appliedphysicsB manuscript No.
(will be inserted by the editor)
arXiv:cond-mat/0201270v1 16 Jan 2002
Phase Fluctuations in Bose-Einstein Condensates
D. Hellweg1 , S. Dettmer1 , P. Ryytty1 , J. J. Arlt1 , W. Ertmer1 , K. Sengstock1 , D. S. Petrov2,3 , G. V.
Shlyapnikov2,3,4 , H. Kreutzmann5 , L. Santos5 , and M. Lewenstein5
1
2
3
4
5
Institut für Quantenoptik, Universität Hannover, Welfengarten 1, 30167 Hannover, Germany
FOM Institute for Atomic and Molecular Physics, Kruislaan 407, 1098 SJ Amsterdam, The Netherlands
Russian Research Center Kurchatov Institute, Kurchatov Square, 123182 Moscow, Russia
Laboratoire Kastler Brossel, Ecole Normale Supérieure, 24 rue Lhomond, 75231 Paris Cedex 05, France
Institut für Theoretische Physik, Universität Hannover, Appelstraße 2, 30167 Hannover, Germany
The date of receipt and acceptance will be inserted by the editor
Abstract We demonstrate the existence of phase fluctuations in elongated Bose-Einstein Condensates (BECs)
and study the dependence of those fluctuations on the
system parameters. A strong dependence on temperature, atom number, and trapping geometry is observed.
Phase fluctuations directly affect the coherence properties of BECs. In particular, we observe instances where
the phase coherence length is significantly smaller than
the condensate size. Our method of detecting phase fluctuations is based on their transformation into density
modulations after ballistic expansion. An analytic theory describing this transformation is developed.
PACS: 03.75.Fi, 32.80.Pj, 05.30.Jp
1 Introduction
Bose-Einstein condensates of weakly interacting gases,
such as alkali atom vapours, constitute very well controllable macroscopic quantum systems. For extremely low
temperatures (T → 0) the condensate is well described
by an effective macroscopic single particle wavefunction,
occupied by millions of atoms [1]. The corresponding
macroscopic phase is related to many fascinating properties of BEC such as its coherence, superfluidity, and
effects known from superconductivity as the Josephson
effect. In particular, the coherence properties are essential for very promising applications of BEC such as matter wave interferometry [2,3,4,5,6,7,8] or atom lasers
[9] which rely on BEC as a source of coherent matter
waves. The phase coherence of condensates was shown
by imaging the interference pattern of two independent
condensates which were brought to overlap [2]. Also using an interference technique it has been found that a
trapped BEC has a uniform spatial phase [3] and therefore the coherence length is just limited by the size of
the condensate. This result has also been obtained by
measuring the momentum distribution in the radial direction of a cigar shaped condensate by a spectroscopic
technique [10]. These experiments have focused on the
coherence properties of almost pure condensates, prepared at temperatures well below the critical temperature. The temperature dependence of the coherence of
a BEC was studied in a ’double slit’ experiment where
an atom laser beam was extracted from a cigar shaped
BEC at different radial positions [4]. A reduction of interference fringe visibility was observed for increasing
temperature with the interference pattern being reproducible and thus indicating that the relative phase of the
condensate fraction was not fluctuating randomly.
Even though the uniform spatial phase of a condensate has been confirmed in several experiments, it is not
an obvious or even general property of BEC at finite
temperature. The ’fragmentation’ into independent condensates with randomly fluctuating relative phase was
often discussed in the context of the nucleation process
of a condensate [11]. There, the equilibrium state of the
system was assumed to be a ’pure’ condensate without
phase fluctuations. However, before the condensate is
completely formed phase fluctuations between different
regions of the condensate are expected even though density fluctuations are suppressed. But also for the equilibrium state of a quantum system it is expected that
low-dimensional (1D and 2D) quantum gases differ qualitatively from the 3D case with respect to statistical and
phase correlation properties [12,13,14]. Low-dimensional
quantum degenerate gases, where the trap energy-level
spacing is larger than the interaction energy between
atoms, have been experimentally realized in atomic hydrogen [15], Sodium [16], and Lithium [17] but phase
correlation properties have not yet been experimentally
investigated. Recently, it was shown theoretically [18]
that for very elongated condensates phase fluctuations
can be pronounced already in the equilibrium state of
the usual 3D ensemble, where density fluctuations are
suppressed. This was experimentally and theoretically
2
investigated in [19]. The phase coherence length in this
case can be much smaller than the axial size of the sample.
In this paper we present detailed experimental and
theoretical studies of phase fluctuations in elongated BEC.
The dependence on the trapping geometry, temperature,
and the number of condensed atoms is investigated. We
demonstrate that the spatial phase of the condensates
undergoes random fluctuations with an average value
determined by the system parameters. In particular, our
measurements show that the average phase fluctuations
increase with tighter radial confinement. Thus, they are
especially pronounced for very elongated geometries as
are used to study the transition from 3D to 1D degenerate quantum gases [16,19].
BEC has recently been achieved in elongated microcircuit geometries in which radial trapping frequencies of
tens of KHz or even MHz are possible [20]. Since this is
more than an order of magnitude bigger than the radial
frequencies at which we have observed phase fluctuations our results should be especially relevant for these
systems. Phase fluctuations should also be considered if
tight confinement waveguides are used for BECs [8] or
guided atom lasers beams, even though our results directly apply only to the equilibrium state of the system.
Note, that phase fluctuations lead to a broadening of
the momentum distribution, which is the Fourier transform of the single particle correlation function, defining
the coherence length. Therefore, only in the absence of
phase fluctuations, a BEC constitutes a state where ultimate control over the motion and position of atoms, limited only by Heisenberg’s uncertainty relation, is achieved.
