
obviously no longer be used for studying the distribution of the total trapping
time after N entries in the trapping zone separated by Nexits.
Toor yy
eo
i
n
3
°
Momentum (hk)
°
-10 4
-20 4
. ea q
_30 Time(I"*) 4
Sr
-0 200000 400000 600000
Figure 10, Monte Carlo wave function simulation of one-dimensional VSCPT (from reference
[61]). Momentum p characterizing the cooled atoms versus time. Each vertical discontinuity cor-
responds to a spontaneous emission jump during which changes in a random way. Between two
successive jumps, p remains constant. The inset shows a zoomed part of the sequence.Claude N. Cohen-Tannoudji 183
It is possible to extend the CLT to broad distributions with power-law wings
[62, 63]. We have applied the corresponding statistics, called “Lévy statistics”,
to subrecoil cooling and shown that one can obtain in this way a better un-
derstanding of the physical processes as well as quantitative analytical predic-
tions for the asymptotic properties of the cooled atoms in the limit when the
interaction time 6 tends to infinity (61, 64]. For example, one predicts in this
way that the temperature decreases as 1/@ when @ — ©, and that the wings
of the momentum distribution decrease as 1/#2, which shows that the shape
of the momentum distribution is closer to a Lorentzian than a Gaussian. This
is in agreement with the experimental observations represented in Fig.8.
(The fit in Fig.8a is an exponential, which is the Fourier transform of a
Lorentzian).
‘One important feature revealed by this theoretical analysis is the non-ergo-
dicity of the cooling process. Regardless of the interaction time @, there are
always atomic evolution times (trapping times in the small zone of Fig.5a
around v= 0) which can be longer than 6. Another advantage of such a new
approach is that it allows the parameters of the cooling lasers to be optimized
for given experimental conditions. For example, by using different shapes for
the laser pulses used in one-dimensional subrecoil Raman cooling, it has
been possible to reach for Cesium atoms temperatures as low as 3 nK [65].
4, AFEW EXAMPLES OF APPLICATIONS
The possibility of trapping atoms and cooling them at very low temperatures,
where their velocity can be as low as a few mm/s, has opened the way to a
wealth of applications. Ultracold atoms can be observed during much longer
times, which is important for high resolution spectroscopy and frequency
standards. They also have very long de Broglie wavelengths, which has given
rise to new research fields, such as atom optics, atom interferometry and
Bose-Einstein condensation of dilute gases. It is impossible to discuss here all
these developments. We refer the reader to recent reviews such as [5]. We will
just describe in this section a few examples of applications which have been
recently investigated by our group in Paris.
4.1 CESIUM ATOMIC CLOCKS
Cesium atoms cooled by Sisyphus cooling have an effective temperature on
the order of 1 4K, corresponding to a rms. velocity of 1 cm/s. This allows
them to spend a longer time Tin an observation zone where a microwave
field induces resonant transitions between the two hyperfine levels g, and g,
of the ground state. Increasing T decreases the width Av ~ 1/T of the micro-
wave resonance line whose frequency is used to define the unit of time. The
stability of atomic clocks can thus be considerably improved by using ultra-
cold atoms [66. 67].
In usual atomic clocks, atoms from a thermal cesium beam cross two
microwave cavities fed by the same oscillator. The average velocity of the
atoms is several hundred m/s, the distance between the two cavities is on the184 Physics 1997
order of 1 m. The microwave resonance between g, and g, is monitored and
is used to lock the frequency of the oscillator to the center of the atomic line.
The narrower the resonance line, the more stable the atomic clock. In fact,
the microwave resonance line exhibits Ramsey interference fringes whose
width Avis determined by the time of flight Tof the atoms from one cavity to
another. For the longest devices, T, which can be considered as the observation
time, can reach 10 ms, leading to values of Av~ 1/T on the order of 100 Hz.
Much narrower Ramsey fringes, with sub-Hertz linewidths can be obtained
in the so-called “Zacharias atomic fountain” [68]. Atoms are captured in a
magneto-optical trap and laser cooled before being launched upwards by a la-
ser pulse through a microwave cavity. Because of gravity they are decelerated,
they return and fall back, passing a second time through the cavity. Atoms
therefore experience two coherent microwave pulses, when they pass through
the cavity, the first time on their way up, the second time on their way down.
