You are on page 1of 5

Monopole excitations of a harmonically trapped one-dimensional Bose gas from the

ideal gas to the Tonks-Girardeau regime


S. Choi,1 V. Dunjko,1 Z. D. Zhang,2, 1 and M. Olshanii1
1
Department of Physics, University of Massachusetts, Boston, MA 02125, USA
2
Department of Physics and Astronomy, SUNY, Stony Brook, NY 11794 USA
Using a time-dependent modified nonlinear Schrödinger equation (m-NLSE) — where the con-
ventional chemical potential proportional to the density is replaced by the one inferred from Lieb-
Liniger’s exact solution — we study frequencies of the collective monopole excitations of a one-
dimensional (1D) Bose gas. We find that our method accurately reproduces the results of a recent
experimental study [E. Haller et al., Science 325, 1224 (2009)] in the full spectrum of interaction
regimes from the ideal gas, through the mean-field regime, through the mean-field Thomas-Fermi
regime, all the way to the Tonks-Giradeau gas. While the former two are accessible by the standard
arXiv:1412.6855v2 [cond-mat.quant-gas] 19 Jan 2015

time-dependent NLSE and inaccessible by the time-dependent local density approximation (LDA),
the situation reverses in the latter case. However, the m-NLSE treats all these regimes within a
single numerical method.

PACS numbers: 67.85.-d, 02.60.Cb

The study of excitations of a material allows us to un- the crossover between different regimes.
derstand its underlying nature and forms the basis of Our numerical method of choice is the modified non-
various spectroscopic methods. In particular, collective linear Schrödinger equation (m-NLSE), where the con-
excitations of ultracold atoms provide a way to infer their ventional chemical potential proportional to the density
character, including the nature of their interatomic in- is replaced by the one inferred from Lieb-Liniger’s ex-
teractions. The goal of this paper is to simulate the act solution. Our m-NLSE unifies the standard time-
monopole oscillations for a Bose gas in a one-dimensional dependent NLSE valid in the ideal gas limit and in the
(1D) harmonic oscillator (HO) potential for all range of neighboring mean-field regime (both before the validity
interaction strengths to demonstrate the continuous tran- of the Thomas-Fermi approximation and in the Thomas-
sition from the bosonic ideal gas and mean-field regimes Fermi regime) and the time-dependent LDA valid for
at weak interaction to the fermionic strongly correlated the mean-field Thomas-Fermi regime, Tonks-Girardeau
limit at large interaction strength. A number of experi- regime, and in between. It should be noted that un-
ments on the excitations of 1D bose gas exist[1–3], and like, for instance, experiments with BECs, not many ex-
yet a unified theoretical description over all interaction periments have so far been performed in the broad area
regimes has not been available. outside of the weakly interacting mean field regime that
So far the experimental regime between Thomas-Fermi would allow us to compare and help refine the numer-
and Tonks-Girardeau that shows√the crossover of the os- ical approach; the experiment of Ref. [2] is a notable
cillation frequency from ω = 3ωz to ω = 2ωz has exception. Granted, our method is more than a naive
been described by the local density approximation (LDA) interpolation between the standard NLSE at weak inter-
combined with sum rules[4], while very recently, it was actions and the LDA at the strong ones, since there ex-
shown that a Hartree approach allows for an accurate ists a parameter region—the mean-field Thomas-Fermi
description of Gaussian Bose-Einstein condensate (BEC) regime—of overlapping ranges of validity of the above
to Thomas-Fermi regimes that √ shows the crossover of fre- methods. Nevertheless, more experimental input would
quency from ω = 2ωz to ω = 3ωz [5]. In the above work, help further justify our approach.
it is further shown that while the two predictions can be Our paper extends Ref. [6] in that we go beyond the
joined almost continuously for the cases involving more steady state solutions and study the dynamics of the har-
than 25 particles, ab initio diffusion Monte Carlo calcu- monically trapped 1D Bose gas over all range of inter-
lations provide a complete all-regimes data for smaller actions, and differs from Ref. [4] in that we numerically
numbers of atoms. In our Letter, we suggest a simple simulate the oscillations directly without resorting to sum
unified numerical description for macroscopic numbers rules which give the upper limit to the excitation frequen-
of atoms in all experimental regimes, using a single wave cies calculated based on the static wave functions. It is
equation. We compare our results with the findings of also important to mention that unlike in [6] and [4], our
a 2009 experiment of Nägerl group[2] that dealt with ul- work goes beyond the local density approximation, and
tracold Cs atoms in a 1D harmonic trap. The authors in doing so is able to capture the system’s behavior at
of the experiment tuned the interaction of the 1D ultra- very low values of the interaction strength.
cold atomic gas via Feshbach resonance while measuring For weakly interacting ultracold atoms such as typical
the change in the ratio of the oscillation frequencies of experimentally produced BECs, the mean field descrip-
the collective compression (ωC ) and dipole (ωD ) modes, tion of the Gross-Pitaevskii Equation (GPE) has been
R = (ωC /ωD )2 . This ratio provides the diagnostics for found to be very effective in describing their collective
2