Fluctuations of the phase of a Bose condensate are
due to thermal excitations and always appear at finite
temperature. Their experimental characterization constitutes a test of many-body theories at finite temperature. If the wavelength of these excitations is smaller
than all dimensions of the condensate they have a 3D
character and do not lead to pronounced phase fluctuations. In elongated condensates however, low energy axial excitations have wavelengths larger than the radial
size of the sample and therefore acquire a pronounced 1D
behaviour leading to axial phase fluctuations, although
density fluctuations of the equilibrium state are still suppressed. The coherence properties can be significantly
altered as compared with previous observations. In particular, the coherence length, i.e. the distance at which
the single particle correlation function falls to e−1 , can
be much smaller than the axial size of the condensate.
This is not in contradiction to previous coherence measurements since they were performed in rather spherical
traps [3], in the radial direction of cigar shaped condensates [4,10], and at low temperature [3,10].
Our method to study phase fluctuations is based on
ballistic expansion. We show that the original phase distribution is mapped into the density distribution during
time-of-flight. The density modulations after ballistic ex-
D. Hellweg et al.
pansion are a direct measure of the original phase fluctuations of the trapped condensate.
The paper is organized as follows. It starts with a
theoretical discussion of phase fluctuations including a
detailed explanation of our analytic theory [19] for the
appearance of density modulations in the ballistic expansion. Then we present our experimental studies demonstrating the existence of phase fluctuations in BECs and
investigating their dependence on system parameters.
Finally, alternative methods to study phase fluctuations
are discussed, including simulations of the interference
of phase fluctuating condensates.
2 Theoretical description
2.1 Phase fluctuating condensates in elongated 3D traps
In a standard equilibrium situation in 3D traps the fluctuations of density and phase of the Bose-Einstein condensate are only important in a narrow temperature
range near the BEC transition temperature Tc (critical fluctuations). Outside this region, the fluctuations
are suppressed and the condensate is phase coherent.
This picture precludes the interesting physics of phasefluctuating condensates, which is present in 2D and 1D
systems (see [12,13] and refs. therein).
In the Thomas-Fermi regime (see e.g. [1]), when the
nonlinear mean field interaction energy dominates the
kinetic energy, the situation can change for elongated
BECs [18]. While density fluctuations in equilibrium will
remain suppressed due to their energetic cost, the situation can be different for the phase fluctuations. In particular, the axial phase fluctuations can manifest themselves even at temperatures far below Tc . Then, as the
density fluctuations are suppressed, the equilibrium state
of the system becomes a condensate with fluctuating phase
(quasicondensate) similar to that in 1D trapped gases
[12]. Decreasing T gradually reduces the phase fluctuations.
2.2 Description of the fluctuating phase
In this section we describe the phase fluctuations along
the lines of Ref. [18]. Let us consider a BEC at T = 0
in the Thomas-Fermi regime, where the mean-field (repulsive) interparticle interaction greatly exceeds the radial (ωρ ) and axial (ωx ) trap frequencies. The density
profile of the zero-temperature condensate has the wellknown shape n0 (ρ, x) = n0m (1 − ρ2 /R2 − x2 /L2 ), where
n0m = µ/g is the maximum condensate density, with
µ being the chemical potential, g = 4πh̄2 a/m, m the
atom mass, and a > 0 the scattering length. Under the
condition ωρ ≫ ωx , the radial size of the condensate,
R = (2µ/mωρ2 )1/2 , is much smaller than the axial size
L = (2µ/mωx2 )1/2 .
Phase Fluctuations in Bose-Einstein Condensates
3
with δ φ̂(r, r′ ) = φ̂(r) − φ̂(r′ ). The operator φ̂(r) is given
by (see, e.g., [23])
X
fj+ (r)âj + h.c.,
(2)
φ̂(r) = [4n0 (r)]−1/2
j
where âj is the annihilation operator of the excitation
with quantum number(s) j and energy ǫj , fj+ = uj + vj ,
and the u, v functions of the excitations are determined
by the Bogoliubov-de Gennes equations.
The “low-energy” axial excitations (with energies ǫj <
h̄ωρ ) have wavelengths larger than R and exhibit a pronounced 1D behavior. Hence, one expects that these excitations give the most important contribution to the
long-wave axial fluctuations of the phase. The solution of
the Bogoliubov-de Gennes equations for such
plow-energy
axial modes gives the spectrum ǫj = h̄ωx j(j + 3)/4
[21], where j is a positive integer. The wavefunctions fj+
of these modes have the form
s
(j + 2)(2j + 3)gn0 (r) (1,1) x
+
,
(3)
Pj
fj (r) =
4π(j + 1)R2 Lǫj
L
(1,1)
are Jacobi polynomials. Note that the conwhere Pj
tribution of the low-energy axial excitations to the phase
operator (2) is independent of the radial coordinate ρ.
2.3 Numerical simulations
In order to simulate numerically the effect of phase fluctuations during the ballistic expansion of the condensate we replace the operators âj and â†j in Eq. (2) by
complex Gaussian random variables αj and α∗j , with the
correlation hαj α∗j ′ i = δjj ′ Nj , where Nj is the occupation
number for the quasiparticle mode j for a given (small)
temperature T . Such a random phase reproduces correctly the phase correlations, which for not too large
|x − x′ |/L ≤ 0.4, behave as
2
|x − x′ |/L,
h[δ φ̂(x, x′ )]2 iT = δL
(4)
2
where the quantity δL
is given by
2
δL
(T ) = 32µT /15N0(h̄ωx )2 ,
(5)
2
and lφ = L/δL
can be interpreted as a phase coherence
length. It is the distance at which the phase factor of the
2p
1p
0p
-1p
-2p
f [rad]
The phase fluctuations can be described by solving
the Bogoliubov-de Gennes equations [21] describing elementary excitations of the condensate.p
One can write the
total field operator of atoms as ψ̂(r) = n0 (r) exp(iφ̂(r)),
where φ̂(r) is the operator of the phase, and the density
fluctuations have been already neglected following the
arguments discussed above. The single-particle correlation function is then expressed through the mean square
fluctuations of the phase (see, e.g. [22]):
p
hψ̂ † (r)ψ̂(r′ )i = n0 (r)n0 (r′ ) exp{−h[δ φ̂(r, r′ )]2 i/2}, (1)
2p
1p
0p
-1p
-2p
2p
1p
0p
-1p
-2p
-0.8
-0.4
0.0
0.4
0.8
x/L
Fig. 1 Typical phase patterns for three different trap aspect ratios: λ = 10 (top), λ = 100 (middle), and λ = 1000
(bottom). For all cases ωx = 2π × 14 Hz, T = 0.6 Tc , and
N = 2 × 105 .
single-particle correlation function (Eq. 1) falls to its 1/e
value. The condition lφ /L = 1 determines then a characteristic temperature Tφ = 15(h̄ωx )2 N0 /32µ, where N0
is the number of condensed atoms. For Tφ < Tc , which
is the case for most of our measurements, one expects
the regime of quasicondensation for the initial cloud in
the temperature interval Tφ < T < Tc [18].
