• Aucun résultat trouvé

Particle number fluctuations in a cloven trapped Bose gas at finite temperature

N/A
N/A
Protected

Academic year: 2021

Partager "Particle number fluctuations in a cloven trapped Bose gas at finite temperature"

Copied!
9
0
0

Texte intégral

(1)

HAL Id: hal-00460982

https://hal.archives-ouvertes.fr/hal-00460982

Submitted on 3 Mar 2010

HAL

is a multi-disciplinary open access archive for the deposit and dissemination of sci- entific research documents, whether they are pub- lished or not. The documents may come from teaching and research institutions in France or abroad, or from public or private research centers.

L’archive ouverte pluridisciplinaire

HAL, est

destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des établissements d’enseignement et de recherche français ou étrangers, des laboratoires publics ou privés.

Particle number fluctuations in a cloven trapped Bose gas at finite temperature

Alice Sinatra, Yvan Castin, Yun Li

To cite this version:

Alice Sinatra, Yvan Castin, Yun Li. Particle number fluctuations in a cloven trapped Bose gas at finite

temperature. Physical Review A, American Physical Society, 2010, 81, pp.053623. �hal-00460982�

(2)

A. Sinatra and Y. Castin

Laboratoire Kastler Brossel, Ecole Normale Sup´erieure, UPMC and CNRS, 24 rue Lhomond, 75231 Paris Cedex 05, France

Yun Li

State Key Laboratory of Precision Spectroscopy, Department of Physics, East China Normal University, Shanghai 200062, China

We study fluctuations in the atom number difference between two halves of a harmonically trapped Bose gas in three dimensions. We solve the problem analytically for non interacting atoms. In the interacting case we find an analytical solution in the Thomas-Fermi and high temperature limit in good agreement with classical field simulations. In the large system size limit, fluctuations in the number difference are maximal for a temperature T ≃0.7Tcwhere Tc is the critical temperature, independently of the trap anisotropy. The occurrence of this maximum is due to an interference effect between the condensate and the non-condensed field.

PACS numbers: 03.75.Hh 03.75.Kk, 67.10.Ba

I. INTRODUCTION

Fluctuations appearing when counting the atoms in a given sub-volume of a quantum system, are a funda- mental feature determined by the interplay between the atomic interactions and quantum statistics. They can be used to investigate many-body properties of the system and in particular non-local properties of the g(2) pair- correlation function. These fluctuations were studied at zero temperature for quantum gases in different regimes and spatial dimensions in [1–5]. Sub-poissonian fluctu- ations appear for non-interacting fermions and for in- teracting bosons. A related issue in condensed matter physics is that of partition noise in electron systems [6].

In cold atoms experiments it is now possible to di- rectly measure the fluctuations in atom number within a given region, as done for example in [7] for a quasi one- dimensional system. However finite temperature plays a major role in experiments. Very recently, experiments were done on an atom chip where a cold gas of Rb atoms, initially trapped in an single harmonic potential well, is split into two parts by raising a potential barrier. Accu- rate statistics of the particle number difference between the left and right wells NL−NR is then performed in the modified potential. By varying the initial tempera- ture of the sample across the transition for Bose-Einstein condensation, they observe a marked peak in the fluctu- ations of the particle number difference below the tran- sition temperature Tc while shot noise fluctuations are recovered forT > Tc. For T ≪Tc they finally get sub- shot noise fluctuations due to the repulsive interactions between atoms [8]. Here we show that the peak of fluc- tuations for T < Tc is in fact a general feature already appearing in a single harmonic well if we look at the fluctuations of the atom number differenceNL−NRbe- tween the left half and the right half of the trap along one axis. We show that, contrarily to what happens to the total number fluctuations, fluctuations in the parti- cle number differenceNL−NR can be computed within

the grand canonical ensemble without any pathology. In the first part of the paper we address the ideal gas case for which we find the complete analytical solution in the grand canonical ensemble. We derive the asymptotic be- haviors for T ≪ Tc and T ≫ Tc and we explain the physical origin of the “bump” in fluctuations of the par- ticle number difference forT < Tc. In the second part of our paper we then address the interacting case.

II. IDEAL GAS: EXACT SOLUTION We consider an ideal gas of bosons in a three- dimensional harmonic potential. The signal we are in- terested in is the particle number difference NL −NR

between the left and right halves of the harmonic poten- tial along one direction, as shown in Fig.1. In terms of

L R

x

FIG. 1: We consider fluctuations of the particle number dif- ferenceNL−NR between the left and the right halves of a three-dimensional harmonic potential.

the atomic field operators:

NL−NR= Z

r∈L

ψψ− Z

r∈R

ψψ . (1)

(3)

2 Due to the symmetry of the problem, NL−NR has a

zero mean value. It is convenient to express its variance in terms of the unnormalized pair correlation function

g(2)(r,r) =hψ(r)ψ(r)ψ(r)ψ(r)i. (2) Normally ordering the field operators with the help of the bosonic commutation relation pulls out a term equal to the mean total number of particles:

Var(NL−NR) = hNi+ 2 Z

r∈L

Z

r∈L

g(2)(r,r)

− Z

r∈L

Z

r∈R

g(2)(r,r)

. (3)

We assume that the system is in thermal equilibrium in the grand canonical ensemble with β = 1/kBT the in- verse temperature and µ the chemical potential. Since the density operator is Gaussian, we can use Wick’s the- orem and express g(2)(r,r) in terms of the first-order coherence functiong(1)(r,r) =hψ(r)ψ(r)i:

g(2)(r,r) =g(1)(r,r)g(1)(r,r) +g(1)(r,r)g(1)(r,r). (4) Theg(1)(r,r) is a matrix element of the one-body density operator

g(1)(r,r) =hr| 1

z−1eβh1−1|ri (5) whereh1 is the single particle Hamiltonian

h1= p2

2m+ X

α=x,y,z

1

2mωα2rα2. (6) To computeg1, it is convenient to expand the one-body density operator in powers of the fugacityz=eβµ[9]:

g(1)(r,r) =hr|

X

l=1

zle−lβh1|ri. (7) On the other hand for a harmonic potential the matrix elements ofe−βh1 are known [10]. We then have:

g(1)(r,r) =

X

l=1

zlm¯ω 2π~

3/2 Y

α=x,y,z

[sinh(lηα)]−1/2 × exp

−mωα

4~

(rα+rα)2tanh lηα

2

+ (rα−rα)2coth

α

2

(8) where we introduced the geometric mean of the oscil- lation frequencies ¯ω = (ωxωyωz)1/3 and ηα = β~ωα. It is convenient to renormalize the fugacity introducing

˜

z=zexp(−P

αηα/2) that spans the interval (0,1). Af- ter some algebra [11], the variance of NL−NR is ex- pressed as a double sum that we reorder as

Var(NL−NR) =hNi+

X

s=1

css (9)

with cs=

s−1

X

l=1

1−π4arctanq

tanh(12x) tanh1

2(s−l)ηx Y

α=x,y,z

1−e−ηαs , (10) withc1= 0. Correspondingly, the mean atom number is expressed as

hNi=

X

l=1

˜ zlY

α

1 + coth(lηα/2)

2 . (11)

This constitutes our analytical solution of the problem in the grand canonical ensemble.

In practice, the forms (9) and (11) are difficult to han- dle in the degenerate regime, since the series converge very slowly when ˜z → 1. A useful exact rewriting is obtained by pulling out the asymptotic behaviors of the summands. For the signal we obtain the operational form

Var(NL−NR) =hNi+chN0i+

X

s=1

(cs−c)˜zs (12) where

c= lim

s→∞cs= 2

X

l=1

1− 4 πarctan

r

tanhlηx

2

! , (13) andhN0i= ˜z/(1−z) is the mean number of condensate˜ particles. The mean atom number is rewritten as

hNi=hN0i+

X

l=1

˜ zl

"

−1 +Y

α

1 + coth(lηα/2) 2

# . (14) In Fig. 2 we show an example of fluctuations of the parti- cle number difference for realistic parameters of an atom- chip experiment. It is apparent that fluctuations are weakly super-poissonian aboveTc and a marked peak of fluctuations occurs forT < Tc. In the following sections we perform some approximations or transformations in order to obtain explicit formulas and get some physical insight.

III. APPROXIMATE FORMULAS FOR kBT≫~ω In this section, we consider the limit of a large atom number and a high temperaturekBT ≫~ωα for allα.

Non-condensed regime:Taking the limitηα≪1 in Eq.

(11) we get

hNi ≃ kBT

~ω¯ 3

g3(˜z) (15) where gα(z) = P

l=1zl/lα is the Bose function. From this equation we recover the usual definition of the critical temperatureTc:

kBTc= [N/ζ(3)]1/3~ω ,¯ (16)

(4)

whereζ(3) =g3(1) withζthe Riemann function. Taking the same limit in (9) gives

Var(NL−NR)≃ hNi

1 +g2(˜z)−g3(˜z) ζ(3)

T3 Tc3

. (17) AtT =Tc this leads to

Var(NL−NR)(Tc)≃Nζ(2)

ζ(3) ≃1.37N , (18) showing that the non-condensed gas is weakly super- poissonian, as already observed in Fig.2.

Alternatively, one may directly take the limit ηα→0 in Eq.(8), yielding

g(1)(r,r)≃

X

l=1

˜ zl l3/2λ3dB

Y

α=x,y,z

exp (

−lmω2α 2kBT

rα+rα 2

2

+ π

2dB(rα−rα)2 )

. (19) This semiclassical approximation coincides with the widely used local density approximation, and allows to recover (15) and (17).