oscillations[7]. More generally, the time-dependent LDA actual scattering length a. This means that the strongly
(or hydrodynamic) equations were found to work when correlated beyond mean field regime is approached when
the atoms are not necessarily weakly interacting, i.e. be- the density n decreases or when a1D < d where d = 1/n is
yond mean field[8, 9], and even in the limit of strong the interatomic distance. The chemical potential for this
correlations of Tonks-Girardeau (TG) regime where the system is calculated using µl = ∂[nǫ(n)]/∂n where the
strong interaction between bosons mimic the Pauli exclu- energy per particle ǫ(n) = h̄2 n2 e(γ(n))/2m comes from
sion principle so as to make them behave as if they are solving the Lieb-Liniger system of equations that arise
free fermions[4]. from applying the Bethe Ansatz[10]. Here, γ(n) is the
The time-dependent LDA equations are: dimensionless Lieb-Liniger interaction parameter propor-
tional to the interaction strength g1D : γ(n) = 2/n|a1D |.
∂ In the 1D mean field and TG regimes the ground
n + ∇(vn) = 0 (1)
 ∂t  state of the LDA (or hydrodynamical) equations can be
∂ 1 2 found analytically. In the mean field limit of na1D ≫ 1,
m v + ∇ µ + Vext (r) + mv = 0 (2)
∂t 2 γ → 0 leading to the well known Thomas-Fermi (TF)
energy functional; there, the chemical potential is given
where n is the density of gas, v is the velocity field, by µ(n) ≈ g1D n = γh̄2 n2 /m. We therefore refer to the
µ = µl [n(r, t)] is the local density-dependent chemical 1D mean field limit as the TF limit. In the TG limit
potential, µl is the chemical potential calculated for a of na1D ≪ 1 where γ → ∞, the chemical potential is
uniform gas at density n(r, t), and Vext (r) is the external µ(n) ≈ π 2 h̄2 n2 /2m. In between the TF and TG limits
confining potential. It is noted that for weakly inter- the chemical potential has to be worked out numerically
acting BECs with negligible quantum depletion, n gives from the Lieb-Liniger system of equations[10] for each
the density of the condensate itself (rather than the to- γ(n), as exact analytical expression cannot be obtained.
tal density of atoms – condensate plus non-condensates). Near the TG limit a more accurate analytical approxi-
Physically, LDA typically corresponds to the case of zero mation for µ(n) can be obtained via large γ expansion
temperature, large N limit, and “macroscopic” dynam- methods[10]:
ics where LDA produced length scales are much larger
π 2 h̄2 2 γ(n)2 [2 + 3γ(n)]
 
