Download arXiv:1601.06197v1 [cond-mat.quant

Survey
yes no Was this document useful for you?
   Thank you for your participation!

* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project

Document related concepts

Quantum state wikipedia , lookup

Coherent states wikipedia , lookup

Max Born wikipedia , lookup

Bell test experiments wikipedia , lookup

Renormalization wikipedia , lookup

Canonical quantization wikipedia , lookup

Double-slit experiment wikipedia , lookup

Scalar field theory wikipedia , lookup

Hydrogen atom wikipedia , lookup

Matter wave wikipedia , lookup

Wave–particle duality wikipedia , lookup

History of quantum field theory wikipedia , lookup

Relativistic quantum mechanics wikipedia , lookup

Hidden variable theory wikipedia , lookup

Scale invariance wikipedia , lookup

T-symmetry wikipedia , lookup

Theoretical and experimental justification for the Schrödinger equation wikipedia , lookup

Atomic theory wikipedia , lookup

Renormalization group wikipedia , lookup

Transcript
Formation of Bose–Einstein condensates
Matthew J. Davis,1, 2, ⇤ Tod M. Wright,1 Thomas Gasenzer,3 Simon A. Gardiner,4 and Nick P. Proukakis5
arXiv:1601.06197v1 [cond-mat.quant-gas] 22 Jan 2016
1
School of Mathematics and Physics, The University of Queensland, St Lucia QLD 4072, Australia
2
JILA, 440 UCB, University of Colorado, Boulder, Colorado 80309, USA
3
Kirchho↵-Institut für Physik, Universität Heidelberg,
Im Neuenheimer Feld 227, 69120 Heidelberg, Germany
4
Joint Quantum Centre (JQC) Durham-Newcastle, Department of Physics,
Durham University, Durham DH1 3LE, United Kingdom
5
Joint Quantum Centre (JQC) Durham-Newcastle,
School of Mathematics and Statistics, Newcastle University,
Newcastle upon Tyne NE1 7RU, United Kingdom
The problem of understanding how a coherent, macroscopic Bose–Einstein condensate (BEC)
emerges from the cooling of a thermal Bose gas has attracted significant theoretical and experimental
interest over several decades. The pioneering achievement of BEC in weakly-interacting dilute
atomic gases in 1995 was followed by a number of experimental studies examining the growth
of the BEC number, as well as the development of its coherence. More recently there has been
interest in connecting such experiments to universal aspects of nonequilibrium phase transitions,
in terms of both static and dynamical critical exponents. Here, the spontaneous formation of
topological structures such as vortices and solitons in quenched cold-atom experiments has enabled
the verification of the Kibble–Zurek mechanism predicting the density of topological defects in
continuous phase transitions, first proposed in the context of the evolution of the early universe.
This chapter reviews progress in the understanding of BEC formation, and discusses open questions
and future research directions in the dynamics of phase transitions in quantum gases.
I.
INTRODUCTION
The equilibrium phase diagram of the dilute Bose gas
exhibits a continuous phase transition between condensed
and noncondensed phases. The order parameter characteristic of the condensed phase vanishes above some critical temperature Tc and grows continuously with decreasing temperature below this critical point. However, the
dynamical process of condensate formation has proved
to be a challenging phenomenon to address both theoretically and experimentally. This formation process is
a crucial aspect of Bose systems and of direct relevance
to all condensates discussed in this book, despite their
evident system-specific properties. Important questions
leading to intense discussions in the early literature include the timescale for condensate formation, and the
role of inhomogeneities and finite-size e↵ects in “closed”
systems. These issues are related to the concept of spontaneous symmetry breaking, its causes, and implications
for physical systems (see, for example, the chapter by
Snoke and Daley in this volume).
In this chapter we give an overview of the dynamics of condensate formation and describe the present
understanding provided by increasingly well controlled
cold-atom experiments and corresponding theoretical advances over the past twenty years. We focus on the
growth of BECs in cooled Bose gases, which, from a
theoretical standpoint, requires a suitable nonequilibrium formalism. A recent book provides a more complete introduction to a number of di↵erent theoretical
⇤
[email protected]
approaches to the description of nonequilibrium and nonzero-temperature quantum gases [1].
We note that the past decade has seen the observation of BEC in a number of diverse experimental systems
beyond ultracold atoms, including exciton-polaritons,
magnons, and phonons, which are covered in other chapters of this volume. Many of the universal aspects of
condensate formation also apply to these systems.
II.
THE PHYSICS OF BEC FORMATION
The essential character of the excitations and collective response of a condensed Bose gas is well described by
perturbative approaches that take as their starting point
the breaking of the U (1) gauge symmetry of the Bose
quantum field. This approach can be extended further
to provide a kinetic description of excitations in a condensed gas weakly perturbed away from equilibrium [2].
The description of the process of formation of a BoseEinstein condensate in a closed system begins, however,
in the opposite regime of kinetics of a non-condensed gas.
Over the past decades, there have been many studies using methods of kinetic theory to investigate the initiation
of Bose–Einstein condensation. It is now well established
that these descriptions break down near the critical point,
and in particular in any situation in which the formation process is far from adiabatic. A number of di↵erent
theoretical methodologies have been applied to the issue
of condensate formation, but most have converged to a
similar description of the essential physics. The prevailing view is that a classical non-linear wave description
— a form of Gross–Pitaevskii equation — can describe
2
the nonequilibrium dynamics of the condensation process, which involves in general aspects of weak-wave turbulence and, in more aggressive cooling scenarios, strong
turbulence. The classical field describes the highly occupied modes of the gas at finite temperature and out of
equilibrium.
A summary of the consensus picture of condensate
formation in a Bose gas cooled from above the critical
temperature is as follows. Well above the critical point
the coherences between particles in distinct eigenstates
of the appropriate single-particle Hamiltonian are negligible and the system is well described by a quantum
Boltzmann kinetic equation for the occupation numbers
of these single-particle modes. As cooling of the gas proceeds due to inter-particle collisions and interactions with
an external bath, if one is present, the occupation numbers of lower-energy modes increase. Once phase correlations between these modes become significant, the system
is best described in terms of an emergent quasiclassical
field, which may in general exhibit large phase fluctuations, topological structures and turbulent dynamics, the
nature of which may vary over time and depend on the
specific details of the system — including its dimensionality, density, and strength of interactions. This regime
is sometimes referred to as a nonequilibrium quasicondensate, in analogy to the phase-fluctuating equilibrium
regimes of low-dimensional Bose systems [3, 4]. The eventual relaxation of this quasicondensate establishes phase
coherence across the sample, producing the state that we
routinely call a Bose–Einstein condensate.
A.
The pre-condensation kinetic regime
Early investigations of the kinetics of condensation of
a gas of massive bosons began with studies of such a system coupled to a thermal bath with infinite heat capacity,
consisting of phonons [5] or fermions [6–8]. These works
inherited ideas from earlier studies of condensation of
photons in cosmological scenarios [9]. In a homogeneous
system, condensation is signified by a delta-function singularity of the momentum distribution at zero momentum (see, e.g., Ref. [10]). Levich and Yakhot found [6]