Figure 1 shows examples of simulated phase distributions for various trap geometries parameterized by the
ratio of trapping frequencies λ = ωρ /ωx . Whereas for a
trap with λ = 10 the fluctuations are rather suppressed,
the phase of the BEC fluctuates for a trap with λ = 1000
by more than 2π. This dramatically changes the properties of the condensate, especially in phase sensitive experiments.
The appearance of stripes in the process of ballistic expansion can be understood qualitatively as follows.
As mentioned above, within the equilibrium state of a
BEC in a magnetic trap the density distribution remains
largely unaffected even if the phase fluctuates [18]. The
reason is that the mean-field interparticle interaction
prevents the transformation of local velocity fields provided by the phase fluctuations into modulations of the
density. However, after switching off the external trap,
the mean-field interaction rapidly decreases and the axial
velocity field is then converted into a particular density
distribution. We have performed numerical simulations
4
D. Hellweg et al.
of the 3D Gross-Pitaevskii equation (GPE) to understand quantitatively how phase fluctuations lead to the
build up of stripes in the density distribution. We assumed that initially (just before opening of the trap) the
condensate had an equilibrium density profile, which has
been calculated using the standard imaginary evolution
of the GP equation. We have imprinted on it a random
fluctuating phase φ(x), as described above, and evolved
the condensate in free space.
[21]), we end up with the equation:
¨ −
δn
µ
h̄2 4
2
∇ δn = 0.
∇
δn
+
2mb2ρ x
4m2 x
Expanding into the Bogoliubov eigenmodes of the system, then
¨ k + ǫ2 (t)δnk = 0
h̄2 δn
(11)
k
where
ǫk (t) =
2.4 Analytic results
Alternatively, the appearance of the stripes and their
statistical properties can be described analytically using
the self-similar solutions of the GP equation valid for the
expanding cloud in the Thomas-Fermi regime (see e.g.
[24]). In the following we shall assume for simplicity that
the condensate is an infinite cylinder. This assumption is
justified since the typical size of the excitations is much
smaller than the axial size of the condensate. Therefore,
at the end of the calculation, the unperturbed density
will be substituted by the corresponding local density.
In this way, the condensate without initial fluctuations
evolve according to the self–similar solution
√
n0
(6)
ψ = 2 eiφ0 ,
bρ (t)
ḃ
m ρ 2
where b2ρ (t) = 1 + ωρ2 t2 , φ0 = 2h̄
bρ ρ , ρ is the radial
coordinate, and n0 is the Thomas–Fermi density profile.
The equations which determine the ballistic expansion in
presence of fluctuations can be obtained by linearizing
around the self–similar solution, for the density n0 + δn,
and the phase φ0 + φ. Introducing this expressions in the
corresponding Gross–Pitaevskii equation, we obtain:
2
˙ = 1 Ôφ − h̄∇x (n0 φ) ,
δn
b2ρ
m
h̄∇2x δn
gn0 δn
1
δn
Ô
−
+
n0 φ̇ =
,
h̄b2ρ
4b2ρ
n0
4m
(7)
(8)
where Ô = −(h̄/m)(∇ρ′ n0 ·∇ρ′ +n0 ∇2ρ′ ), and ρ′ = ρ/bρ .
By combining Eqs. (7) and (8), we obtain:
!
2 4
2
gn
∇
δn
h̄
Ôφ
∂
δn
h̄
∇
δn
0
x
x
¨ −
δn
+ 2 ∇2x Ô .
=
+
mb2ρ
4m2
∂t b2ρ
4bρ m
n0
(9)
The last term at the rhs of Eq. (9) can be neglected
if µ/h̄ωρ ≫ 1 (i.e. in the Thomas–Fermi limit). The
third term in the lhs of the equation (quantum pressure term) is for short times smaller than the second
one on the q
lhs, but becomes comparable with it for times
t ∼ (1/ωρ ) µm/h̄2 k 2 . This becomes an important point
as discussed below. Averaging over the radial profile (employing the expansion in powers of ρ as discussed in Ref.
(10)
s
h̄4 k 4
µh̄2 k 2
+
,
2
2mbρ
4m2
(12)
is the Bogoliubov spectrum.
part of the specFor short times ωρ t <
∼ 1 the phonon
q 2 2
µh̄ k
trum is dominant, i.e. ǫk (t) ≃
2mb2ρ . The resulting
equation can be solved in terms of hypergeometric functions. For larger times, the free particle part of the spec2 2
k
, and the equation can be
trum dominates, i.e. ǫk = h̄2m
solved in terms of Bessel functions. The two regimes may
be matched, using the asymptotic expansions of both hypergeometric and Bessel functions. In this way, we end
up with the analytic expression for the relative density
fluctuations
!
X
ǫ2j τ
δ̂n
−(ǫj /h̄ωρ )2
=2
φ̂j ,
(13)
τ
sin
n0
µh̄ωρ
j
p
where the sum extends over the axial modes ǫj = h̄ωx j(j + 3)/4,
τ = ωρ t, and φ̂j is the contribution of the j–th mode to
the phase operator in Eq. (2).
By substituting the operators by the corresponding
c–numbers as discussed above, we have checked that the
analytical expressions agree very well with the numerical
results (see also Fig. 2).