Bose-condensed regime: In this regime ˜z→1 so we use the splittings (12) and (14). Setting ˜z= 1 and taking the limitηα→0 in each term of the sum overl in (14), we obtain the usual condensate fraction

hN0i

hNi ≃1−T3

Tc3. (20)

The same procedure may be applied to the sum oversin (12). The calculation of the small-ηlimit ofcrequires a

0 0.2 0.4 0.6 0.8 1 1.2 1.4

T/Tc 0

4 8 12 16 20 24

Var(NL-NR)/<N>

1 1.2 1.4

1 1.1 1.2 1.3 1.4

omx=1,omy=omz=2

FIG. 2: (Color online) Normalized variance of the particle number differenceNL−NR as a function of temperature in a cigar-shaped trap with ωy = ωz = 2ωx. The number of particles ishNi = 6000 (black lines) and hNi = 13000 (red lines). The inset is a magnification of theT > Tcregion. Solid lines: exact result (9) together with (11). Dashed line forT >

Tc: approximate result (17) together with (15). Dashed line for T < Tc: approximate result resulting from the improved estimates (26) and (25). The temperatureT is expressed in units of the critical temperatureTcdefined in Eq. (16).

different technique: Contrarily to the previous cases, the sum in (13) is not dominated by values of the summation index≪1/ηαand explores high values ofl∼1/ηx. As a remarkable consequence, the local-density approximation fails in this case [12]. We find that one rather has to replace the sum overlby an integral in (13):

c≃ Z +∞

0

dl f(l) = 2 ln 2 ηx

(21) withf(x) = 2−(8/π) arctanp

tanh(xηx/2). This leads to the simple formula forT < Tc:

Var(NL−NR)≃ hNi

1 +ζ(2)−ζ(3) ζ(3)

T3 Tc3 +2 ln 2

kBTc

x

1−T3 Tc3

T Tc

. (22) The second line of Eq.(22) is a new contribution involving the macroscopic value ofhN0ibelow the critical temper- ature. SincekBTc≫~ωxhere, it is the dominant contri- bution to the fluctuations of the particle number differ- ence. It clearly leads to the occurrence of a maximum of these fluctuations, at a temperature which remarkably is independent of the trap anisotropy:

T Tc

max

≃2−2/3≃0.63. (23) The corresponding variance is strongly super-poissonian in the large atom-number limit:

[Var(NL−NR)]max≃ hNi

"

1 +3 4ln 2

2hNi ζ(3)

1/3

¯ ω ωx

# . (24) For the parameters of the upper curve in Fig. 2, these approximate formulas lead to a maximal variance over hNiequal to≃27.5, whereas the exact result is≃22.2, located atT /Tc≃0.61. We thus see that finite size cor- rections remain important even for the large atom num- ber hNi = 13000. Fortunately, it is straightforward to calculate the next order correction. For the condensate atom number we simply expand the summand in (14) up to orderηα2, and we recover the known result [13]:

hN0i

hNi ≃1−T3 Tc3−T2

Tc2 3ζ(2) 2ζ(3)

m

kBTc

(25) with the arithmetic mean ωm = P

αωα/3. For c we use the Euler-Mac Laurin summation formula, applied to the previously defined functionf(x) over the interval (1,+∞), and we obtain

c= 2 ln 2

ηx −1 +Aηx1/2+O(η3/2x ) (26) with A = −25/2ζ(−1/2)/π ≃ 0.374 [14]. Using these more accurate formulas for hN0i and c in the second

(5)

4 term of the right-hand side of (12) leads to an excellent

agreement with the exact result, see the dashed lines in Fig. 2 practically indistinguishable from the solid lines.

Note that the effect of the −1 correction in (26) is to change the shot noise term 1 in the square brackets (22) and (24) into 1− hN0i/hNi.

IV. A PHYSICAL ANALYSIS SINGLING OUT THE CONDENSATE MODE

To investigate the contribution of different physical ef- fects on our observable, it is convenient to go back to the expression (3) and split the field operator into the condensate and the non-condensed part:

ψ(r) =φ(r)a0+δψ(r), (27) where φ(r) is the ground mode wavefunction of the harmonic potential. The pair correlation function g(2) is then expressed as the sum of three contributions, g(2)(r,r) =gI(2)+gII(2)+gIII(2), sorted by increasing powers ofδψ:

g(2)I = φ2(r)φ2(r)ha0a0a0a0i (28) g(2)II = h

φ(r)φ(r)ha0a0δψ(r)δψ(r)i+r↔ri + h

φ2(r)ha0a0δψ(r)δψ(r)i+r↔ri (29) gIII(2) = hδψ(r)δψ(r)δψ(r)δψ(r)i. (30) Averages involving different numbers of operatorsa0and a0 vanish since the system is in a statistical mixture of Fock states in the harmonic oscillator eigenbasis.