that an initially non-degenerate equilibrium ideal Bose
gas brought in contact with a bath with a temperature
below Tc would develop such a singularity at zero momentum only in the limit of an infinite evolution time (see
also Ref. [11]). These same authors subsequently found
that the introduction of collisions between the bosons
lead to the “explosive” development of a singular peak
at zero momentum after a finite evolution time [7, 8].
They were careful to point out, however, the approximations involved in their treatment of interactions, and indeed that the development of such coherence invalidates
the assumptions underlying the quantum Boltzmann description, conjecturing that “the system in the course of
phase transition passes through a stage which may be
identified as a period of strong turbulence” [7].
Experimental attempts in the 1980s to achieve Bose
condensation of spin-polarized hydrogen (see the chapter by Greytak and Kleppner for an overview and recent developments), and excitons in semiconductors such
as Cu2 O, inspired renewed theoretical interest in Bosegas kinetics. Eckern developed a kinetic theory [12] for
Hartree–Fock–Bogoliubov quasiparticles appropriate to
the relaxation of the system on the condensed side of
the transition. Snoke and Wolfe revisited the question of
the kinetics of approach to the condensation transition
by undertaking numerical calculations of the quantum
Boltzmann equation [13]. They found in particular that
the bosonic enhancement of scattering rates in the degenerate regime o↵set the increased number of scattering events required for rethermalisation in this regime,
such that re-equilibration of a shock-cooled thermal distribution takes place on the order of three to four kinetic
collision times, ⌧kin = (⇢ vT ) 1 , where ⇢ is the particle
density, is the collisional cross section, and the mean
thermal velocity vT = (3kB T /m)1/2 . These results imply that a Boltzmann-equation description of this early
kinetic regime is valid even for short-lived particles such
as excitons, as the particle lifetime is long compared to
this equilibration timescale.
Over time a comprehensive picture of the process of
condensation of a quench-cooled gas has emerged, and
comprises three distinct stages of nonequilibrium dynamics: a kinetic redistribution of population towards lower
energy modes in the non-condensed phase, development
of an instability that leads to nucleation of the condensate and a subsequent build-up of coherence, and finally
condensate growth and phase ordering. In the midst of
increasingly intensive e↵orts to achieve Bose condensation in dilute atomic gases, by then including the new
system of alkali-metal vapours, these stages were analysed in more detail in the early 1990s, beginning with
a series of papers by Stoof [14–18], and by Svistunov,
Kagan, and Shlyapnikov [19–22].
In Ref. [19], Svistunov discussed condensate formation
in a weakly interacting, dilute Bose gas, with so-called
gas parameter ⇣ = ⇢1/3 a ⌧ 1, where a is the scattering
length. In a closed system, a cooling quench generically
leads to a particle distribution which, below some energy
scale "0 , exceeds the equilibrium occupation number corresponding to the total energy and particle content. Energy and momentum conservation then imply that a few
particles scattered to high-momentum modes carry away
a large fraction of the excess energy associated with this
over-occupation, allowing the momentum of a majority
of the particles to decrease. Should the characteristic
energy scale of the overpopulated regime be sufficiently
small, "0 ⌧ ~2 ⇢2/3 /m ⇠ kB Tc , mode-occupation numbers in this regime will be much larger than unity, and
the subsequent particle transport in momentum space towards lower energies is described by the quantum Boltzmann equation in the classical-wave limit [19–21]. This
is valid for modes with energies above the scale set by
the chemical potential µ = g⇢ ⇠ ⇣kB Tc of the ultimate
3
equilibrium state, where g = 4⇡~2 a/m is the interaction constant for particles of mass m. At lower energies, the phase correlations between momentum modes
become significant, and a description beyond the quantum Boltzmann equation is required.
We note that for open systems such as excitonpolariton condensates, the quasi-coherent dynamics of
such low-energy modes will in general be sensitive to the
driving and dissipation corresponding to the continual
decay and replenishment of the bosons. Such external
coupling can dramatically alter the behaviour of the system, and its e↵ects on condensate formation dynamics
are a subject of current research — see, e.g., Refs. [23–
25] and the chapter by Altman et al.. Hereafter, unless
otherwise specified, the theoretical developments we discuss pertain to closed systems in which the bosons undergoing condensation are conserved in number during
the formation process.
By assuming the scattering matrix elements in the
wave Boltzmann equation to be independent of the mode
energies, Svistunov discussed several di↵erent transport
scenarios within the framework of weak-wave turbulence,
in analogy to similar processes underlying Langmuirwave turbulence in plasmas [26]. He concluded that
the initial kinetic transport stage of the condensation
process evolves as a weakly non-local particle wave in
momentum space. Specifically, he proposed that the
particle-flux wave followed the self-similar form n(", t) ⇠
"1 (t) 7/6 f ("/"1 (t)), with "1 (t) ⇠ (t t⇤ )3 , and scaling
function f falling o↵ as f (x) / x ↵ for x
1, with
↵ = 7/6. Following the arrival of this wave at time
t⇤ ' t0 + ~"0 /µ2 , a quasi-stationary wave-turbulent cascade forms in which particles are transported locally,
from momentum shell to momentum shell, from the scale
"0 of the energy concentration in the initial state to the
low-energy regime " . µ where coherence formation sets
in.
The wave-kinetic (or weak-wave turbulence) stage of
condensate formation following a cooling quench was
investigated in more detail by Semikoz and Tkachev
[27, 28], who solved the wave Boltzmann equation numerically and found results consistent with the above
scenario, albeit with a slightly shifted power-law exponent ↵ ' 1.24 for the wave-turbulence spectrum. Later
dynamical classical-field simulations of the condensation
formation process by Berlo↵ and Svistunov [29] further
corroborated the above picture.
B.
The formation of coherence
It has been known for some time that a kinetic Boltzmann equation model is unable to describe the development of a macroscopic zero-momentum occupation in the
absence of seeding or other modifications [6, 13, 19]. In
any event, the quantum Boltzmann equation ceases to
be valid in the high-density, low-energy regime in which
condensation occurs. The two-body scattering receives
significantR many-body corrections once the interaction
energy g k.p dk nk of particles with momenta below a
given scale p exceeds the kinetic energy at that scale, and
these are indeed the prevailing conditions when phase coherence emerges and the condensate begins to grow [19].
In a series of papers [14, 15, 17, 18], Stoof took account
of these many-body corrections and developed a theory of
condensate nucleation resting on kinetic equations incorporating a ladder-resummed many-body T -matrix determined from a one-particle-irreducible (1PI) e↵ective action or free-energy functional. In the 1PI formalism the
propagators appearing in the e↵ective action are taken
as fixed, determined in this case by the initial thermal
Bose number distribution and the spectral properties of
a free gas.
Constructed within the Schwinger–Keldysh closedtime-path framework, the method allows the determination of the time evolution of the self-energy and thus of an
e↵ective chemical potential for the zero-momentum mode
through the phase transition. During the kinetic stage,
once the system has reached temperatures below the
interaction-renormalized critical temperature, the selfenergy renders the vacuum state of the zero-momentum
mode metastable. Stoof found that this modification of
the self-energy occurs on a time scale ⇠ ~/kB Tc and gives
rise to a small seed population in the zero mode, n0 ⇠
⇣ 2 ⇢, within the kinetic time scale ⌧kin ⇠ ~/(⇣ 2 kB Tc ).