than both the interparticle distance and the healing
length. These conditions, although they superficially ap- µ(n) ≈ n . (4)
2 m 3[2 + γ(n)]3
pear strict, are found not too difficult to meet in practice.
The excitation of ultracold atoms in the framework For comparison with experiments, it is convenient to
of quantum hydrodynamics is modeled analytically by define the effective γ instead of γ(n) via the maximum
putting all the time-dependence in the deviation from the steady state density of the atomic cloud at z = 0 in the
steady-state value: n(r, t) = n0 (r)+δn(r,
  t) and µ(r, t) = TG limit and the actual a1D :
∂µl
µl [n0 (r)]+δµl (r, t) where δµl (r, t) = ∂n δn. The 2 π
n=n0 γef f. = = (5)
¨ = ∇ [n∇δµl (r, t)] /m
equation to be solved is then δnn nT G (0)|a1D | α
where the double dot denotes the second order time p
derivative, and from this the corresponding excitation where nT G (0) = 2N mωz /h̄/π is the analytical TG
frequencies can be worked out[8] (see also [9] for more density in the center of the trap[6], and for con-
details). In order to find the required chemical poten- venience
p we defined dimensionless parameter α =
tial for 1D Bose gas in HO potential encompassing all N mωz /2h̄|a1D | that parametrizes the regimes of inter-
regimes of quantum correlations, we use the formalism of action strength. This naturally introduces a set of γ in-
Lieb and Liniger[10] that considers the Hamiltonian with dependent of the density profile, and we shall use γef f. as
zero range 1D repulsive potentials: our parameter in our simulation. In Fig. 1 the chemical
potential per atom, µ(γef f. )/N in HO energy units (h̄ωz )
N
h̄2 X ∂ 2 X N calculated from the Lieb-Liniger system of equations[10]
H =− 2
+ g1D δ(zi − zj ) (3) is presented as a function of log10 γef f. . The near TG
2m i=1 ∂zi i<j formula of Eq. (4) is also shown in the figure as a com-
parison. It is found that the general shape of µ(γef f. )/N
A more recent work [11] has shown that the interaction is independent of the number of atoms N , and it is seen
strength g1D is related to the 1D scattering length a1D that as γef f. → ∞, µ/N → 1 as to be expected for the
2h̄2
as g1D = ma 1D
where the interplay between strong in- TG limit. The analytical expression of Eq. (4) is seen to
teractions and confinement to low dimensional geome- be a good approximation for γ > 10 and also for γ < 0.01.
try amplifies the effect of quantum correlations, and that
given an experimentally relevant cigar-shaped confining For simulation purposes, we turn the time-dependent
trap with longitudinal frequency ωz and transverse fre- LDA equations Eqs. (1) and (2) into modified non-
quency ω⊥ , a1D = (−a2⊥ /2a)[1 − C(a/a⊥ )], where a⊥ = linear Schrödinger Equation (m-NLSE) for the wave
n(r, t)eiΦ(r,t) , where we write
p p
2h̄/mω⊥ is the size of the transverse ground state wave function Ψ(r, t) =
function and C = 1.4603, is inversely proportional to the the density n(r, t) = |Ψ(r, t)|2 and the velocity field
3

1
calculated numerically from the Lieb-Liniger system of
equations[10] at each spatio-temporal step as ñ(z, t)
0.8 evolves.
The simulation was done by first finding the ground
0.6 state solution for various values of α starting from the
µ(γeff.)/N

ground state for a HO (N = 0) and adiabatically in-


0.4 creasing N up to the desired number of atoms. Once
the steady state solutions are found, the monopole ex-
0.2
citation can then be simulated in many different ways,
including an addition by hand of the exact Bogoliubov
0
excitation modes or sinusoidal driving of the confining
−1.5 −1 −0.5 0 0.5
log10 γeff.
1 1.5 2
potential. In this paper, we directly excite monopole
oscillations by suddenly quenching the confining poten-
FIG. 1: Chemical potential per atom, µ(γef f. )/N calculated tial, from V (z) = 21 mωz2 z 2 to V (z) = 1.25 × 21 mωz2 z 2 for
from the Lieb-Liniger system of equations[10] as a function some short time (τ = 0.25π/ωz ) then back in the orig-
of log10 γef f. (Solid line) and the near TG limit expression of inal trap frequency. The simulation was then run until
Eq. (4) (Dashed line) The chemical potential is in HO units τ = 400π/ωz while measuring the time-dependent width
while γef f. is a dimensionless parameter. (variance) of the wave function
Z Z 2
v(r, t) = ∇Φ(r, t)[12, 13]. Here, the standard NLSE 2 2
h∆z i = z |ψ(z, t)| dz −2 2
z|ψ(z, t)| dz . (8)
nonlinear term, g1D |Ψ(r, t)|2 Ψ(r, t) is replaced by µ(n =
|Ψ(r, t)|2 )Ψ(r, t). For sufficiently smooth density distri-
butions, the two equations are equivalent. Furthermore, It was found that except for a short transient, the width
besides being better numerically tractable (e.g. the sharp h∆z 2 i follows a sinusoidal variation over time, owing to
edges of the atomic clouds are automatically regularized), the harmonic confining potential. From this sinusoidally
the m-NLSE offers the following benefit. At very low den- varying time-dependent width, the Fourier frequency
sities, where the size of the cloud become comparable to components were obtained numerically. We have con-
the size of the one-body quantum ground state of the firmed our procedure outlined above, particularly with
trap, the time-dependent LDA fails while the standard respect to finding the steady state solutions, the quench-
NLSE is naturally valid there; but this is exactly what ing method to excite monopole oscillations, and the accu-
the m-NLSE converts to. Thus, the m-NLSE allows for racy of the frequency measurement, by carrying out our
a formal numerical unification of the two regimes. simulation using Eq. (4) as the chemical potential and
Note that while at the ideal gas point and in the subse- finding excellent agreement with the full ab initio many-
quent mean-field regime, the m-NLSE correctly describes body calculation prediction for the monopole oscillation
the density evolution at all length scales, in the strongly frequency near the TG limit[15] given by
correlated regime, the m-NLSE must be regarded merely  2 " !#2
as a simulator for the time-dependent LDA equations, ωC 1 192 1
= 2− +O 2 . (9)
and any features of a healing length size or smaller must ωD γef f. 45π γef f.
be treated as artifacts of the computational method. This
question is discussed in Ref.[14]: it is shown in particular We plot in Fig. 2 the steady state density nss (z)
that the interference fringes produced by the m-NLSE in with atom number N = 25 for γef f. ≈ 0.01, 1, and
the TG regime have nothing to do with reality. γef f. → ∞ and the corresponding chemical potential
We are going to work with a wavefunction ψ(r, t) = µ(z). The position-dependent chemical potential gives
R +∞
N −1/2 Ψ(r, t) normalized to unity, −∞ |ψ|2 dz = 1, and, an idea of the effective potential experienced by the
wave function due to interatomic interaction. This func-
accordingly, with a probability distribution ñ = n/N =
tion was found to almost vanish for γef f. < 1, leav-
|ψ|2 . Also, in our case, the external potential is given by
ing an effectively interaction-free system of atoms. We
a one-dimensional harmonic potential Vext (r) = 12 mωz2 z 2 .
also plot the steady state harmonic oscillator energy
Accordingly, we are going to work in the harmonic oscil- h
∂2
i
EHO = ψ ∗ (z) − 12 ∂z 1 2
R
lator system of units: h̄ = m = ω = 1. The m-NLSE 2 + 2z ψ(z)dz, interaction en-
equation becomes
R ∗
ergy EI = ψ (z)µ(z)ψ(z)dz, and the total energy
EK + EI as a function of log10 γef f. . Additionally, we
1 ∂2
 