The term gI(2) originates from the condensate mode only. Its contribution to Var(NL−NR) is zero for symme- try reasons. This is a crucial advantage, because it makes our observable immune to the non-physical fluctuations of the number of condensate particles in the grand canon- ical ensemble, and legitimates the use of that ensemble.

For the same symmetry reasons, the second line of gII(2) has a zero contribution to Var(NL−NR). The termg(2)III originates from the non-condensed gas only. Below Tc

this gas is saturated (˜z≃1) to a number of particles scal- ing asT3, see (20), and its contribution to Var(NL−NR), maximal at T =Tc, makes the fluctuations in the par- ticle number difference only weakly super-poissonian, as already discussed.

Below Tc, the first line of gII(2) is thus the important term. It originates from a beating between the conden- sate and the non-condensed fields. Its contribution to Var(NL −NR) can be evaluated from Wick’s theorem [15]:

VarII(NL−NR) = 4hN0i Z

r∈L

Z

r∈L

φ(r)φ(r)g(1)(r,r)

− Z

r∈L

Z

r∈R

φ(r)φ(r)g(1)(r,r)

. (31)

Using Eq. (8) and setting ˜z≃1 ing(1), after some alge- bra, we obtain forT < Tc

VarII(NL−NR)≃ hN0ic. (32) We can thus give a physical meaning to the mathematical splitting (12) forT < Tc: The second term and the sum overs in the right-hand side of (12) respectively corre- spond to the condensate-non-condensed beating contri- bution VarII(NL−NR) and to the purely non-condensed contribution VarIII(NL−NR).

V. CLASSICAL FIELD APPROXIMATION AND INTERACTING CASE

In this section we show for the ideal gas that the clas- sical field approximation [16, 17, 19, 20] exactly gives the high temperature limit (kBT ≫~ωx) of the ampli- tudec in the condensate-non-condensed beating term of Var(NL−NR) (32). We then use the classical field approximation to extend our analysis to the interacting case.

A. Ideal gas: test of the classical field approximation

It is useful to rewrite (31) as an integral over the whole space introducing the sign functions(x). One then rec- ognizes two closure relations onrandr and obtains:

c= 2hφ|s(x) 1

z−1eβh1−1s(x)|φi. (33) Correspondingly, in the classical field limit:

cclass = 2hφ|s(x) kBT h1−P

α

~ωα

2

s(x)|φi. (34) Inserting a closure relation on the eigenstates of the har- monic oscillator|niin (33), we are then led to calculate the matrix elements in one dimension

xh0|s(x)|nix= 2~

xn 1/2

φx0(0)φxn−1(0). (35) To obtain this result, we introduced the raising opera- tor ax of the harmonic oscillator along x and we eval- uated the matrix element xh0|[s(x), ax]|n−1ix in two different ways. First, it is equal ton1/2xh0|s(x)|nixsince ax|n−1ix=n1/2|nix. Second it can be deduced from the commutator [Y(x), ax] = [~/(2mωx)]1/2δ(x) where Y(x) is the Heaviside distribution. From the known values of φxn(0) (see e.g. [3]), we thus obtain:

c= 4 π

X

m∈N

he(2m+1)ηx−1i−1 (2m)!

22m(2m+ 1)(m!)2. (36)

(6)

The equivalent for the classical field is readily computed and we obtain:

cclass =2 ln 2 ηx

(37) showing that the classical field approximation gives the right answer for the dominant contribution to our ob- servable. Moreover, going to the first order beyond the classical field approximation, that is including the −1/2 term in the expansion 1/[exp(u)−1] =u−1−1/2 +O(u), equation (33) readily gives the term−1 in (26). In what follows we will use the classical field approximation to treat the interacting case.

B. Classical field simulations for the interacting gas In Fig.3 we show results of a classical field simulation in presence of interactions for two different atom numbers (blue circles and black triangles). The non interacting case for one atom number (red circles and red curve) is shown for comparison. We note that the assumption of an ideal Bose gas is nowadays realistic: Recently, the use of a Feshbach resonance has allowed to reach a scattering length of a = 0.06 Bohr radii [18]. We estimate from the Gross-Pitaevskii equation for a pure condensate that interactions are negligible if 12N gR

|φ|4 ≪ ~ωmin where ωmin is the smallest of the three oscillation frequencies ωα. For the parameters of Fig. 3 (red curve) this results in the well-satisfied conditionN≪105.

In presence of repulsive interactions, the peak of fluc- tuations in the particle number difference at T < Tc is still present, approximately in the same position, but its amplitude is strongly suppressed with respect to the ideal gas case. Another notable effect is that the dependence of the curve on the atom number is almost suppressed in the interacting case.