He argued that, following this seeding, the system undergoes an unstable semi-classical evolution of the lowenergy modes. Taking interactions between quasiparticles into account he found that thepsquared dispersion
!(p)2 becomes negative for p . ~ an0 (t), i.e., below
a momentum scale of the order of the inverse healing
length associated with the density n0 (t) of the existing
condensed fraction. As a result, the condensate grows
linearly in time over the kinetic time scale ⌧kin . The
growth process eventually ceases due to the conservation
of total particle number, whereafter the final kinetic equilibration of quasi-particles takes place over a time scale
⇠ ~/(⇣ 3 kB Tc ) as discussed previously by Eckern [12], and
by Semikoz and Tkachev [28].
C.
Turbulent condensation
The semi-classical scenario of Stoof is built on the assumptions that the cooling quench has driven the system to the critical point in a quasi-adiabatic fashion, and
that the neglect of thermal fluctuations and nonequilibrium over-occupations in the self-energy is justified [18].
However, as previously pointed out in Ref. [7], a more
vigorous quench may drive the system into an intermediate stage of strong turbulence, where the coherences
between wave frequencies lead to the formation of coherent structures, such as vortices, that have a significant
influence on the subsequent dynamics. The main processes and scales governing this stage were discussed in
detail by Kagan and Svistunov [21, 22]. As a result of
4
excess particles being transported kinetically into the coherent regime (wave numbers below the final inverse healp
ing length, k . ⇠ 1 ⇠ a⇢), the density and phase of the
Bose field fluctuate strongly on length scales shorter than
⇠. The growing population at even smaller wave numbers
then implies, according to Refs. [19–21], the formation
of a quasicondensate over the respective length scales,
as the coherent evolution of the field according to the
Gross–Pitaevskii equation causes the density fluctuations
to strongly decrease at the expense of phase fluctuations.
This short-range phase-ordering occurs on a time scale
⌧c ⇠ ~/µ ⇠ ~/(⇣kB Tc ). Depending on the flux of excess particles entering the coherent regime, this leads to
quasicondensate formation over a minimum length scale
lv > ⇠ (see Sects. IV A and IV B) [30]. The phase, however, remains strongly fluctuating on larger length scales
due to the formation of topological defects — vortex lines
and rings. These vortices appear in the form of clumps of
strongly tangled filaments [31] with an average distance
between filaments of order lv . If the cooling quench is sufficiently strong to drive the system near a non-thermal
fixed point, cf. Sect. IV B, this quasicondensate is characterised by new universal scaling laws in space and time.
The work of Kagan and Svistunov laid the foundations for studying the role of superfluid turbulence in the
process of Bose–Einstein condensation. Kozik and Svistunov have subsequently elucidated the decay of the vortex tangle via the transport of Kelvin waves created on
the vortex filaments through their reconnections, which
can itself assume a wave-turbulent structure [32–35].
III.
CONDENSATE FORMATION
EXPERIMENTS
A.
Growth of condensate number
We now provide a historical overview of both experiments and theory related to condensate formation in ultracold atomic gases. The first experiments to achieve
Bose–Einstein condensation in 1995 [36, 37] reached the
phase space density necessary for quantum degeneracy
using the technique of evaporative cooling [38] — the
steady removal of the most energetic atoms, followed by
rethermalisation to a lower temperature via atomic collisions. These experiments, which concentrated on the
BEC atom number as the conceptionally simplest observable, provided an indication of the time scale for condensation in trapped atomic gases, in the range of milliseconds to seconds. This gave the impetus for the development of a quantum kinetic theory by Gardiner and Zoller
using the techniques of open quantum systems. They
first considered the homogeneous Bose gas [39], before
extending the formalism to trapped gases [40, 41]. Their
methodology split the system into a “condensate band”,
containing modes significantly a↵ected by the presence of
a BEC, and a “non-condensate band” containing all other
levels. A master equation was derived for the condensate
band using standard techniques [42], yielding equations
of motion for the occupations of the condensate mode
and the low-lying excited states contained in the condensate band. A simple BEC growth equation derived from
this approach provided a reasonable first estimate of the
time of formation for the 87 Rb and 23 Na BECs of the
JILA [36] and MIT [37] groups, respectively.
The first experiment to explicitly study the formation dynamics of a BEC in a dilute weakly interacting
gas was performed by the Ketterle group at MIT, using their newly developed technique of non-destructive
imaging [43]. Beginning with an equilibrium gas just
above the critical temperature, they performed a sudden evaporative cooling “quench” by removing all atoms
above a certain energy. The subsequent evolution led to
the formation of a condensate, with a characteristic Sshaped curve for the growth in condensate number. This
was interpreted as evidence of bosonic stimulation in the
growth process, and they fitted the simple BEC growth
equation of Ref. [44] to their experimental observations.
However, the measured growth rates did not fit the theory all that well.
Gardiner and co-workers subsequently developed an
expanded rate-equation approach incorporating the dynamics of a number of quasiparticle levels [45, 46]. This
formalism predicted faster growth rates, mostly due to
the enhancement of collision rates by bosonic stimulation, but still failed to agree with the experimental data.
One limitation of this approach was that it neglected the
evaporative cooling dynamics of the thermal cloud, instead treating it as being in a supersaturated thermal
equilibrium.
The details of the evaporative cooling were simulated
in two closely related works by Davis et al. [47] and Bijlsma et al. [48]. The former was based on the quantum
kinetic theory of Gardiner and Zoller, while the latter
emerged as a limit of the field-theoretical approach of
Stoof [17, 18] and the “ZNG” formalism previously developed for nonequilibrium trapped Bose gases [49] by
Zaremba, Nikuni, and Griffin. The latter authors used a
broken-symmetry approach to derive a quantum Boltzmann equation for non-condensed atoms coupled to a
Gross–Pitaevskii equation for the condensate [49, 50],
thereby extending their two-fluid model for trapped
BECs [50], which was based on the pioneering work of
Kirkpatrick and Dorfman [51–54]. The ZNG methodology has since been used successfully and extensively to
study a variety of nonequilibrium phenomena in partially
condensed Bose gases, such as the temperature dependence of collective excitations, as reviewed in Ref. [2]. As
this methodology is explicitly based on symmetry breaking, it cannot address the initial seeding of a BEC, or
any critical physics arising from fluctuations. However,
it can model continued growth once a BEC is present.
The works of Davis et al. [47] and Bijlsma et al. [48]
both introduced approximations to the formalisms they
were built on, assuming that the condensate grew adiabatically in its ground state, and treating all non-
5
condensed atoms in a Boltzmann-like approach. Both
papers boiled down to simulating the quantum Boltzmann equation in the ergodic approximation, in which
the phase-space distribution depends on the phase-space
variables only through the energy [55]. Despite the di↵erent approaches, the calculations were in excellent agreement with one another — yet still quantitatively disagreed with the MIT experimental data [43]. This disagreement has remained unexplained.
A second study of evaporative cooling to BEC in a dilute gas was performed by the group of Esslinger and
Hänsch in Munich [56]. In this experiment the Bose
cloud, which was again initially prepared in an equilibrium state slightly above Tc , was subjected to a continuous rf field inducing the ejection of high-energy atoms