The time interval between the two pulses can now be on the order of 1 sec,
i.e, about two orders of magnitude longer than with usual clocks. Atomic
fountains have been realized for sodium [69] and cesium [70]. A short-term
relative frequency stability of 1.3 x 10-!°r ~'/2, where Tis the integration time,
has been recently measured for a one meter high Cesium fountain [71, 72].
For T= 104s, Av/v ~ 1.3 x 10°! and for t= 3 x 104s, Av/v~ 8 x 10-'6 has been.
measured, In fact such a stability is most likely limited by the Hydrogen maser
which is used as a reference source and the real stability, which could be
more precisely determined by beating the signals of two fountain clocks, is ex-
pected to reach Av/v~ 10-16 for a one day integration time. In addition to the
stability, another very important property of a frequency standard is its ac-
curacy. Because of the very low velocities in a fountain device, many system-
atic shifts are strongly reduced and can be evaluated with great precision.
With an accuracy of 2 x 10718, the BNM-LPTF fountain is presently the most
accurate primary standard [73]. A factor 10 improvement in this accuracy is
expected in the near future.
To increase the observation time beyond one second, a possible solution
consists of building a clock for operation in a reduced gravity environment.
Such a microgravity clock has been recently tested in a jet plane making
parabolic flights. A resonance signal with a width of 7 Hz has been recorded
in a 10-%g environment. This width is twice narrower than that produced on
earth in the same apparatus. This clock prototype (see Fig.11) is a compact
and transportable device which can be also used on earth for high precision
frequency comparison.
Atomic clocks working with ultracold atoms could not only provide an im-
provement of positioning systems such as the GPS. They could be also used
for fundamental studies. For example, one could build two fountains clocks,
‘one with cesium and one with rubidium, in order to measure with a high ac-
curacy the ratio between the hyperfine frequencies of these two atoms.
Because of relativistic corrections, the hyperfine splitting is a function of Za.
where qs the fine structure constant and Zis the atomic number [74]. Since
Zis not the same for cesium and rubidium, the ratio of the two hyperfineClaude N. Cohen-Tannoudji 185
Figure 11. The microgravity clock prototype. The left part is the 60 cm x 60 em x 15 cm optical
bench containing the diode laser sources and the various optical components. The right part is
the clock itself (about one meter long) containing the optical molasses, the microwave cavity and
the detection region.
structures depends on @. By making several measurements of this ratio over
long periods of time, one could check Dirac’s suggestion concerning a possi-
ble variation of o with time. The present upper limit for6./at in laboratory
tests [74] could be improved by two orders of magnitude.
Another interesting test would be to measure with a higher accuracy the
gravitational red shift and the gravitational delay of an electromagnetic wave
passing near a large mass (Shapiro effect [75]).
4.2 GRAVITATIONAL CAVITIES FOR NEUTRAL ATOMS
We have already mentioned in section 1.3 the possibility of making atomic
mirrors for atoms by using blue detuned evanescent waves at the surface of a
piece of glass. Concave mirrors (Fig.12a) are particularly interesting because
the transverse atomic motion is then stable if atoms are released from a point
located below the focus of the mirror. It has been possible in this way to ob-
serve several successive bounces of the atoms (Fig-12b) and such a system can
be considered as a “trampoline for atoms” [37]. In such an experiment, it is a
good approximation to consider atoms as classical particles bouncing off a
concave mirror. In a quantum mechanical description of the experiment, one
must consider the reflection of the atomic de Broglie waves by the mirror.
Standing de Broglie waves can then be introduced for such a “gravitational
cavity”, which are quite analogous to the light standing waves for a Fabry-
Perot cavity [76]. By modulating at frequency 0/27 the intensity of the eva-
nescent wave which forms the atomic mirror, one can produce the equivalent
of a vibrating mirror for de Broglie waves. The reflected waves thus
have a modulated Doppler shift. The corresponding frequency modulation186, Physics 1997
of these waves has been recently demonstrated [77] by measuring the energy
change AE of the bouncing atom, which is found to be equal to 7Q, where n
= 0,4 142, ... (Figs.12c and d). The discrete nature of this energy spectrum
isa pure quantum effect. For classical particles bouncing off a vibrating mir-
ror, AE would vary continuously in a certain range.