∂ψ(z) 1 2 plot as a function of log10 γef f. the initial, maximum and
i = − + z + µ[ñ(z, t)] ψ(z) (6)
∂t 2 ∂z 2 2 minimum width of the wave function h∆z 2 i attained dur-
ing the monopole oscillation. It is not surprising that the
where the chemical potential is given by initial width of the wave function follows the trend of the
total energy of the system as a function of γef f. since
N2 2
 
∂ the increasing repulsion between the atoms makes the
µ[ñ(z, t)] = ñ 3 + ñ e(γ(n = N ñ)) (7)
2 ∂ ñ wave function profile wider. The amplitude of oscillation
4

15 30
In conclusion, we found—using a direct comparison
with the experimental data[2]—that using a single triv-
10 20
ially modified nonlinear Schrödinger equation one can
n (z)

µ(z)
ss

consistently simulate the 1D Bose gas in the full spectrum


5 10

0 0
−10 0 10 −10 0 10
z z 4

30 400 3.9

300 3.8
20 4
<∆z >
Energy

2 3.7
200
3.8

2
(ωC/ωD)
10 3.6
100
3.5 3.6
0 0
−1 0 1 2 −1 0 1 2 3.4 3.4
log γ log γ
10 eff. 10 eff.
3.3 3.2