from the sample. By adjusting the frequency of the applied field and thus the energies of the removed atoms,
these authors were able to investigate the growth of
the condensate for varying rates of evaporative cooling.
Davis and Gardiner extended their earlier approach [47]
to include the e↵ects of three-body loss and gravitational
sag on the cooling of the 87 Rb cloud in this experiment [57]. Their calculations yielded excellent agreement
with the experimental data of Ref. [56] within its statistical uncertainty for all but the slowest cooling scenarios
considered. An example is shown in Fig. 1(a).
In 1997 Pinkse et al. [58] experimentally demonstrated
that adiabatically changing the trap shape could increase
the phase-space density of an atomic gas by up to a factor of two and conjectured that this e↵ect could be exploited to cross the BEC transition in a thermodynamically reversible fashion. This scenario was subsequently
realised in the MIT group by Stamper-Kurn et al. [59]
by slowly ramping on a tight “dimple” trap formed from
an optical dipole potential on top of a weaker harmonic
magnetic trap. This experiment was the setting for the
first application of a stochastic Gross–Pitaevskii methodology [60], previously developed from a nonequilibrium
formalism for Bose gases by Stoof [18]. This is based
on the many-body T-matrix approximation, and uses the
Schwinger–Keldysh path integral formulation of nonequilibrium quantum field theory to derive a Fokker-Planck
equation for both the coherent and incoherent dynamics
of a Bose gas. The classical modes of the gas were represented by a Gross–Pitaevskii equation, with additional
dissipative and noise terms resulting from a collisional
coupling to a thermal bath with a temperature T and
chemical potential µ.
Proukakis et al. [61] subsequently used this methodology to study the formation of quasicondensates in a onedimensional dimple trap. A much later experiment [62]
investigated the dynamics of condensate formation following the sudden introduction of a dimple trap, and
included quantum-kinetic simulations that were in good
agreement with the data.
A novel method of cooling a bosonic cloud to condensation was introduced in 2003 by the Cornell group
at JILA [63], who demonstrated the evaporative cool-
ing of an atomic Bose cloud brought in close proximity
to a dielectric surface, due to the selective adsorption of
high-energy atoms. More recently, similar experiments
have been undertaken by the Durham [64] and Tübingen
groups [65], with the observed rates of loss in the latter
case explained accurately by non-ergodic ZNG-method
calculations of the evaporative cooling dynamics. Example results are shown in Fig. 1(b).
B.
Other theories for condensate formation
For completeness, here we briefly outline other theoretical methods that can be applied to condensate formation. A generalised kinetic equation for thermally excited
Bogoliubov quasiparticles was obtained by ImamovicTomasovic and Griffin [66] based on the application of
the Kadano↵–Baym nonequilibrium Green’s function approach [67] to a trapped Bose gas. This kinetic equation
reduces to that of Eckern [12] in the homogenous limit
and to that of ZNG [49] when the quasiparticle character of the excitation spectrum is neglected. Walser et
al. [68, 69] derived a kinetic theory for a weakly interacting condensed Bose gas in terms of a coarse graining of the N -particle density operator over configurational variables. Neglecting short-lived correlations between colliding atoms in a Markov approximation, they
obtained kinetic equations for the condensate and noncondensate mean fields which were subsequently shown
to be microscopically equivalent [70] to the nonequilibrium Green’s function approach of Ref. [66]. Exactly the
same kinetic equations were derived by Proukakis [71],
within the formalism of his earlier quantum kinetic formulation [72, 73], based on the adiabatic elimination of
rapidly-evolving averages of non-condensate operators,
ideas which fed into the development of the ZNG kinetic model [74]. Although elegant, these formalisms
have not provided a tractable computational methodology for modelling condensate formation away from the
quasistatic limit.
A non-perturbative method for the many-body dynamics of the Bose gas far from equilibrium has been
developed by Berges, Gasenzer, and co-workers [75–78].
This two-particle irreducible (2PI) e↵ective-action approach provides a systematic way to derive approximate
Kadano↵–Baym equations consistent with conservation
laws such as those for energy and particle number. In
contrast to 1PI methods, single-particle correlators are
determined self-consistently by these equations. This
approach allows the description of strongly correlated
systems, and has been exploited in the context of turbulent condensation [79–81] where it provides a selfconsistently determined many-body T matrix. This 2PI
e↵ective-action approach is useful for studying strongly
interacting systems such as 1D gases with large coupling
constant [82], or relaxation and (pre-)thermalization of
strongly correlated spinor gases [83].
PHYSICAL REVIEW A 90, 023614 (2014)
EVAPORATIVE COOLING OF COLD ATOMS AT SURFACES
6
105 N
105 N
possible collision partners. Because the density of the atom
1.5
cloud can vary considerably, we use an adaptive Cartesian
1.0
grid in real space as outlined in [77], while keeping a global
time step.
0.5
Our initial state is a thermal cloud in equilibrium with a
1.0
temperature T . This state is calculated using self-consistent
Hartree-Fock as outlined in [78]. In addition to the thermal
0.5 1.5
t (s)
cloud, the initial state requires a small condensate “seed”
to allow for C12 collisions, and hence condensate growth;
0.5
the number of atoms in the seed is obtained using the
Bose-Einstein distribution, assuming µc = 0 [38].
Interactions between the surface and the atoms are modeled
by calculating the single-correction function [Eq. (7)] for the
generalized Gross-Pitaevskii equation [Eq. (1)] and combining
0
0
1
2
3
it with a linear imaginary potential to remove condensate
atoms, effective from the position where the trap opens [79].
t (s)
In addition, we annihilate test particles that are beyond this
opening point, resulting in an atom loss for the thermal cloud.
FIG. 2. (Color online) Total atom number N against time t for
FIG. 1.These
(a) Growth
of an atomic
modelled
the trap-surface
quantum Boltzmann
The experiment
threewith
different
separations: xequation.
two processes
lead toBose–Einstein
a reduction in condensate
the total atom
s = 68 µm (gold solid
87
6
began with
a
Rb
Bose
gas
in
an
elongated
harmonic
trap
with
N
=
(4.2±0.2)⇥10
atoms
at
a
temperature
of
Ti =curve).
(640 ±30)
i
curve),
30
µm
(red
dashed
curve),
and
15
µm
(black
dotted
number in the system.
nK, before turning on rf evaporative cooling with a truncation energy
of 1.4k
solid and dotted
show correspond
the theoretical
B T . The
Points
correspond
to experimental
data andlines
the curves
calculations with a starting number of Ni = 4.2 ⇥ 106 and Ni = 4.4to⇥simulations.
106 atomsThe
respectively.
Taken
Ref.
[56]. (b)for
Surface
dot-dashed gold
curvefrom
shows
a simulation
IV.
RESULTS
evaporation leading to the formation of a BEC, showing the total
for three
distances.
68 µmnumber
without collisions,
i.e.,di↵erent
C12 = C22cloud-surface
= 0. The gray vertical
xs =atom
The lines Having
are the set
results
of computational
ZNG simulations,
are from dashed
experiment.
The
dot-dash
gold
line
is reaches
for a ZNG
line marks
the point
when the
atom
cloud
its finalsimulation
hold
up our
model,the
wepoints
now employ
position atthe
t = 0.
Thedynamics
gray hashedofarea
the shift
of the
neglecting
in the
thermalcooling.
cloud, demonstrating
that modelling
full
theshows
thermal
cloud
is curve
necessary
it tocollisions
study surface
evaporative
We show the results
when
thetotal
surface
position is
varied bycloud
±2.5 µm.
The inset
a
for a quantitative
understanding
of the
experiment.
inset
the
number,
thermal
number,
andshows
condensate
of simulations
for two different
geometries.
In Sec.The
IV A,
we shows
breakdown
the cloud
number,directly
from top