107
%
%
Number of atoms
a
0 0102 03 04 05
Bouncing time (s)
(a) (b)
Probe
absorption
(arb. un.)
time
06 10-15 20
(ce) (d)
Figure 12. Gravitational cavity for neutral atoms (from references [37] and [77]). Trampoline for
atoms (a) - Atoms released from a magneto-optical trap bounce off a concave mirror formed by
‘a blue detuned evanescent wave at the surface of a curved glass prism. Number of atoms at the
initial position of the trap versus time after the trap has been switched off (b). Ten successive
bounces are visible in the figure. Principle of the experiment demonstrating the frequency
modulation of de Broglie waves (c). The upper trace gives the atomic trajectories (vertical posi-
tion zversus time). The lower trace gives the time dependence of the intensity of the evanescent
wave, The first pulse is used for making a velocity selection. The second pulse is modulated in in-
tensity. This produces a vibrating mirror giving rise to a frequency modulated reflected de
Broglie wave which consists in a carrier and sidebands at the modulation frequency. The energy
spectrum of the reflected particles is thus discrete so that the trajectories of the reflected partic-
les form a discrete set. This effect is detected by looking at the time dependence of the absorp-
tion ofa probe beam located above the prism (d).Claude N. Cohen-Tannoudji 187
4.3 BLOCH OSCILLATIONS
In the subrecoil regime where 6p becomes smaller than fk, the atomic co-
herence length h/&p becomes larger than the optical wavelength A= h/hk =
2m /k of the lasers used to cool the atom. Consider then such an ultracold
atom in the periodic light shift potential produced by a non resonant laser
standing wave. The atomic de Broglie wave is delocalized over several periods
of the periodic potential, which means that one can prepare in this way
quasi-Bloch states. By chirping the frequency of the two counterpropagating
laser waves forming the standing wave, one can produce an accelerated stand-
ing wave. In the rest frame of this wave, atoms thus feel a constant inertial
force in addition to the periodic potential. They are accelerated and the de
Broglie wavelength Ay, = h/M(v) decreases. When Ayp = 4,,,., the de Broglie
wave is Bragg-reflected by the periodic optical potential. Instead of increasing
linearly with time, the mean velocity (x) of the atoms oscillates back and forth.
Such Bloch oscillations, which are a textbook effect of solid-state physics, are
more easily observed with ultracold atoms than with electrons in condensed
matter because the Bloch period can be much shorter than the relaxation
time for the coherence of de Broglie waves (in condensed matter, the relaxa-
tion processes due to collisions are very strong). Fig.13 shows an example of
Bloch oscillations [78] observed on cesium atoms cooled by the improved
subrecoil Raman cooling technique described in [65].
9
a
=
3
=
e OQ
S
&S
=-05
~2 -1 QO 1 2
Time
Figure 13. Bloch oscillations of atonis in a periodic optical potential (from reference [78]. Mean
velocity (in units of the recoil velocity) versus time (in units of half the Bloch period) for ultra-
cold cesium atoms moving in a periodic optical potential and submitted in addition to a constant
force.
5, CONCLUSION
We have described in this paper a few physical mechanisms allowing one to
manipulate neutral atoms with laser light. Several of these mechanisms can
be simply interpreted in terms of resonant exchanges of energy, angular and
linear momentum between atoms and photons. A few of them, among the188 Physics 1997
most efficient ones, result from a new way of combining well known physical
effects such as optical pumping, light shifts, coherent population trapping.
We have given two examples of such cooling mechanisms, Sisyphus cooling
and subrecoil cooling, which allow atoms to be cooled in the microKelvin and
nanoKelvin ranges. A few possible applications of ultracold atoms have been
also reviewed. They take advantage of the long interaction times and long de
Broglie wavelengths which are now available with laser cooling and trapping
techniques.
One can reasonably expect that further progress in this field will be
made in the near future and that new applications will be found. Concerning
fundamental problems, two directions of research at least look promising.