2.5 Averages
From Eq.(13) one obtains a closed relation for the mean
square density fluctuations σ 2 by averaging (δn/n0 )2
over different realizations of the initial phase. Taking
into account that h(αj + α∗j )(αj ′ + α∗j ′ )i/4 = Nj δjj ′ /2,
then the mean square fluctuations are given by:
*
2 +
T
δn(x, t)
=
C(N0 , ωρ , ωx , x, t)2 ,
(14)
n0 (x, t)
Tφ
where
C(N0 , ωρ , ωx , x, t)2 =
2 (j+3/2)2
∞
x
(j + 3/2)2
1X 2
− ω
ωρ
2
sin
e
2 j=1
4α(1 − (x/L)2 )
(j + 2)(2j + 3)
(1,1) x
(Pj
)2 ,
j(j + 1)(j + 3)
L
with α = µ/h̄ωx2 t.
ln(2ωρ t)
(15)
density [a.u.]
Phase Fluctuations in Bose-Einstein Condensates
5
2
2 √
convoluted by a resolution function ζ(x) = e−x /σ / πσ,
where σ characterizes the experimental resolution (Fig. 2).
Other experimental limitations can be easily taken
into account by following a similar procedure. In particular, a smoothing of the observed density ripples is produced if the laser that integrates the condensate column
density in the absorption detection scheme is not exactly
parallel to the density stripes. This effect can be easily
incorporated into the calculations, by filtering the density distribution with a cut–off function in the Fourier
spectrum of the form exp(ikRbρ (t)θ) sin(kRbρ (t)θ)/kT bρ (t)θ,
where θ is the angle between the laser and the density
ripples. We return to this point in the experimental section.
0
3 Experimental Results
3.1 Experimental Setup
0
-1.0
-0.5
0.0
0.5
1.0
x/L
Fig. 2 Typical density profile after 25 ms time-of-flight.
Top: numerical simulation without taking into account a
limited experimental resolution. Bottom: numerical simulation (dotted line) compared with the analytic theory (solid
line), both taking into account an experimental resolution of
σ = 3 µm. All profiles were calculated for ωx = 2π × 14 Hz,
ωρ = 2π × 508 Hz, N0 = 2 × 105 , and T = 0.5 Tc . For all the
figures the same initial phase pattern was used.
(1,1)
(x))2 ≃ 4(1 −
Using the quasiclassical average (Pj
x2 )−3/2 /πj and transforming the sum over j into an integral, for the central part of the cloud (x ≈ 0) we obtain
*
2 +
2
δn(0, t)
σBEC
=
=
n0 (0, t)
n0
v s
u
r
2
u
√
T
ln τ t
h̄ωρ τ
− 2 .
1+ 1+
ǫTφ
π
µ ln τ
(16)
Note that Eq. (16) provides a direct relation between
the observed density fluctuations and temperature, and
thus can be used for thermometry at very low T .
This closed expression would provide an accurate description of the density fluctuations in the absence of any
experimental limitation. However, in practice, the observation of the density fluctuations is limited by the spatial resolution of the experiment. This fact can be easily
taken into account in the calculations by substituting
(1,1)
(x/L) in Eq. (15) by the corresponding function
Pj
The experiment was performed with Bose-Einstein condensates of 87 Rb atoms in the |F = 2, mF = +2i hyperfine ground state. As described previously [25,26] we
load a magneto-optical trap with a few times 109 atoms
from a chirp slowed thermal beam. This is followed by
a short period of subdoppler cooling and optical pumping into the desired magnetic sublevel. The atomic cloud
is then loaded into an Ioffe-Pritchard type (cloverleaf)
magnetic trap and finally adiabatically compressed to
allow efficient rf evaporative cooling. The fundamental
frequencies of our magnetic trap are ωx = 2π × 14 Hz
and ωρ = 2π × 365 Hz along the axial and radial direction, respectively. Due to the highly anisotropic confining potential with an aspect ratio λ = ωρ /ωx of 26 the
condensates are already elongated along the horizontal
x-axis.
In order to study the dependence of phase fluctuations on the trapping geometry we realized a wide range
of radial confinement strengths. To allow for stronger
radial confinement we used a holographically generated
blue detuned Laguerre-Gauss mode (TEM∗01 ) laser beam
to form a combined magnetic and optical dipole potential trap [8]. In this combined trap BECs were produced
in a two step evaporation procedure. The atoms were
first cooled in the ’pure’ magnetic trap to a temperature
slightly above the transition temperature Tc . The optical
dipole potential was then turned on adiabatically, and finally the desired temperature was achieved by rf evaporation in the combined potential. On the other hand less
elongated condensates were produced by using a higher
offset-field for the magnetic trap leading to a weaker radial confinement.
Our measurements were performed for an axial trap
frequency of ωx = 2π × 14 Hz and radial frequencies ωρ
between 2π × 138 Hz and 2π × 715 Hz corresponding to
aspect ratios λ between 10 and 51. After rf evaporative
cooling to the desired temperature, we wait for 1 sec
6
D. Hellweg et al.
TOF
1.0
a)
l = 10
N0 [105]
0.5
1.0
0.0
b)
l = 22
2.0
10.0 ms
18.4 ms
100
200
2.0
300
0.0
200
400
l = 26
600
l = 51
N0 [105]
0.4
c)
1.0
0.2
29.3 ms
0.0
200
T [nK]
density [a.u.]
d)
400
-100
0
100
-100
x [mm]
T < 200 nK
0
100
x [mm]
T
» 400 nK
Fig. 3 (a)-(c): Absorption images for various times-of-flight
and temperatures with ωx = 2π×14 Hz and ωρ = 2π×365 Hz.
(d): Density profiles for the clouds displayed in (c) integrated
along the radial direction. In the case of T < 200 nK no thermal component was visible and the temperature was estimated to be < Tc /2.