to bottom
respectively,
function
time. Taken
fromofRef.
[65].atom numbers against time for xs = 68 µm
compare
theory with
experimentasto aexamine
theof
extent
from the point when the cloud reaches its holding position. The solid
to which the model captures the important physical processes.
curve shows the total atom number, the dashed curve corresponds to
We then go on to consider a simpler model system in Sec. IV B,
thermal atoms, and the dotted curve corresponds to the condensate
with
a
view
to
optimizing
parameters
to
create
the
purest
or
C. Other pioneering condensate-formation
In number.
2004, the Sengstock group observed the formation
atom
largest condensates.
experiments
of a BEC at constant temperature [87]. Working with a
The experiments were performed using the apparatus
spin-1 system, they prepared a partially condensed gas
described in [17]. Clouds of 87 Rb atoms were loaded into
temperatures are slightly above the critical temperature for
consisting
of mF = ±1 states. Spin collisions within the
There
number
of experimental
other
an are
atoma chip
trap with
frequencies ωx methods
= 2π ×16 rads
s−1
condensation, Tc , for
an ideal gas [80].
BEC
components
the mthese
0 state, which
−1
F =experimental
than evaporative
cooling to
in the axial direction
andincrease
ωy = ωz the
= 2πphase-space
×85 rads s denin the
We performed the populated
simulations using
then
quickly
thermalised.
When
the
population
direction.
initially prepared
with the
sity of radial
a quantum
gasThe
andcloud
formwas
a condensate.
We briefly
parameters [81]. We plot the theoretical and experimental atomof the
m
component
reached
the series”).
criticalWe
number
trap
center
at
a
distance
x
≈
135
µm
from
a
silicon
surface,
F = 0against
numbers
time in Fig.
2 (a “time
considera new
s
mention them here for completeness.
BEC
emerged.
The
experiment
was
modelled
withitsa simdefined
as
the
x
=
0
plane.
At
this
point,
there
was
negligible
the
time
t
=
0
to
be
the
point
when
the
cloud
reaches
An experimental technique that has proved to be exoverlap between the cloud and the surface. The cloud was then
finalrate
hold equation.
position, indicated by the gray vertical dashed line.
ple
tremelytransported
useful foralong
multi-component
quantum gases is the
the x axis at a variable speed to a variable
Since the absolute surface position may vary due to drifts, we
methoddistance,
of sympathetic
which
an atomic
gas
surface andinheld
for a variable
hold time.
xs , from the cooling,
In the same
Ketterle’s
group
performed
performed
a rangeyear,
of simulations
withMIT
varying
xs to obtain
the an
is cooled
by virtue
of the
its remaining
collisional
interaction
with
a
In order
to measure
atom
number N , the
cloud
best fit. In thisinsense,
thethey
simulations
served
as a calibration
experiment
which
distilled
a BEC
from one trap
secondwas
gasswiftly
of atoms,
the after
firstwhich
either
broughtdistinguished
back to its initialfrom
position,
we
tool: for thetoxsanother
≈ 14, 29,[88].
and A
72non-zero-temperature
µm curves, the best fits BEC
minimum
isotopically
or by
internal measurements
quantum numbers,
which is
performed
time-of-flight
and CCD imaging.
were formed
obtained in
withana optical
simulateddipole
cloud-surface
was
trap, separation
before a of
second
15, 30, and 68 µm, respectively, well within the experimental
itself subject to, e.g., evaporative cooling. This techtrap with a greater potential depth was brought nearby.
uncertainties. Figure 2 shows the evolution of the total number
nique was first demonstrated
bycurves
Myatt et al. [84] in a
A. Loss
Atoms
energy
were
87
of atoms,ofNsufficient
, remainingthermal
in the cloud
during
the able
coursetoofcross
the the
gas comprising two distinct1. spin
Time states
series of Rb, and was
barrier
between
the
two
potential
minima,
populating
simulation, with curves corresponding to numerical results and the
subsequently employed to cool a single-component Fermi
second
trap. Eventually
the firstdata.
condensate evaporated,
points corresponding
to experimental
We begin by considering atom loss curves as a function of
gas to degeneracy by Schreck et al. [85].
andThe
a second
formed
in the
global
gold solidcondensate
curve and gold
star points
are new
for the
xs = trap
time when the cloud is brought into overlap with the surface.
In a Insimilar
spirit, inthe
2009
thewas
Inguscio
group
insurface
Flo68 µm hold point, the red dashed curve and red open circles
the experiments,
cloud
transported
to the
minimum.
rence used
exchange
are for the xs = 30 µm hold point, and the black dotted curve
in 1 s entropy
and held stationary
at abetween
final holdcomponents
point for up to of
2.5as.
87
41
two-species
RbK were
Boseconsidered:
gas mixture
BEC
Finally,
we mention
a xrecent
Three hold
points
xs ≈ to
14, induce
29, and 72
µm.
and
black crosses
are for the
µm hold point.by
Tothe
givegroup
s = 15 experiment
werecomponents
estimated from [86].
the pointThe
wheretwo
the trap
completely
an Schreck
idea of how
surface position
the remaining
in oneThese
of the
gases
were
of
atthe
Innsbruck,
whoaffects
demonstrated
theatom
first exopened
of the atoms by
werecooling,
lost to theafter
surface.
Reference
number, we vary
the surface
±2.5 µm
the coolbrought
closeand
to all
degeneracy
which
the
perimental
production
of position
a BECbysolely
by for
laser
measurements
thatpotential
temperature-related
drifts could
68 µm
curve,
shown
as the
hashed
area in Fig.
strength
of the 41 Krevealed
trapping
was adiabatically
ing
[89].
This
feat
wasgray
made
possible
by 2.
laser cooling
shift the position of the surface by up to 10 µm, hence the
For all values of xs , we observe a nontrivial
curve;
84
increased, by introducing an optical dipole potential to
on a narrow-linewidth transition of Sr, loss
resulting
in a
given87values for xs are approximate; this is the dominant source
the loss rates increase to a maximum as the cloud is brought
which the
Rb
component
was
largely
insensitive.
In
a
low
Doppler-limit
temperature
of
just
350
nK.
A
“lightto the surface. Once the cloud reaches its final position, the
of error. The initial cloud temperatures were 130 nK for xs ≈
single-component
this
increase
shift”
laser beam
was
at thethese
centre
of the
≈ 29 µm,
andwould
140 nKlead
for xsto≈ an
72 µm.
These
losses swiftly
reduce.
Theintroduced
transfer between
regimes
is trap
14 µm and xs system
(a)
(b)
in the temperature and leave the phase-space density unso that the atoms in that region no longer responded to
a↵ected. However, in the dual-species setup the 87 Rb023614-3
the laser cooling, after which an additional dimple trap
cloud acted as a thermal reservoir, suppressing the temwas introduced to confine the atoms. Repeatedly cycling
perature increase of the 41 K component and causing it
the dimple trap on and o↵ resulted in the formation of
to cross the BEC threshold.
several condensates [89].
7
D.
Low-dimensional Bose systems and phase
fluctuations
The experiments described above were in the threedimensional (3D) realm, in which long-wavelength phase
fluctuations are strongly suppressed away from the vicinity of the phase transition. In lower dimensional systems
such fluctuations are enhanced, leading to dramatic modifications to the physics of the degenerate regime. In a
two-dimensional (2D) system, thermal fluctuations of the
phase erode the long-range order associated with true
condensation, leaving only so-called quasi-long-range order characterised by correlation functions that decay algebraically with spatial separation [3]. A more complete
analysis reveals the importance of vortex-antivortex pairs
in this phase-fluctuating “quasi-condensed” regime [90].
Such pairs undergo a so-called Berezinskii–Kosterlitz–
Thouless (BKT) deconfinement transition at some finite
temperature, above which even quasi-long-range order is
lost and superfluidity is extinguished. Two-dimensional
Bose systems are of particular interest due to their natural realisation in systems such as liquid helium films
and the fact that the degenerate Bose quasiparticles such
as excitons and polaritons in semiconductor systems are
typically confined in a planar geometry. An insightful
overview of BKT physics can be found in the chapter by