First, a better control of “pure” situations involving a small number of atoms
in well-defined states exhibiting quantum features such as very long spatial
coherence lengths or entanglement. In that perspective, atomic, molecular
and optical physics will continue to play an important role by providing a
“testing bench” for improving our understanding of quantum phenomena. A
second interesting direction is the investigation of new systems, such as Bose
condensates involving a macroscopic number of atoms in the same quantum
state. One can reasonably hope that new types of coherent atomic sources
(sometimes called “atom lasers”) will be realized, opening the way to interest-
ing new possibilities.
Itis clear finally that all the developments which have occurred in the field
of laser cooling and trapping are strengthening the connections which can be
established between atomic physics and other branches of physics, such as
condensed matter or statistical physics. The use of Lévy statistics for analyzing
subrecoil cooling is an example of such a fruitful dialogue. The interdiscipli-
nary character of the present researches on the properties of Bose conden-
sates is also a clear sign of the increase of these exchanges.
REFERENCES,
[1] A Kastler, J. Phys. Rad. 11, 255 (1950).
[2] A. Ashkin, Science, 210, 1081 (1980).
[3] V.S. Letokhov and V.G. Minogin, Phys.Reports, 73, 1 (1981).
[4] S. Stenholm, Rev.Mod.Phys. 58, 699 (1986)
[5] CS. Adams and E. Riis, Prog. Quant. Electr. 21, 1 (1997).
[6] C. Cohen-Tannoudji and W. Phillips, Physics Today 43, No 10, 33 (1990).
[7] W. Heitler, “The quantum theory of radiation”, 3rd ed. (Clarendon Press, Oxford,
1954)
[8] JP. Barrat and C,Cohen-Tannoudji, J.Phys.Rad. 22, 329 and 443 (1961).
[9] C.Cohen-Tannougji , Ann. Phys. Paris 7, 423 and 469 (1962).
[10] C. Cohen-Tannoudji, J. Dupont Roc and G. Grynberg, “Atom-photon interactions
Basic processes and applications”, (Wiley, New York, 1992).
[11] SH. Autler and C.H. Townes, Phys.Rev. 100, 703 (1955).
[12] C. Cohen-Tannoudji, C.RAcad Sci. (Fr) 252, 394 (1961).
[13] C. Cohen-Tannoudji and J. Dupont Roc, Phys.Rev. A5, 968 (1972).
[14] W.D.Phillips and H.Metcalf, Phys.Rev.Lett. 48, 596 (1982).
[15] J.VProdan, WD Phillips and H.Metcalf, Phys.Rev.Lett. 49, 1149 (1982).
[16] W. Ertmer, R. Blatt, J.L. Hall and M. Zhu, Phys.Rev.Lett. 54, 996 (1985).a7)
(18)
19]
[20]
[21]
[22]
[23]
[24]
[25]
[26]
[27]
(28)
[29}
[30}
[31
[32]
[33]
[34]
(35)
[36]
(37)
(38)
39]
(40}
(4)
(42)
(431
(44)
(45)
[46]
[47]
(48)
[491
(50]
(51)
Claude N. Cohen-Tannoudji 189
G.A. Askarian, Zh.Eksp.TeorFiz, 42, 1567 (1962) (SovPhysJETP, 15, 1088 (1962)].
C. Cohen-Tannoudji, “Atomic motion in laser light”, in “Fundamental systems in
quantum optics”, J. Dalibard, J.-M. Raimond, and J. Zinn-Justin eds, Les Houches ses
sion LIII (1990), (North-Holland, Amsterdam 1992) p.1
J. Dalibard and C. Cohen-Tannoudji, J.Opt.Soc.Am. B2, 1707 (1985).
‘.W. Hansch and A.L. Schawlow, Opt. Commun. 18, 68 (1975).
D,Wineland and H.Dehmelt, Bull.Am.PhysSoc. 20, 637 (1975)
S. Chu, L. Hollberg, J.E. Bjorkholm, A. Cable and A. Ashkin, Phys. Rev. Lett. 55, 48
(1985).
VS. Letokhov, V.G. Minogin and B.D. Pavlik, Zh.Eksp.Teor-Fiz. 72, 1828 (1977)
[Sov.Phys JETP, 45, 698 (1977)].