To compute the normally ordered contribution in (3) we generate 800 stochastic fields in the canon- ical ensemble sampling the Glauber-P function that we approximate by the classical distribution P ∝ δ N−R

|ψ|2

exp{−βE[ψ, ψ]} where E[ψ, ψ] is the Gross-Pitaevskii energy functional

E[ψ, ψ] = Z

ψh1ψ+g

2|ψ|4, (38) where the coupling constant g = 4π~2a/m is propor- tional to the s-wave scattering length a. The approxi- mate Glauber-P function is sampled by a brownian mo- tion simulation in imaginary time [21]. The simula- tion results are plotted as a function of T /Tcclass where the transition temperature in the classical field simu- lations, extracted by diagonalization of the one-body density matrix, slightly differs from Tc given by (16):

Tcclass= 1.15Tc.

0 0,5 1 1,5

T/Tc 1

10

Var (NL- NR) / N

FIG. 3: (Color online) Symbols: Classical field simulations forN= 17000 (circles) andN= 6000 (triangles)87Rb atoms with interactions (blue and black), and without interactions for comparison (red). Red solid line: analytical prediction (9) in the non interacting case. Blue and black solid lines: analyt- ical prediction (54) in the interacting case. The oscillation fre- quencies alongx, y, zare ωα/2π[Hz] = (234,1120,1473). For the interacting case thes-wave scattering length isa= 100.4 Bohr radii.

C. Analytical treatment for the interacting gas In the interacting case, we perform the same splitting as in (27) except that now the condensate field ψ0 = hN0i1/2φsolves the Gross-Pitaevskii equation

(h1+gψ020=µψ0. (39) For the dominant contribution of g(2) to the signal, for T < Tc, we then obtain

g(2)II(r,r) = 2hN0iφ(r)φ(r)

(r)Λ(r)i+hΛ(r)Λ(r)i + hN0i

φ2(r)hΛ(r)Λ(r)i+r↔r

, (40) where we have neglected fluctuations of N0, set a0 = hN0i1/2e and introduced Λ(r) = e−iθδψ(r) [22]. The second line in (40) brings no contribution to the signal for symmetry reasons. Note that the terms ing(2) that are cubic in the condensate field vanish sincehΛ(r)i= 0.

In the Bogoliubov [23] and classical field approxima- tion, the non-condensed field Λ has an equilibrium prob- ability distribution

P(Λ,Λ)∝exp

−β 2

Z

,Λ)ηL Λ

Λ

(41) where the matrix ηL is given in [25]. Splitting Λ into its real and imaginary parts ΛR and ΛI, which turn out to be independent random variables, we then obtain the probability distribution for ΛR:

P(ΛR)∝exp

−β Z

ΛRR

(42) with

H=h1+ 3gψ20−µ . (43)

(7)

6 The relevant part ofg(2)II, first line in (40), can then be ex-

pressed in terms of ΛR only. Passing through the eigen- states of H and proceeding as we have done to obtain (33)-(34), we then obtain

cclass = 2kBThφ|s(x)1

Hs(x)|φi (44) that is the equivalent of (34) for the interacting gas. We then have to solve the equation

Hχ=s(x)φ . (45)

In the absence of the sign function the solutionφaof (45) is known [22, 26]:

φa(r) = ∂hN0iψ0(r)

hN0i1/2µ(hN0i) (46) whereµ is the derivative of the chemical potential with respect to hN0i. This can be obtained by taking the derivative of the Gross-Pitaevskii equation (39) with re- spect tohN0i.

In the presence ofs(x), the spatially homogeneous case may be solved exactly. One finds that the solution χ differs froms(x)φa only in a layer around the planex= 0 of width given by the healing length ξ. In the trapped case, in the Thomas-Fermi limit where the radius of the condensate is much larger than ξ, we reach the same conclusion and we use the separation of length scales to calculateχ approximately. Setting

χ(r) =f(r)φa(r) (47) we obtain the still exact equation

− ~2 2m

φa

φ ∆f−~2 m

gradφa

φ ·gradf +f =s(x). (48) We expect thatfvaries rapidly, that is over a length scale ξ, in the directionxonly, so that we take ∆f ≃∂x2f and we neglect the term involving gradf. Also f deviates significantly froms(x) only at a distance.ξfromx= 0, so thatφa/φmay be evaluated inx= 0 only. This leads to the approximate equation

− 1

κ2(y, z)∂x2f+f =s(x), (49) with

~2κ2(y, z)

2m = φ(0, y, z)

φa(0, y, z)≃2(µ−U) (50) whereU is the trapping potential and the approximation of κ was obtained in the Thomas-Fermi approximation gψ02≃µ−U. Equation (49) may be integrated to give

χ(r)≃s(x)φa(r)h

1−e−κ(y,z)|x|i

. (51)

Neglecting the deviation ofχ from s(x)φa in the thin layer aroundx= 0, we finally set

χ(r)≃s(x)φa(r) (52) and obtain the simple result

cclas ≃ kBT

hN0(hN0i). (53) For harmonic trapping where µ scales as hN0i2/5, and assuming that hN0i is well approximated by the ideal gas formula (20), we then obtain:

Var (NL−NR)≃ hNi

"

1 + kBTc 2 5µ(hNi)

1−T3

Tc3 3/5

T Tc

# . (54) The analytic prediction (54) is plotted as a full line (black and blue for two different atom numbers) in Fig.3. We note a good agreement with the numerical simulation.