FIG. 2: Top left panel: Plot of the steady state density 3.2 3
nss (z) = |ψ(z)|2 vs. z with atom number N = 25 for −3 −2 −1 0 1 2
3.1 log10 γeff.
γef f. ≈ 0.01 (dotted line), 1 (dashed line), and γef f. → ∞
(solid line) Top right panel: the corresponding chemical po- 3
0 50 100 150 200 250 300
tential µ(z). It is noted that µ(z) for γef f. ≈ 0.01 (dotted γeff.
line) is vanishingly small. Bottom left panel: Plot of the
harmonic oscillator energy EHO (solid line), interaction en- FIG. 3: (Color online) Plot of the monopole oscillation
ergy EI (dashed line) and the total energy EK + EI (dotted frequency squared (ωC /ωD )2 as a function of γef f. =
line) as defined in the text as a function of log10 γef f. Bot- p
tom left panel: the initial width of the wave function h∆z 2 i 2π/ 2N mωz /h̄|a1D |. Inset: same figure as a function of
(dotted line), minimum h∆z 2 i (dashed line), and maximum log10 γef f. to magnify the region near γef f. ≈ 0—the region
h∆z 2 i (solid line) attained during the monopole oscillation as that lies outside of the range of validity of LDA. Stars with er-
a function of log10 γef f. . The harmonic oscillator units are ror bars: the experimental data from Nägerl group[2]. Green
used throughout. squares: numerical simulation, Red crosses: sum rules result
[4]. It must be mentioned that in the region of γef f. ≪ 1, the
result [4] is not expected to match the experiment since it has
is seen to also grow as a function of γef f. ; however tak- been built on LDA a priori.
ing into account the change in the initial width itself the
oscillation amplitude remains constant at approximately
20% of the initial width regardless of γef f. . of interaction regimes from the ideal gas, through the
Experimentally, the group of Nägerl started with a mean-field regime, through the mean-field Thomas-Fermi
BEC of between 1 × 104 to 4 × 104 Cs atoms and by regime, all the way to the Tonks-Giradeau gas. While the
using 2D optical lattice that forms an array of vertical former two are accessible by the standard time-dependent
tubes, trapped them in about 3000 to 6000 1D tubes with NLSE and inaccessible by the time-dependent local den-
8 to 25 atoms in the center tube. Our numerical simu- sity approximation (LDA), the situation reverses in the
lation parameters are within the range of experimental latter case. At the same time, the m-NLSE treats all
parameters: we cover the same range of γef f. as in the these regimes using a single numerical method. In addi-
experiment and we use N = 25. In Fig. 3 we plot the ex- tion, it is also interesting to note here that the hydrody-
perimentally measured frequencies (ωC /ωD )2 with error namic simulation was found to agree very well with the
bars as a function of γef f. and superpose the results from results of many-body calculation that considers essen-
our simulation as well as the prediction. The near-ideal tially microscopic oscillations in the near TG limit[15].
gas region corresponds to the frequency interval from As is usually the case in developing viable theoretical
(ωC /ωD )2 ≈ 4 to 3. The point (ωC /ωD )2 ≈ 3 (γ ≈ 1) is models, more experiments are needed to be done and
the mean-field Thomas-Fermi point. For higher γef f. , the compared with our NLSE to establish the range of valid-
system slowly approaches a TG plateau of (ωC /ωD )2 ≈ 2. ity of our approach. However, many experimental situa-
The sum rule formula of Ref. [4], which was built using tions involving large amplitude motion, such as problems
LDA, works well in both mean-field Thomas-Fermi and in quantum transport should satisfy the LDA and hence
in the TG regimes, but naturally fails for the near-ideal render the NLSE of Eqs. (6) and (7) a fully valid theo-
gas. Our approach however captures it. retical model.

[1] H. Moritz, T. Stöferle, M. Köhl, and T. Esslinger Phys. [2] E. Haller et al., Science 325, 1224 (2009).
Rev. Lett. 91, 250402 (2003).
5

[3] B. Fang, G. Carleo, and I. Bouchoule Phys. Rev. Lett. [9] K. Merloti, R. Dubessy, L. Longchambon, M. Olshanii,
113, 035301 (2014). and Hélène Perrin, Phys. Rev. A 88, 061603 (2013).
[4] C. Menotti and S. Stringari, Phys. Rev. A 66, 043610 [10] E. H. Lieb and W. Liniger Phys. Rev. 130, 1605 (1963);
(2002). E. H. Lieb ibid. 130, 1616 (1963).
[5] A. I. Gudyma, G. E. Astrakharchik, and M. B. Zvonarev [11] M. Olshanii Phys. Rev. Lett. 81, 938 (1998).
arXiv:1412.4408 [12] S. Stringari, Phys. Rev. Lett. 77, 2360 (1996).
[6] V. Dunjko, V. Lorent, and M. Olshanii Phys. Rev. Lett. [13] S. Stringari, Phys. Rev. A 58, 2385 (1998).
86, 5413 (2001). [14] M. D. Girardeau and E. M. Wright, Phys. Rev. Lett. 84,
[7] F. Dalfovo, S. Giorgini, L. P. Pitaevskii, and S. Stringari 5239 (2000).
Rev. Mod. Phys. 71, 463 (1999). [15] Z. D. Zhang et al. Phys. Rev. A 89, 063616 (2014).
[8] L. Pitaevskii and S. Stringari, Phys. Rev. Lett. 81, 4541
(1998).

You might also like