600
0.0
200
400
600
T [nK]
Fig. 4 Distribution of phase fluctuating condensates as a
function of temperature and condensate number for four different trap geometries. White points indicate the absence of
detectable structures (σBEC < 1.5σT ), black points the existence of significant structures (σBEC > 2σT ).
mation of the density modulations as a function of temperature and trapping geometry. All measurements were
performed for times-of-flight between 18 ms and 25 ms.
For a quantitative analysis we compare the observed density distribution with the Thomas-Fermi distribution expected for a condensate without fluctuations. For each
(with rf ’shielding’) to allow the system to reach an equiimage the observed density distribution was integrated
librium state. We then switch off the trapping potential
along the radial direction [Fig. 3(d)] and then fitted by a
within 200 µs and wait for a variable time-of-flight bebimodal function with the integrated parabolic Thomasfore detecting the atomic cloud by resonant absorption
Fermi density distribution and a Gaussian for the therimaging with the imaging axis perpendicular to the long
mal cloud. In Fig 3(d) we show the integrated density
axis of the trap.
distribution close to the BEC region, thus the thermal
component is hardly visible. The standard deviations of
the experimental data from the fit were calculated in the
3.2 Ballistic expansion measurements
central region of the BEC (half width of full size), σBEC ,
and in the thermal wings, σT . The standard deviation in
Figure 3 shows typical images of the ballistically exthe thermal wings characterizes our detection noise. To
panded clouds for various times-of-flight t and temperaccount for particle number changes the standard deviatures T < Tc . The usual anisotropic expansion of the
ations were normalized to the fitted peak density n0 .
condensate due to the elongated trapping geometry is
Figure 4 shows the distribution of phase fluctuating
clearly visible. The line density profiles reflect the parabolic
condensates
as a function of temperature and condensate
shape of the BEC density distribution. As predicted by
number
for
various
trap geometries. In this figure white
the theory, we also observe pronounced stripes in the
points
indicate
the
absence of detectable structures in
density distribution. On average these stripes are more
the
density
distribution
(σBEC < 1.5σT ), whereas black
pronounced for high temperatures (right column of Fig. 3)
points
indicate
that
significant
structures larger than the
and long times-of-flight [Fig. 3(c)] indicating the buildexperimental
noise
level
(σ
BEC > 2σT ) were observed,
up of the stripes during the ballistic expansion.
i.e. the existence of phase fluctuations could be clearly
detected. The temperature and particle number of each
condensate were determined by 2D fits to the absorption
3.3 Experimental characterization of phase fluctuations
images. The temperature was determined from the width
of a Gaussian distribution fitted to the thermal wings,
In order to determine the dependence of phase fluctuathe corresponding condensate number from the integral
tions on the experimental conditions we study the for-
Phase Fluctuations in Bose-Einstein Condensates
0.1
0.1
200
400
600
0.0
0
l = 36
200
400
600
[ (sBEC /n0)2 ]1/2
sBEC/n0
0.2
0
0.20
l = 26
l = 10
0.2
0.0
7
0.16
0.12
l=10
l=22
l=26
l=36
l=51
0.08
l = 51
0.2
0.1
0.1
0.04
sBEC/n0
0.2
0.00
0.0
0
200
400
T [nK]
600
0.0
0.03
0.06
0.09
0.12
0.15
2
C T/ Tf
0
200
400
600
T [nK]
Fig. 5 Measurement of σBEC /n0 versus temperature in four
different trap geometries. The dotted lines represent the average detection noise σT /n0 . All data was taken for T < Tc .
over the Thomas-Fermi part of a bimodal fit. The statistical uncertainty in the temperature determination is
typically 15 % and less than 10 % for the particle numbers.
It is clearly visible in Fig. 4 that the number of
condensates showing detectable phase fluctuations is increasing with the aspect ratio. Furthermore, the number of realizations without detectable phase fluctuations
is growing for low temperatures and high particle numbers. Finally, there is no clear transition line between the
two regimes. There is rather a broad region in which we
observe both, condensates with and without detectable
phase fluctuations indicating the statistical character of
the phase fluctuations. Whereas for the weakest radial
confinement with λ = 10, no significant structures were
observed outside the close vicinity of the critical temperature, we detected pronounced phase fluctuations in a
broad temperature range in the case of our tightest trap
λ = 51 [Fig. 4]. According to Eq. 5 phase fluctuations can
be reduced below any detection level for sufficiently low
temperatures. However, for high aspect ratio traps this
requires very low temperatures at high particle numbers
in the condensate, making it experimentally difficult to
access. For all traps the BECs were produced by evaporating in the final potential except in the case of the
λ = 10 trap. This trap was realized by evaporating in the
tighter λ = 26 trap and then adiabatically reducing the
radial confinement. Thus the measurements show that
by changing the trapping potential adiabatically we are
able to decrease the amount of phase fluctuations. The
dynamics of the appearance and disappearance of the
phase fluctuations remains to be studied systematically
in future work.
Fig. 6 Average standard deviation of the measured line
densities [ (σBEC /n0 )2 ]1/2
exp as a function of temperature. The
temperature-axis is scaled according to the expected dependence on the other experimental parameters (Eq. 14). The
dashed line is a fit of a square root function to the data reflecting the theoretical expected dependence.
Figure 5 shows the temperature dependence of σBEC
for various trap geometries. Since the initial phase of a
Bose condensate is mapped into its density distribution
after ballistic expansion, the quantity σBEC is a direct
measure of the initial phase fluctuations. Note however,
that this method reflects the instantaneous phase of the
BEC at the time of release and therefore images taken at
the same initial conditions can look significantly different. Indeed, we observe a large spread of our experimental data (see Fig. 5). The scatter in these cases is mainly
due to the statistical character of the phase fluctuations
as well as due to the uncertainties in the temperature determination. For all traps the highest temperature data
correspond to values close to Tc (T ≈ 0.9Tc) which is increasing for tighter confinement. On average the phase
fluctuations continuously decrease with falling temperature and get more pronounced with increasing aspect
ratio. The reduction of the observed phase fluctuations
for lower temperatures is due to both, the reduced excitation spectrum at lower temperature and the increasing number of condensed atoms. Due to the loss of particles caused by evaporation, lowering the temperature
reduces the total number of particles but the fraction of
condensed atoms relative to the total particle number is
increased as shown in Fig. 4. It is not possible to determine a cut-off for the phase fluctuations, rather they
decrease until they cannot be resolved below our noise
limit. Hence all experiments with BECs at finite temperature in tightly confining elongated potentials will
be subject to axial phase fluctuations.