From (54) we can extract the position of the maximum in the normalized fluctuations:

T Tc

max

=9801/3

14 ≃0.709 (55)

as well as their amplitude:

[Var (NL−NR)]max≃ hNi

1 + 1.36 kBTc

µ(hNi)

. (56) Note the very weak dependence ofkBTc/µ(hNi) on the atom number, scaling as hNi−1/15 [27]. We point out that our analysis in the interacting case is quite general and can be applied to atoms in any even trapping po- tential provided that the potential does not introduce a length scale smaller than the healing lengthξ.

D. First quantum corrections to the classical field Expanding the non-condensed fields Λ(r) and Λ(r) over Bogoliubov modes:

Λ(r) Λ(r)

=X

j

bj

uj(r) vj(r)

+bj

vj(r) uj(r)

(57) and using (40), the coefficient c giving the dominant contribution (32) to the signal for T < Tc, can be split asc=cth+c0. With the thermal contribution

cth= 2X

j

¯

nj|hφ|s(x)(|uji+|vji)|2≥0 (58)

where ¯nj = hbjbji = 1/[exp(βǫj)−1], ǫj being the en- ergy of the Bogoliubov modej, and the zero temperature contribution

c0= 2X

j

|hφ|s(x)|vji|2+hφ|s(x)|ujihvj|s(x)|φi. (59)

(8)

Performing the classical field approximation that is set- ting ¯nj = kBT /ǫj in (58), exactly gives the expression (44) ofcclass [31]. Introducing the first quantum correc- tion−1/2 to the occupation number ¯njand including the quantum contribution, we obtain the first quantum cor- rection tocclass in the interacting caseδc=c0+δcth: δc=X

j

hφ|s(x)|vjihvj|s(x)|φi−hφ|s(x)|ujihuj|s(x)|φi. (60) Using the closure relation P

j|ujihuj| − |vjihvj| = 1−

|φihφ| [22], we obtain

δc=−1. (61)

This exactly corresponds to the first correction in (26) for the ideal gas. We conclude that also in the interacting case, the first quantum correction tocclass has the effect of changing the shot noise term (the first term equal to one) in the square brackets of (54) into 1− hN0i/hNi.

VI. CONCLUSION

We have studied fluctuations in the difference of num- ber of particles in the left (x <0) and right (x >0) halves of a three dimensional harmonically trapped Bose gas as a function of the temperature across the critical temper- ature for condensation. Both for the ideal gas and the interacting gas, fluctuations are weakly super poissonian forT > Tc. If one lowers the temperature fromTc down to zero, fluctuations increase, reach a maximum and then decrease again as the non-condensed fraction vanishes.

We have solved this problem analytically for the ideal gas case, and we have found an approximate solution in the interacting case in the Thomas-Fermi limit when the temperature is larger than the quantum of oscillation in the trap. Remarkably, the local density approximation fails for this problem for the ideal gas. On the contrary, we show that the classical field approximation correctly gives the high temperature contribution to the fluctua- tions in the particle number difference both for the ideal gas and the interacting gas.

For a large ideal gas, the maximum of normalized fluc- tuations Var(NL−NR)/hNiis located atT /Tc = 2−2/3 independently on the trap oscillation frequencies and its

amplitude approximately scales asN1/3ω/ω¯ x. For the in- teracting case in the Thomas-Fermi regime the maximum of fluctuations in the relative number of particle subsists approximately for the same value ofT /Tc but its ampli- tude is strongly reduced as well as its dependence onN scaling asN−1/15.

Finally, we give a physical interpretation of the

“bump” in fluctuations forT < Tcas due to a beating be- tween the condensate modeφ0(r) and the non-condensed modesδψ(r), as it is apparent in Eqs. (29,40). For sym- metry reasons, only the antisymmetric component of the non-condensed fieldδψA(x) =−δψA(−x) contributes to Var(NL−NR). The measurement can then be seen in

-

φ0(x) - δψA(x) φ0(x) + δψΑ(x)

N

L

N

R

φ0(x) δψΑ(x)

signal

FIG. 4: The measurement can be seen as a balanced homo- dyne detection of the non-condensed field where the conden- sate acts as a local oscillator

a pictorial way as a balanced homodyne detection of the non-condensed field where the condensate field acts as a local oscillator (Fig.4). Since one or the other of these two fields vanishes forT tending to 0 orTc, the beating effect, and thus Var(NL−NR), are obviously maximal at some intermediate temperature.

Acknowledgments

We thank J. Reichel, J. Est`eve, K. Maussang and R.