DJ. Wineland and W. Itano, Phys.Rev. A20, 1521 (1979).
J.P. Gordon and A. Ashkin, Phys.Rev. A21, 1606 (1980).
PD. Lett, RN. Watts, C.1. Westbrook, W. Phillips, P.L. Gould and HJ. Metcalf, Phys.
Rev. Lett. 61, 169 (1988).
ELL. Raab, M. Prentiss, A. Cable, S. Chu and D-E. Pritchard, Phys.Rev.Lett. 59, 2631
(1987).
C. Monroe, W. Swann, H. Robinson and C.E. Wieman, PhysRevLett. 65, 1571
(1990).
S. Chu, J.E. Bjorkholm, A. Ashkin and A. Cable, Phys.Rev.Lett. 57, 314 (1986).
C.S. Adams, HJ. Lee, N. Davidson, M. Kasevich and S. Chu, Phys. Rev. Lett. 74, 3577
(1995).
A. Kuhn, H. Perrin, W. Hansel and C. Salomon, OSA TOPS on Ultracold Atoms and
BEG, 1996, Vol. 7, p.58, Keith Burnett (ed.), (Optical Society of America, 1997).
VS. Letokhov, Pis'ma.Eksp.Teor Fiz. 7, 348 (1968) (JETP Lett, 7, 272 (1968) ].
C. Salomon, J. Dalibard, A. Aspect, H. Metcalf and C. Cohen-Tannoudji, Phys. Rev.
Lett. 59, 1659 (1987).
RJ. Cook and RK. Hill, Opt. Commun. 43, 258 (1982)
VIL Balykin, V.S. Letokhoy, Yu. B. Ovchinnikov and A.I. Sidorov, Phys. Rev. Lett. 60,
2137 (1988)
M.A. Kasevich, D.S. Weiss and S. Chu, Opt. Lett. 15, 607 (1990).
C.G. Aminoff, AM. Steane, P. Bouyer, P. Desbiolles, J. Dalibard and C. Cohen-
‘Tannoudji, Phys. Rev. Lett. 71, 3083 (1993).
J. Dalibard and C. Cohen-Tannoudji, J. Opt. Soc. Am. B6, 2023 (1989).
PJ. Ungar, D.S. Weiss, E. Riis and S. Chu, JOSA B6, 2058 (1989).
C. Salomon, J. Dalibard, W. Phillips, A. Clairon and S. Guellati, Europhys. Lett. 12,
683 (1990)
Y. Castin, These de doctorat, Paris (1991).
Y. Castin and K. Molmer, Phys. Rev. Lett. 74, 3772 (1995).
Y. Castin and J. Dalibard, Europhys.Lett. 14, 761 (1991).
P, Verkerk, B. Lounis, C. Salomon, C. Cohen-Tannoudji,
Grynberg, Phys. Rev. Lett. 68, 3861 (1992)
PS. Jessen, C. Gerz, PD. Lett, W.D. Phillips, S.L. Rolston, RJ.C. Spreeuw and C1.
Westbrook, Phys. Rev. Lett. 69, 49 (1992).
G. Grynberg and C. Triché, in Proceedings of the International School of Physics
“Enrico Fermi”, Course CXXXI, A. Aspect, W. Barletta and R. Bonifacio (Eds), p.243,
IOS Press, Amsterdam (1996); A. Hemmerich, M. Weidemiiller and T.W. Hansch,
same Proceedings, p.503.
PS. Jessen and LH. Deutsch, in Advances in Atomic, Molecular and Optical Physics,
37, 95 (1996), ed. by B. Bederson and H. Walther.
‘A Aspect, E. Arimondo, R. Kaiser, N. Vansteenkiste, and C. Cohen-Tannoudji, Phys.
Rev. Lett. 61, 826 (1988).
M. Kasevich and S. Chu, Phys. Rev. Lett. 69, 1741 (1992).
Alzetta G., Gozzini A., Moi L., Orriols G., Il Nuovo Cimento 36B, 5 (1976).