8
3.4 Evaluation of averaged phase fluctuations
Due to the statistical character of phase fluctuations every experimental realization has a different phase distribution. Therefore, we average the observed standard deviations for a small range of experimental parameters to
obtain general information about the phase fluctuations
and to compare our measurements to the theoretical prediction. According to Eq. (14) the standard deviation
of the fluctuations is expected to depend on the square
root of the temperature. This behaviour can not be observed in Fig. 5 since the condensate number changes
with temperature. Therefore, we use that equation to
scale out the influence of condensate number, trapping
potential, and time-of-flight. In figure 6 the measured
standard deviations are plotted as function of T /Tφ C 2 ,
where Tφ and C (C taken at x=0) are calculated from
the measured number of condensed atoms, trapping frequencies, and expansion times. The figure shows that
all data points fall on an universal square root shaped
curve. This means, that our experimental results follow
the expected dependence very well.
As a main result, with the direct link of the phase
fluctuations in the magnetic trap to the observed density modulations given by Eq. (16), our measurements
confirm the predicted behaviour of phase fluctuations in
elongated BECs.
As explained above, a precise quantitative comparison needs to take into account a reduced contrast due to
the limited experimental resolution and a possible tilt of
the detection laser beam with respect to the structures.
If we account for a resolution of our imaging system of
σ = 3µm, the predicted standard deviations are about a
factor of 2 bigger than the observed deviations. A possible reason might be a small tilt of the detection laser
beam. For our parameters a tilt of only 4◦ reduces the
observed modulations by a factor of approximately 2.
Note, that our measurements were performed in elongated geometries but our condensates still were in a
3D-regime in the sense that the chemical potential was
greater than the transverse level spacing. Nevertheless,
most of our measurements, which exhibit fluctuations
well above our noise level, correspond to the regime of
quasicondensation in which the phase coherence length
is smaller than the condensate size. For instance, for
λ = 51, T = 0.5 Tc, and N0 = 3 × 104 , one obtains
µ = 3.4h̄ωρ and lφ ≈ L/3.
4 Discussion of alternative methods
In principle, phase fluctuations may be observed by all
experiments on BEC which rely on its phase properties.
Hence, various methods may be used to study them. We
have presented a method which is based on the transfer of phase fluctuations into density modulations after
ballistic expansion. A great advantage of our method
D. Hellweg et al.
is its applicability to very different trap geometries and
that the sensitivity depends on the chosen time-of-flight
t (see Eq. (16)). In principle for longer t the method becomes more and more sensitive for initial phase fluctuations. On the other hand the decreasing density reduces
the signal-to-noise ratio leading to an optimum time-offlight for given parameters. In our case a time-of-flight
of 25 ms allows us to measure phase fluctuations of only
2
δL
≈ π/7. Moreover, time-of-flight methods constitute a
standard experimental tool to study properties of BECs.
Therefore, understanding the formation of stripes in the
density distribution is of great importance. It is instructive to discuss two of the most successful methods used
to study the coherence properties of BEC, namely Bragg
spectroscopy [10] and interferometry [3], with respect to
the measurement of phase fluctuations.
4.1 Bragg spectroscopy
The velocity field of a condensate is proportional to
the gradient of its phase. Therefore, phase fluctuations
lead to a broadening of the momentum distribution of a
trapped condensate which can be measured using Bragg
spectroscopy. Since in elongated condensates phase fluctuations are predominantly provided by axial excitations
the momentum distribution along the axial direction has
to be measured. The average momentum distribution P
is given by the Fourier transform (F T ) of the integrated
R
single-particle correlation function G(ra ) ≡ d3 rhψ̂ † (r)ψ̂(r + ra )i
[27]. Using Eq. (1) and the definition of the phase coherence length we get
Z
|r |
p
− a
d3 r n0 (r) n0 (r + ra )
P(p) = F T exp lφ
hp
i
|r |
− la
= F T exp φ ⋆ F T h n0 (r) n0 (r + ra )i(17)
r
which is a convolution of the Fourier transform of the
density distribution and that of a Gaussian with the
width of the phase coherence length lφ . Thus, the first
term is inversely proportional to lφ , the second inversely
proportional to the condensate size. In the absence of
phase fluctuations the momentum distribution is limited by the size of the condensate, i. e. , Heisenberguncertainty limited. In the regime of quasicondensates
where the phase coherence length is smaller than the
condensate size, the momentum distribution is dominated by the phase coherence length, thus increasing
the momentum width significantly with respect to the
Heisenberg limit. However, for a trapped condensate the
width of the Bragg resonance is not only determined by
this momentum width (Doppler width) but also by the
inhomogeneous density distribution, leading to a spatially inhomogeneous mean field-shift of the resonance
and therefore broadening it.
We will now briefly estimate the relative contribution
of the momentum width and the mean-field width to the
Phase Fluctuations in Bose-Einstein Condensates
4.2 Interferometry
Interferometry has been successfully used to study and
to characterize the spatial phase of Bose condensates.
Therefore, it is natural to consider possibilities to extend such methods to determine the amount of phase
fluctuations in trapped atomic gases. In analogy to classical optics, there are several methods to use interferometry to determine the coherence properties of a condensate wave function. These include, e.g., the double slit
experiment with outcoupled atom laser beams [4] and
interference between two overlapping condensates [2]. In
the latter case, two spatially separated condensates were
prepared independently and their interference was measured after a ballistic expansion. The results indicated a
flat phase distribution of the initial wavefunction for the
experimental parameters used in those measurements.