Long for stimulating discussions. The teams of A.S. and Y.C. are parts of IFRAF. L. Y. acknowledges financial support from the National Basic Research Program of China (973 Program) under Grant No. 2006CB921104.

[1] M. Budde, K. Mølmer, Phys. Rev. A70, 053618 (2004).

[2] Y. Castin, “Simple theoretical tools for low dimension Bose gases”, Lecture notes of the 2003 Les Houches Spring School, Quantum Gases in Low Dimensions, M.

Olshanii, H. Perrin, L. Pricoupenko, Eds., J. Phys. IV France116, 89 (2004).

[3] Y. Castin, “Basic tools for degenerate Fermi gases”, Lec- ture notes of the 2006 Varenna Enrico Fermi School on Fermi gases, M. Inguscio, W. Ketterle, C. Salomon, Eds.,

SIF (2007).

[4] G.E. Astrakharchik, R. Combescot, L.P. Pitaevskii, Phys. Rev. A76, 063616 (2007).

[5] R. Brustein, A. Yarom, Journ. of Stat. Mech., P07025 (2008).

[6] L.S. Levitov, H.-W. Lee, and G.B. Lesovik, J. Math.

Phys.37, 10 (1996).

[7] J. Esteve, J.-B. Trebbia, T. Schumm, A. Aspect, C.I.

Westbrook, I. Bouchoule, Phys. Rev. Lett. 96, 130403

(9)

8 (2006).

[8] K. Maussang et al.in preparation; Presentation of J. Re- ichel at the BEC 2009 conference, San Feliu de Guixols, Spain.

[9] Y. Castin, p.1-136, in Coherent atomic matter waves, Lecture notes of Les Houches summer school, edited by R. Kaiser, C. Westbrook, and F. David, EDP Sciences and Springer-Verlag (2001).

[10] L. Landau, E. Lifchitz, p. 103-104, §30 in Physique th´eorique, Tome V : Physique Statistique, 4th edition, Mir (1994).

[11] For f(x, x) = exp[−α(x + x)2 − β(x − x)2] and for β, α > 0, let I ≡ R0

−∞dxR0

−∞dxf(x, x) − R0

−∞dxR

0 dxf(x, x). Using polar coordinates we ob- tainI=21αβarctan2βαβα . For β > α this is alsoI=

1αβ

hπ

412arctan2β−ααβi

=1αβ

hπ

4 −arctanq

α β

i . [12] Using (31) and (32), with the local density approxima-

tion (19) for g(1) (with ˜z = 1), one obtains cLDA =

4 π

P

l=1

arctan[12(lηx2 )1/2(1lηx2 )]

Πα(1+lηα2 ) .This result is qualita- tively incorrect: It depends onωy and ωz whereas the exact expression (13) does not. It is quantitatively incor- rect even in the isotropic case, where cLDA ∼(1/2)/ηx

forηx→0, to be compared to (21).

[13] S. Grossmann, M. Holthaus, Z. Naturforsch. A50, 921 (1995); S. Grossmann, M. Holthaus, Phys. Lett. A208, 188 (1995).

[14] From the expansion f(x) = 2 −(4√

2/π)(ηxx)1/2 + O(ηxx)3/2, the k-th order derivative of f in x = 1 is f(k)(1) ≃ −(2/π)3/2η1/2x (−1)k+1Γ(k−1/2), and the integral R1

0 f(x) is readily evaluated to leading order in ηx. This leads to (26), with the numerical coeffi- cient A = (2/π)3/2[S + π1/2/3], where S is a sum involving Bernoulli numbers B2k, S = P

k≥1Γ(2k − 3/2)B2k/(2k)! ≃ 0.1461. On the other hand, using (i) the integral representation of the Riemannζfunction, see

§9.512 withq= 1 in [28], (ii) the definition of Bernoulli numbers§9.610 in [28], and (iii) the definition of the Γ function §8.310(2) in [28], we obtain ζ(z) = π1Γ(1− z) sin(πz)P

n∈N(−1)nΓ(z+n−1)Bn/n!,∀z∈C\Z. Then the desired identity follows from the fact thatB0 = 1, B1 = −1/2, B2k+1 = 0∀k ≥ 1. An alternative tech- nique is to obtain from (36) the integral representation:

A = (2/π)3/2R+∞

0 dx x−3/2

(ex−1)−1−x−1+ 1/2 . Using the integral representation of the Riemannζfunc- tion, see §9.512 with q = 1 in [28], we obtain A =

−25/2ζ(−1/2)/π.

[15] In principle in (31) one should have introduced the first-order coherence function of the non-condensed field g(1)(r,r)− hN0iφ(r)φ(r), but this does not affect the result due to symmetry reasons.

[16] Yu. Kagan, B.V. Svistunov, and G.V. Shlyapnikov, Sov.

Phys. JETP75, 387 (1992); Yu. Kagan and B. Svistunov, Phys. Rev. Lett.793331 (1997).