Kim, Nitsche and Yamamoto in this volume.
There have been numerous experimental realisations
of (quasi-)2D Bose gases in cold-atom experiments [91–
96], with most notable the observations of thermally activated vortices via interferometric measurements [91] and
the direct probing of the equation of state and scale invariance of the 2D system [96] (see the chapter by Chin
and Refs. [97–101] for related theoretical considerations).
Further details and a lengthy discussion of the interplay
between BKT and BEC in homogeneous and trapped
systems can be found in Ref. [102]. Although theoretical works on the dynamics of such systems have existed
for some time, little experimental work on the formation
dynamics of condensates in these systems has been undertaken (aside from the quasi-2D Kibble-Zurek works
discussed in the following section). Considerable discussion is currently taking place regarding the emergence
and nature of the BKT transition in driven-dissipative
polariton condensates: experimentalists have observed
evidence for quasi-long-range order [103, 104] (see Kim
et al.’s chapter), but the nature of the transition and its
“nonequilibrium” features are topics of current debate
[24, 25] (see also the chapter by Keeling et al.).
In one dimension, the e↵ects of phase fluctuations are
even more pronounced, leading to the complete destruction of long-range order and superfluidity at any finite
temperature. Many experiments with cold atoms in elongated “cigar-shaped” traps have investigated the physics
of such (quasi-) one-dimensional systems, though again,
little work has been done on the formation dynamics
of these degenerate samples. We note, however, that
quasicondensate regimes somewhat analogous to those
of (quasi-) one-dimensional systems can be realized in
elongated 3D traps [105]. In such a regime, the Bose gas
behaves much as a conventional three-dimensional Bose
condensate, except that the coherence length of the gas is
shorter than the system extent along the long axis of the
trap. A study of condensate formation in this regime was
performed by the Amsterdam group of Walraven [106] in
2002 in an elongated 23 Na cloud. Similarly to the MIT
experiment [43], they performed rapid quench cooling of
their sample from just above the critical temperature.
However, the system was in the hydrodynamic regime in
the weakly trapped dimension, i.e., the mean distance between collisions was much shorter than the system length.
It was argued that the system rapidly came to a local thermal equilibrium in the radial direction, resulting
in cooling of the cloud below the local degeneracy temperature over a large spatial region and generating an
elongated quasicondensate. However, the extent of this
quasicondensate along the long axis of the trap was larger
than that expected at equilibrium, leading to large amplitude oscillations. The momentum distribution of the
cloud was imaged via “condensate focussing”, with the
breadth of the focal point giving an indication of the
magnitude of the phase fluctuations present in the sample. This interesting experiment was somewhat ahead of
its time, with theoretical techniques unable to address
many of the nonequilibrium aspects of the problem.
In 2007 the group of Aspect from Institut d’Optique
also studied the formation of a quasicondensate in an
elongated three-dimensional trap via continuous evaporative cooling [107] in a similar fashion to the earlier work
by Köhl et al. [56]. As well as measuring the condensate
number, they also performed Bragg spectroscopy during the growth to determine the momentum width and
hence the coherence length of the system. They found
that the momentum width they measured rapidly decreased with time to the width expected in equilibrium
for the instantaneous value of the condensate number.
Modelling of the growth of the condensate population
using the methodology of Ref. [57] produced results in
good agreement with the experimental data, apart from
an unexplained delay of 10–50 ms, depending on the rate
of evaporation.
IV.
CRITICALITY AND NONEQUILIBRIUM
DYNAMICS
As Bose–Einstein condensation is a continuous phase
transition, the theory of critical phenomena [108] predicts
that in the vicinity of the critical point the correlations of
the Bose field obey universal scaling relations. In particular, the scaling of correlations at and near equilibrium is
governed by a set of universal critical exponents and scaling functions, independent of the microscopic parameters
of the gas. For a homogeneous system close to criticality, standard theory predicts that the correlation length
⇠, relaxation time ⌧ , and first-order correlation function
8
G(x) = h
⇠=
†
⇠0
,
|✏|⌫
(x) (0)i obey scaling laws
⌧=
⌧0
,
|✏|⌫z
G(x) = ✏⌫(d
2+⌘)
F(✏⌫ x),
(1)
with ✏ = T /Tc 1 the reduced temperature, ⌫ and z
the correlation length and dynamical critical exponents,
⌘ the scaling dimension of the Bose field, and F a universal scaling function. The static Bose gas belongs to
the XY (or O(2)) universality class, and is thus expected
to have the same critical exponents as superfluid helium,
i.e., in 3D, ⌫ ' 0.67 and ⌘ = 0.038(4) [109]. The critical
dynamics of the system are expected to conform to those
of the di↵usive model denoted by F in the classification
of Ref. [110], implying a value z = 3/2 for the dynamical
critical exponent.
The influence of critical physics is significantly reduced
in the conditions of harmonic confinement typical of experimental Bose-gas systems, as compared to homogeneous systems. Within a local-density approximation,
the inhomogeneous thermodynamic parameters of the
system imply that only a small fraction of atoms in the
gas enter the critical regime, and so global observables
are relatively insensitive to the e↵ects of criticality. Nevertheless, a few experiments have attempted to observe
aspects of the critical physics of trapped Bose gases.
In a homogeneous gas the introduction of interparticle interactions has no e↵ect on the critical temperature
at the mean-field level, but the magnitude and even the
sign of the shift due to critical fluctuations was debated
for several decades (see Ref. [111] and references therein)
before being settled by classical-field Monte-Carlo calculations [112, 113]. An experiment by the Aspect group
carefully measured a shift in critical temperature of the
trapped gas, but was unable to unambiguously infer any
beyond-mean-field contribution to this shift [114, 115]. A
later experiment by the group of Hadzibabic made use of
a Feshbach resonance to control the interaction strength
in 41 K, and found clear evidence of a positive beyondmean-field shift [116] (see also the chapter by Smith in
this volume).
In 2007 the ETH Zürich group of Esslinger revisited
their experiments on condensate formation and the coherence of a three-dimensional BEC with a new tool: the
ability to count single atoms passing through an optical cavity below their ultra-cold gas [117]. They outcoupled atoms from two di↵erent vertical locations from
their sample as it was cooled, realising interference in
the falling matter waves. By monitoring the visibility of
the fringes, they were able to measure the growth of the
coherence length as a function of time. Using the same
optical cavity setup, the Esslinger group subsequently
measured the coherence length of their Bose gas as it
was driven through the critical temperature by a small
background heating rate, and determined the correlationlength critical exponent to be ⌫ = 0.67±0.13 [118]. Their
results are shown in Fig. 2(a). Classical-field simulations
of their experiment were in reasonable agreement, determining ⌫ = 0.80 ± 0.12 [119].
Although an important topic in its own right, the
greatest significance of the equilibrium theory of critical
fluctuations to studies of condensate formation is that it
provides a basis for generalisations of concepts such as
critical scaling laws and universality classes to the domain of nonequilibrium physics. In the remainder of this
section we discuss two such extensions: the Kibble–Zurek
mechanism (KZM), and the theory of non-thermal fixed
points.
A.
The Kibble–Zurek mechanism
The theory of the Kibble–Zurek mechanism leverages