In general, interferometric measurements rely on the superposition of two wavefunctions and the interpretation
of the resulting fringe pattern. The existence of a randomly fluctuating phase as shown in Fig.1 complicates
the interpretation.
To visualize possible experimental results, let us first
consider the following situation. A given phase fluctuating condensate is coherently split in two equal parts
giving one of them a small additional velocity. Absorption images taken of the overlapping condensates will
show interference fringes. For condensates with a constant phase the images would show regularly spaced lines.
However due to the phase fluctuations these lines will be
shifted according to the relative fluctuating phase, leading to a random, spatially varying fringe spacing. Similar to the ballistic expansion measurements these images
n [a.u.]
a)
1.0
0.0
n [a.u.]
b)
1.0
0.0
3.0
n [a.u.]
total linewidth of the Bragg resonance. For typical parameters of N = 105 , wρ = 2π × 365 Hz, wx = 2π × 14 Hz
the momentum widthp
resulting from the size of the con2h̄k
= 57 Hz, where k =
densate is ∆νsize = 21/8 2πmL
2π/λ is the absolute value of the wavevector of the Bragg
beams. For p
these parameters the mean field broadening
is ∆νmf = 8/147 µh = 536 Hz. Hence the total width
q
2
2
∆νmomentum
+ ∆νmf
=
of the resonance is ∆ν =
539 Hz. Here, we have taken the expressions for the widths
from Ref. [10] and assumed Bragg spectroscopy with
counterpropagating beams since this leads to the highest momentum resolution. For a temperature T = 0.5Tc
the phase coherence length is, according to Eq. (5), lφ ≈
1.4L (δL2 ≈ π/5). Though this increases the momentum
width roughly by a factor of 2, it leads only to a total linewidth of approximately 547 Hz. Thus, the contribution of the momentum width to the total linewidth
in this case is less than 2%, making it very difficult
to detect. In contrast, our measurements show that the
method of ballistic expansion is capable of detecting even
such small phase fluctuations.
9
c)
2.0
1.0
0.0
-100
-50
0
50
100
x [mm]
Fig. 7 (a, b) Examples of independently calculated density
distributions of phase fluctuating condensates (solid line) for
ωρ = 2π×715 Hz, ωx = 2π×14 Hz, N = 5×104 and T = 0.6Tc
compared to that of a condensate at T = 0 (dashed line)
for a time-of-flight of 15 ms. (c) Interference pattern of the
superposition of the condensates shown in a) and b) with a
relative displacement of ∆x = 40µm.
will be different from shot to shot and allow only for a
statistical evaluation.
So far most interferometric measurements with BECs
have been performed after a ballistic expansion. An additional complication arises in these experiments, since
both density and phase of the sample evolve during the
free expansion time. Let us assume that two independent quasicondensates are brought to partial overlap after a time-of-flight with zero relative velocity (using e. g.
Bragg pulses [5,7,8]). At this point modulations have
appeared in the density distribution and the phase has
evolved from its initial random pattern. Both parts now
contribute to the formation of stripes in the interference
pattern, making a direct interpretation difficult.
Figure 7 shows an example of an expected absorption image for our experimental conditions. Both wavefunctions Ψ1 and Ψ2 of the phase fluctuating condensates were calculated after a time-of-flight of 15 ms using the GP equation. The interference pattern is given
by |Ψ1 (x) + Ψ2 (x + ∆x)|2 , where ∆x is the displacement
of the two quasicondensates. Fig. 7 a) and b) show the
density distribution of the two quasicondensates (solid
line) displaced by ∆x = 40µm ≈ 2lφ in comparison with
the Thomas-Fermi distribution of a condensate at T = 0
(dashed line). Note, that the structures shown there are
10
the density modulations due to the phase fluctuations
only. The resulting interference pattern of these two independent condensates is then displayed in Fig. 7 c). At
T = 0 the interference pattern is due to the small axial
expansion velocity, which leads to a quadratic phase. In
the case of the quasicondensates the profile looks considerably different, i. e. the phase fluctuations lead to
a significant change of the experimental outcome. Due
to the difficult interpretation of the density distribution,
interferometric methods together with a long expansion
time are not suited very well for analysis of phase fluctuations. However, interferometric methods that work with
a short expansion time or even with trapped condensates might be very useful for visualization of the spatial profile of phase fluctuations. Such methods might
include, e.g. preparation of two independent quasicondensates with a small vertical separation. After a short
time-of-flight the radial expansion of the clouds leads
to their partial superposition, and consequently interference in the overlap region. In the case of constant relative
phase in the axial direction horizontal stripes appear in
the overlap region. Due to the phase fluctuations the
relative phase can, however, change as a function of axial position, which is seen as bending of the interference
fringes in the vertical direction. Therefore, the relative
phase profile of the clouds can be imaged by measuring
the shape of the interference fringes.
5 Conclusion
We have presented detailed experimental and theoretical studies of phase fluctuations in the equilibrium state
of BECs. A strong dependence on the trap geometry
and temperature has been found in agreement with the
theoretical prediction. We have shown that phase fluctuations are a general property of elongated condensates. By measuring the phase fluctuations and comparing the temperature with Tφ we have demonstrated instances, where the phase coherence length was smaller
than the axial size of the condensate, i.e. the initial cloud
was a quasicondensate. Our results set severe limitations on applications of BECs in interferometric measurements, and for guided atom laser beams. Our experimental method combined with the theoretical analysis
provides a method of BEC thermometry. Further studies
of the effect of phase fluctuations, e. g. on the superfluid
properties, will give additional insight to the behaviour
of degenerate quantum gases at finite temperature.
This work is supported by the Deutsche Forschungsgemeinschaft within the SFB 407 and the Schwerpunktprogramm ”Wechselwirkungen ultrakalter atomarer und
molekularer Gase”, and the European Science Foundation (ESF) within the BEC2000+ programme. DSP acknowledges support from the Alexander von Humboldt
Foundation, from the Dutch Foundations NWO and FOM,
and from the Russian Foundation for Basic Research.