Arimondo E., Orriols G., Lett. Nuovo Cimento 17, 383 (1976).
|-¥. Courtois and G.190 Physics 1997
[52] A. Aspect, E. Arimondo, R. Kaiser, N. Vansteenkiste, and C. Cohen-Tannoudji, J. Opt.
Soc. Am. B6, 2112 (1989).
[53] F. Bardou, B. Saubamea, J. Lawall, K. Shimizu, O. Emile, C. Westbrook, A. Aspect, C.
Cohen-Tannoudji, C. R. Acad. Sci. Paris 318, 87-885 (1994)
[54] B. Saubamea, T.W. Hijmans, S. Kulin, E. Rasel, E. Peik, M. Leduc and C. Cohen-
Tannoudji, Phys.Rev.Lett. 79, 3146 (1997)
[55] J. Lawall, F. Bardou, B. Saubamea, K. Shimizu, M. Leduc, A. Aspect and C. Cohen-
‘Tannoudji, Phys. Rev. Lett. 73, 1915 (1994).
[56] J. Lawall, S. Kulin, B. Saubamea, N. Bigelow, M. Leduc and C. Cohen-Tannoudji,
Phys. Rev. Lett. 75, 4194 (1995).
[57] M.A. Ol’shanii and V.G. Minogin, Opt. Commun. 89, 393 (1992).
[58] S. Kulin, B. Saubamea, E. Peik, J. Lawall, T:W. Hijmans, M. Leduc and G .Cohen-
‘Tannoudji, Phys. Rev. Lett. 78, 4185 (1997).
[59] C. Cohen-Tannoudji and J. Dalibard, Europhys. Lett. 1, 441 (1986).
[60] P. Zoller, M. Marte and D.F. Walls, Phys. Rev. A35, 198 (1987).
(61) F. Bardou, J-P. Bouchaud, O. Emile, A. Aspect and C. Cohen-Tannoudji, Phys. Rev.
Lett. 72, 203 (1994).
[62] B.V. Gnedenko and A.N. Kolmogorov, “Limit distributions for sum of independent
random variables” ( Addison Wesley, Reading, MA, 1954).
[63] J.P. Bouchaud and A. Georges, Phys. Rep. 195, 127 (1990).
[64] F. Bardou, Ph. D. Thesis, University of Paris XI, Orsay (1995)
[65] J. Reichel, F. Bardou, M. Ben Dahan, E. Peik, S. Rand, C. Salomon and C. Cohen-
‘Tannoudji, Phys. Rev. Lett. 75, 4575 (1995).
[66] K Gibble and S. Chu, Metrologia, 29, 201 (1992)
[67] S.N. Lea, A. Clairon, C. Salomon, P. Laurent, B. Lounis, J. Reichel, A. Nadir, and G.
Santarelli, Physica Scripta T51, 78 (1994)
[68] J. Zacharias, Phys.Rev. 94, 751 (1954). See also : N. Ramsey, Molecular Beams, Oxford
University Press, Oxford, 1956.
[69] M. Kasevich, E. Riis, S. Chu and R. de Voe, Phys. Rev. Lett. 63, 612 (1989)
[70] A. Clairon, C. Salomon, S. Guellati and W.D. Phillips, Europhys. Lett. 16, 165 (1991).
[71] S. Ghezali, Ph. Laurent, $.N. Lea and A. Clairon, Europhys. Lett. 36, 25 (1996)
[72] S. Ghezali, Thése de doctorat, Paris (1997).
[73] E. Simon, P. Laurent, C. Mandache and A. Clairon, Proceedings of EFTF 1997,
Neuchatel, Switzerland.
[74] J. Prestage, R. Tjoelker and L. Maleki, Phys. Rev. Lett. 74, 3511 (1995).
[75] LShapiro, Phys. Rev. Lett. 18, 789 (1964).
[76] H. Wallis, J. Dalibard and C. Cohen-Tannoudji, Appl. Phys. B54, 407 (1992).
(77] A. Steane, P. Szriftgiser, P. Desbiolles and J. Dalibard, Phys. Rev. Lett. 74, 4972 (1995).
[78] M. Ben Dahan, E. Peik, J. Reichel, Y. Castin and C. Salomon, Phys. Rev. Lett. 76, 4508
(1996).