the well-established results of the equilibrium theory of
criticality to make immediate predictions for universal
scaling behaviour in the nonequilibrium dynamics of passage through a second-order phase transition. The underlying idea — that causally disconnected regions of
space break symmetry independently, leading to the formation of topological defects — was first discussed by
Kibble [121], who predicted that the distribution of defects following the transition would be determined by
the instantaneous correlation length of the system as it
passes through the Ginzburg temperature [122]. Zurek
later emphasised [123] the importance of dynamic critical phenomena [110] in such a scenario. In particular,
the scaling relations (1) imply that both the correlation
length and the characteristic relaxation time of the system diverge as the critical point is approached (✏ ! 0),
imposing a limit to the size of spatial regions over which
order can be established during the transition. Topological defects will thus be seeded, with a density determined
by the correlation length at the time the system “freezes”
during the transition, and will subsequently decay in the
symmetry-broken phase. The more rapidly the system
passes through the critical point, the shorter the correlation length that is frozen in, and therefore more topological defects will form. A dimensional analysis predicts
that a linear ramp ✏(t) = t/⌧Q of the reduced temperature through the critical point on a characteristic time
scale ⌧Q results in a distribution of spontaneously formed
defects with a density nd that scales as [124]
(p d)⌫/(1+⌫z)
n d / ⌧Q
,
(2)
where d is the dimensionality of the sample and p is the
intrinsic dimensionality of the defects.
Zurek initially described the KZM in the context of
vortices in the -transition of superfluid 4 He [123]. Although vortices are observed in the wake of this transition, it is difficult to identify them as having formed due
to the KZM rather than being induced by, e.g., inadvertent stirring [124] (see also the chapter by Pickett in this
volume). The prospect of generating vorticity in atomic
BECs by means of the KZM was first discussed by Anglin
and Zurek in 1999 [125]. However, it was not until the
2008 experiment of the Anderson group at the University
9
8
10
6
(b)
(a)
2.0
1
5
4
0.1
3
10-3
{ HmmL
Correlation length ξ (µm)
7
10-2
2
1.0
1
0
0
0.01
0.02
0.03
0.04
0.05
0.06
0.07
Reduced temperature (T-Tc)/Tc
0.1
0.4
tQ HsL
1.6
FIG. 2. Critical phenomena in BECs. (a) Divergence of the equilibrium correlation length ⇠ as a function of the reduced
temperature, and the fitting of the critical exponent, giving the result ⌫ = 0.67 ± 0.13. Inset: Double logarithmic plot of the
same data. Taken from Ref. [118]. (b) Log-log plot of the dependence of the correlation length, here labelled `, as a function
of the characteristic time ⌧Q of the quench through the BEC phase transition. The solid line corresponds to a Kibble–Zurek
b
power-law scaling ` / ⌧Q
with b = 0.35 ± 0.04, in agreement with the beyond-mean-field prediction b = 1/3 of the so-called F
model [110] and inconsistent with the mean-field value b = 1/4. This in turn implies a value z = 1.4 ± 0.2 for the dynamical
critical exponent. Taken from Ref. [120].
of Arizona [126] that spontaneously formed vortices were
first observed in such a system (see also Ref. [127]).
The observations of spontaneous vortices in Ref. [126]
were supported by numerical simulations using the
stochastic projected Gross-Pitaevskii equation description of Gardiner and Davis [128]. Their results are
shown in Fig. 3. This formalism is essentially a variant of the Gardiner-Zoller quantum kinetic theory, obtained by making a high-temperature approximation to
the condensate-band master equation and then exploiting the quantum-classical correspondence of the Wigner
representation to obtain a stochastic classical-field description of the condensate band [128, 129]. Although derived using di↵erent theoretical techniques, the resulting
description is similar to the stochastic Gross-Pitaevskii
equation of Stoof [18, 60], both in terms of its physical content and its computational implementation —
see, e.g., discussion in Refs. [130, 131]. A related phasespace method originating in quantum optics known as
the positive-P representation has also been applied to ultracold gases [132]. This has been used to investigate
cooling of a small system towards BEC by Drummond
and Corney [133], who observed features consistent with
spontaneously formed vortices. Despite formally being
a statistically exact method, for interacting systems it
tends to su↵er from numerical divergences after a relatively short evolution time.
It seems likely that spontaneously formed vortices and
other defects were present in earlier BEC-formation experiments, but not observed due to the practical difficulties inherent in resolving these defects in experimental imaging — and indeed the fact that these experiments were not attempting to investigate whether such
structures were present. Another difficulty in identifying quantitative signatures of the KZM in experimental
BECs is the inhomogeneity of the system in the experimental trapping potential, which is typically harmonic.
From the point of view of a local-density approximation, this inhomogeneity implies that the instantaneous
coherence length and relaxation timescale are spatially
varying quantities, and that the transition occurs at different times in di↵erent regions of space as the system
is cooled. Following preliminary reports of the experimental observation of dark solitons following the formation of a quasi-one-dimensional BEC by the group of
Engels at the University of Washington [134], Zurek applied the framework of the KZM to a quasi-1D BEC in
a cigar-shaped trap to estimate the scaling of the number of spontaneously generated solitons as a function of
the quench time [135, 136]. Witkowska et al. [137] numerically studied cooling leading to solitons in a comparable one-dimensional geometry. Zurek’s methodology
for inhomogeneous systems was applied by del Campo
et al. [138] to strongly oblate geometries in which vortex
filaments behave approximately as point vortices in the
plane, an idealisation of the geometry of the experiment
of Weiler et al. [126].
Lamporesi et al. [139] recently reported the spontaneous creation of Kibble–Zurek dark solitons in the formation of a BEC in an elongated trap, and found the scaling of the number of observed defects with cooling rate
in good agreement with the predictions of Zurek [135]. It
was later realised that the apparent solitons were actually
solitonic vortices [140]. The e↵ects of inhomogeneity in
such experiments can be mitigated by the realisation of
“box-like” flat-bottomed trapping geometries. The Dalibard group in Paris has observed the formation of spontaneous vortices in a quasi-2D box-like geometry, and
found scaling of the vortex number with quench rate in
good agreement with the predictions of the KZM [141].
10
b
(b)
6
(a)
a
1
0.8
4
0.6
3
0.4
2
0.2
1
0
2
3
4
5
6
Vortex probability
N0 (105 atoms)
5
(c)
c
d
(d)
0
Time (s)
FIG. 3. Spontaneous vortices in the formation of a Bose–Einstein condensate. (a) Squares: Experimentally measured condensate
population as a function of time. Solid line: Condensate number from stochastic Gross–Pitaevskii simulations. Dashed line:
probability of finding one or more vortices in the simulations as a function of time, averaged over 298 trajectories. The shaded
area indicates the statistical uncertainty in the experimentally measured vortex probability at t = 6.0 s. It was observed in
experiment that there was no discernible vortex decay between 3.5 s and 6.0 s (b) Experimental absorption images taken after
59 ms time-of-flight showing the presence of vortices. (c) Simulated in-trap column densities at t = 3.5 s [indicated by the left
vertical dotted line in (a).] (d) Phase images through the z = 0 plane, with plusses (open circles) representing vortices with
positive (negative) circulation. Adapted from C. N. Weiler et al. [126].
We also note further work by the Dalibard group [142]
verifying the production of quench-induced supercurrents
in a toroidal or “ring-trap” geometry [143] analogous to