[17] K. Damle, S.N. Majumdar and S. Sachdev, Phys. Rev.

A54, 5037 (1996).

[18] M. Fattori, C. D’Errico, G. Roati, M. Zaccanti, M. Jona- Lasinio, M. Modugno, M. Inguscio, G. Modugno, Phys.

Rev. Lett.100, 080405 (2008).

[19] K. G´oral, M. Gajda, K. Rz¸a˙zewski, Opt. Express8, 92 (2001); D. Kadio, M. Gajda and K. Rz¸a˙zewski, Phys.

Rev. A72, 013607 (2005).

[20] M.J. Davis, S.A. Morgan and K. Burnett, Phys. Rev.

Lett.87, 160402 (2001).

[21] E. Mandonnet PhD thesis University Paris 6, URL http://tel.archives-ouvertes.fr/tel-00011872/fr/ . [22] Y. Castin, R. Dum, Phys. Rev. A57, 3008 (1998).

[23] The accuracy of the Bogoliubov approximation to calcu- late the pair distribution function was checked by exact Quantum Monte Carlo calculations for a trapped gas [24].

[24] M. Holzmann, Y. Castin, Eur. Phys. J. D7, 425 (1999).

[25] A. Sinatra, C. Lobo, Y. Castin, J. Phys. B 35, 3599 (2002).

[26] P. Villain, M. Lewenstein, R. Dum, Y. Castin, L. You, A.

Imamoglu, T.A.B. Kennedy, Journal of Modern Optics, 44, 1775 (1997).

[27] Using the formulation in terms of Bogoliubov modes pre- sented in section V D, one can show that the scaling cclass ∝ kBT /µ for kBT ≫ ~ωα can also be obtained from an hydrodynamic approximation to the Bogoliubov mode functionsujandvj. Introducing the hydrodynam- ical variablesρandSsuch that the time-dependent con- densate fieldψ=ρ1/2exp(iS/~), one linearizes the time- dependent Gross-Pitaevskii equation around the station- ary solution ψ0. Differentiating ρ = ψψ and ψ/ψ = exp(2iS/~), and settingδψ=ujandδψ=vj, we obtain δρj0(uj+vj) and δS/~= (uj−vj)/(2iψ0).For real mode functions the normalization conditionR

u2j−vj2= 1 together with Euler’s equation−iǫjδSj/~=−gδρjgives the normalization condition 2gǫ

j

R

U(r)<µd3r δρ2j = 1.Let us consider for simplicity the isotropic case (see e.g. [29]

for the general anisotropic case). In the hydrodynamic approximation, δρis a product of a polynomial P(r/R) where R is the Thomas-Fermi radius, a spherical har- monic and a normalization factorNj. The coefficients of the polynomial are numbers that were calculated in [30].

We then find thatNj2scales asǫj/(gR3) so that the ma- trix element squared|hφ|s(x)(|uji+|vji)|2scales asǫj/µ.

On the other hand ¯nj ≃kBT /ǫj for the low frequency modes. We then find the scalingcclass ∝kBT /µ.

[28] I.S. Gradshteyn and I.M. Ryzhik,Table of integrals, Se- ries, and Products, 5th edition, Academic Press (San Diego, 1994).

[29] S. Sinha, Y. Castin, Phys. Rev. Lett.87, 190402 (2001).

[30] S. Stringari, Phys. Rev. Lett.77, 2360 (1996).

[31] This may be checked explicitly by calculating the inverse ofηLfrom the spectral decomposition (ηL)−1=L1η= P

j(1/ǫj) |uji

|vji

(huj|,hvj|) + |vji

|uji

hvj|,huj|

withη=

1 0 0 −1

, see e.g. [22].

Références

Documents relatifs

We report on measurements of the critical temperature of a harmonically trapped, weakly inter- acting Bose gas as a function of atom number.. Our results exclude ideal-gas behavior

This allows us to separate spatially the condensed and normal components in time of flight, and to measure reliably temperatures as low as 0.2 T c 0 and thermal fractions as low

L’archive ouverte pluridisciplinaire HAL, est destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des

4: (Color online) Normalized cloud radius R TF /R ∗ (filled purple circles) and normalized atom number (open black squares) as a function of the inverse scattering length a ho /a at

Here, using fully non-perturbative semi- classical field simulations and a powerful formulation of Bogoliubov theory in terms of the time dependent con- densate phase operator [12],

In section II, we recapitulate the general determinantal properties of non- interacting fermions in a harmonic potential and present the first derivation of the limiting distribution

In 4.1 we present a modulus-phase reformulation of the Bogoliubov theory generalized for a bimodal condensate; in 4.2 and 4.3 we derive the expansion of the spin squeezing parameter

3 we plot the barrier height Ef (relative to the spherical rinimum) versustemperature T. We see that it leads, indeed, to a drastic overestimation of the finite-temperature