D. Hellweg et al.
References
1. See e.g. F. Dalfovo, S. Giorgini, L. P. Pitaevskii, and S.
Stringari, Rev. Mod. Phys. 71, 463 (1999).
2. M. R. Andrews, C. G. Townsend, H.-J. Miesner, D. S.
Durfee, D. M. Kurn, and W. Ketterle, Science 275, 637
(1997).
3. E. W. Hagley, L. Deng, M. Kozuma, M. Trippenbach,
Y. B. Band, M. Edwards, M. R. Doery, P. S. Julienne, K.
Helmerson, S. L. Rolston, and W. D. Phillips, Phys. Rev.
Lett. 83, 3112 (1999).
4. I. Bloch, T. W. Hänsch, and T. Esslinger, Nature 403,
166 (2000).
5. J. E. Simsarian, J. Denschlag, Mark Edwards, Charles W.
Clark, L. Deng, E. W. Hagley, K. Helmerson, S. L. Rolston,
and W. D. Phillips, Phys. Rev. Lett. 85, 2040 (2000).
6. D. S. Hall, M. R. Matthews, C. E. Wieman, and E.A.
Cornell, Phys. Rev. Lett. 81, 1543 (1998).
7. Y. Torii, Y. Suzuki, M. Kozuma, T. Sugiura, T. Kuga, L.
Deng, and E. W. Hagley, Phys. Rev. A 61, R041602 (2000).
8. K. Bongs, S. Burger, S. Dettmer, D. Hellweg, J. Arlt,
W. Ertmer, and K. Sengstock, Phys. Rev. A 63, R31602
(2000).
9. M.-O. Mewes, M. R. Andrews, D. M. Kurn, D. S. Durfee,
C. G. Townsend, and W. Ketterle, Phys. Rev. Lett. 78,
582 (1997); B. P. Anderson, and M. A. Kasevich, Science
282, 1686 (1998); E. W. Hagley, L. Deng, M. Kozuma,
J. Wen, K. Helmerson, S.L. Rolston, and W. D. Phillips,
Science 283, 1706 (1999); I. Bloch, T. W. Hänsch, and T.
Esslinger, Phys. Rev. Lett. 82, 3008 (1999); F. Gerbier, P.
Bouyer, and A. Aspect, Phys. Rev. Lett. 86, 4729 (2001).
10. J. Stenger, S. Inouye, A. P. Chikkatur, D. M. StamperKurn, D. E. Pritchard, and W. Ketterle, Phys. Rev. Lett.
82, 4569 (1999).
11. Yu. Kagan, in Bose-Einstein Condensation, edited by A.
Griffin, D. W. Snoke, and S. Stringari (Cambridge University Press, Cambridge, 1995), p. 202–225; and references
therein.
12. D. S. Petrov, G. V. Shlyapnikov, and J. T. M. Walraven,
Phys. Rev. Lett. 85, 3745 (2000).
13. D. S. Petrov, M. Holzmann, and G. V. Shlyapnikov,
Phys. Rev. Lett. 84, 2551 (2000).
14. Yu. Kagan, V. A. Kashurnikov, A. V. Krasavin, N. V.
Prokof’ev, and B.V. Svistunov, Phys. Rev. A 61, 43608
(2000); N. J. van Druten, and W. Ketterle, Phys. Rev. Lett.
79, 549 (1997); H. Monien, M. Linn, and N. Elstner, Phys.
Rev. A 58, R3395 (1998); M. Olshanii, Phys. Rev. Lett.
81, 938 (1998); and references therein.
15. A. I. Safonov, S. A. Vasilyev, I. S. Yasnikov, I. I. Lukashevich, and S. Jaakkola, Phys. Rev. Lett. 81, 4545 (1998).
16. A. Görlitz, J. M. Vogels, A. E. Leanhardt, C. Raman,
T. L. Gustavson, J. R. Abo-Shaeer, A. P. Chikkatur, S.
Gupta, S. Inouye, T. P. Rosenband, D. E. Pritchard, and
W. Ketterle, cond-mat/0104549.
17. F. Schreck, L. Khaykovich, K. L. Corwin, G. Ferrari, T. Bourdel, J. Cubizolles, and C. Salomon, condmat/0107442.
18. D. S. Petrov, G. V. Shlyapnikov, and J. T. M. Walraven,
Phys. Rev. Lett. 87, 050404 (2001).
19. S. Dettmer, D. Hellweg, P. Ryytty, J. J. Arlt, W. Ertmer,
K. Sengstock, D. S. Petrov, G. V. Shlyapnikov, H. Kreutzmann, L. Santos, and M. Lewenstein, cond-mat/0105525.
Phase Fluctuations in Bose-Einstein Condensates
20. J. Reichel, and T. W. Hänsch, private communication;
C. Zimmermann, private communication.
21. S. Stringari, Phys. Rev. A 58, 2385 (1998).
22. V. N. Popov, Functional Integrals in Quantum Field Theory and Statistical Physics, (D. Reidel Pub., Dordrecht,
1983).
23. S. I. Shevchenko, Sov. J. Low Temp. Phys. 18, 223
(1992).
24. Yu. Kagan, E. L. Surkov, and G. V. Shlyapnikov, Phys.
Rev. A54, R1753 (1996); Y. Castin and R. Dum, Phys.
Rev. Lett. 77, 5315 (1996).
25. K. Bongs, S. Burger, G. Birkl, K. Sengstock, W. Ertmer,
K. Rza̧żewski, A. Sanpera, and M. Lewenstein, Phys. Rev.
Lett. 83, 3577 (1999).
26. S. Burger, K. Bongs, S. Dettmer, W. Ertmer, K. Sengstock, A. Sanpera, G. V. Shlyapnikov, and M. Lewenstein,
Phys. Rev. Lett. 83, 5198 (1999).
27. See for example C. Cohen-Tannoudji, Lecture notes
of the Euroschool Bose-Einstein Condensates and Atom
lasers, Cargese, France (2000).
11