the annular sample of superfluid helium considered in
Zurek’s original proposal [123].
Experimental investigations of the KZM in dilute
atomic gases have largely focused on the imaging of defects in the wake of the phase transition — either following time-of-flight expansion [126, 139, 141] or in situ
[140]. However, the accurate extraction of critical scaling behaviour from such observations is hampered by the
large background excitation of the field near the transition, and the relaxation (or “coarsening”) dynamics of
defects in the symmetry-broken phase. An alternative
approach is to make quantitative measurements of global
properties of the system following the quench. Performing quench experiments in a three-dimensional box-like
geometry, the Hadzibabic group in Cambridge [120] made
careful measurements of the scaling of the correlation
length with quench time. From the measured scaling
law, these authors were able to infer a beyond-mean-field
value z = 1.4±0.2 for the dynamical critical exponent for
this universality class. Some of the results of Ref. [120]
are displayed in Fig. 2(b).
The possibilities for the trapping and cooling of multicomponent systems in atomic physics experiments have
naturally lead to investigations of the spontaneous formation of more complicated topological defects during
a phase transition. Although such experiments have sofar largely focused on the formation of defects following
a quench of Hamiltonian parameters [144, 145], the formation of nontrivial domain structures following gradual
sympathetic cooling in immiscible 85 Rb-87 Rb [146] and
87
Rb-133 Cs [147] Bose-Bose mixtures has also been observed. The competing growth dynamics of the two immiscible components in the formation of such a binary
condensate have recently been investigated theoretically
in the limit of a sudden temperature quench [148] (see
also Refs. [149–151] for related critical scaling in other
Hamiltonian quenches). These investigations indicate
the rich nonequilibrium dynamics possible in these systems, including strong memory e↵ects on the coarsening
of spontaneously formed defects and the potential “microtrapping” of one component in spontaneous defects
formed in the other.
B.
Non-thermal fixed points
A general characterisation of the relaxation dynamics
of quantum many-body systems quenched far out of equilibrium remains a largely open problem. In particular,
it is interesting to ask to what extent analogues of the
universal descriptions arising from the equilibrium theory of critical fluctuations may exist for nonequilibrium
systems. A recent advance towards answering such questions has been made in the development of the theory of
non-thermal fixed points: universal nonequilibrium configurations showing scaling in space and (evolution) time,
characterised by a small number of fundamental properties. The theory of such fixed points transposes the
concepts of equilibrium and di↵usive near-equilibrium
renormalisation-group theory to the real-time evolution
of nonequilibrium systems. These developments provide,
for example, a framework within which to understand the
turbulent, coarsening, and relaxation dynamics following
Now: Strong cooling quench!
11
log nk
the creation of various kinds of defects and nonlinear patterns in a Kibble-Zurek quench.
The existence and significance of non-thermal scaling solutions in space and time was discussed by Berges
nQ ≈ (ρa 3)-1/2
and collaborators in the context of reheating after earlyinitial distribution after quench
universe inflation [79, 80] and then generalised by Berges,
Gasenzer, and coworkers to scenarios of strong matterwave turbulence [81, 152]. For the condensation dynamparticles removed
by cooling quench
ics of the dilute Bose gas discussed here, the presence of a
non-thermal fixed point can exert a significant influence
in the case of a strong cooling quench [30, 153, 154].
Q
As an illustration, we consider a particle distribukξ
log k
tion that
p drops abruptly above the healing-length scale
k⇠ = 8⇡a⇢, as depicted on a double-logarithmic scale
FIG. 4. Sketch of the evolution of the single-particle momenin Fig. 4 (dashed line). In order for the influence of the
tum distribution nk (t) of a Bose gas close to a non-thermal
non-thermal fixed point to be observed, the decay of n(k)
Synthetic · Zagreb
· 30 Sep
2015
fixed
point
(after Ref. [153]). Starting from the extreme initial
above Q ' k⇠ is assumed to be much steeper than the
distribution (dashed line, see main text for details) produced,
quasi-thermal scaling that develops in the kinetic stage
e.g., by a strong cooling quench, a bidirectional redistribution
of condensation following a weak quench [27, 28], as disof particles in momentum space (arrows) builds up a quasicussed in Sect. II. Such a distribution would, e.g., recondensate in the infrared while refilling the thermal tail at
sult from a severe cooling quench of a thermal Bose
large momenta. The particle transport towards zero momengas initially just above the critical temperature where
tum is characterised by a self-similar scaling evolution in space
and time, n(k, t) = (t/t0 )↵ n([t/t0 ] k, t0 ), with characteristic
T > |µ|/kB such that the Bose-Einstein distribution has
scaling exponents ↵, . Note the double-logarithmic scale.
developed a Rayleigh–Jeans scaling regime where n(k) ⇠
2mkB T /(~k)2 . The modulus of the chemical potential of
this state determines the momentum scale Q where the
phenomenological grounds in Ref. [21], and other types of
flat infrared scaling of the distribution goes over to the
(quasi-)topological excitations in low-dimensional, spinor
Rayleigh–Jeans scaling at larger k. If this chemical poand gauge systems [158–162].
tential is of the order of the ground-state energy of the
post-quench fully condensed gas, (~Q)2 /2m ' |µ| ' g⇢,
The dynamics in the vicinity of the fixed point are
with g = 4⇡~2 a/m, then the energy of the entire gas
characterised by an anomalously slow relaxation of the
is concentrated at the scale Q ' k⇠ after the quench.
total vortex line length, which exhibits an algebraic deThis is a key feature of the extreme nonequilibrium inicay ⇠ t 0.88 (see Ref. [158] for analogous results in the
tial state from which a non-thermal fixed point can be
2D case). At the same time, the condensate population
approached. We note that, if in this state there is no
grows as n0 (t) ⇠ t2 [30, 153], a significant slowing comsignificant zero-mode occupation n0 , the respective ocpared to the ⇠ t3 behaviour observed for weakly nonequicupation number at Q is on the order of the inverse of
librium condensate formation [30, 163].
the diluteness parameter, nQ ⇠ ⇣ 3/2 .
In the vicinity of the fixed point the momentum distribution is expected to follow a self-similar scaling beIn analogy to the weak-wave-turbulence scenario [19,
haviour in space and time in the infrared, n(k, t) =
20, 27, 28] discussed in Sect. II A, the initial overpop(t/t0 )↵ n([t/t0 ] k, t0 ). For a 3D Bose gas, these scaling
ulation of modes with energies ⇠ (~Q)2 /2m leads to
exponents have recently been numerically determined to
inverse particle transport while energy is transported
be ↵ = 1.66(12),
= 0.55(3), in agreement with the
to higher wavenumbers, as indicated by the arrows in
analytically predicted values ↵ = d,
= 1/2 [153].
Fig. 4 [30, 153, 154]. However, the inverse transport
This behaviour, here corresponding to the dilution and
involves non-local scattering and thus does not reprerelaxation of vortices leading to a build-up of the consent a cascade. Furthermore, in contrast to the case of
densate population, represents the generalisation of crita weak quench [14, 15, 18–20, 28, 29], in which weakical slowing-down to real-time evolution far away from
wave turbulence produces a quasi-thermal momentum
thermal equilibrium. At very late times, the system
distribution that relaxes quickly to a thermal equilibleaves the vicinity of the non-thermal fixed point, typrium distribution, here the inverse transport is characically when the last topological patterns decay, and fiterised by a strongly non-thermal power-law scaling in
nally approaches thermal equilibrium [32–35, 158, 164].
the infrared. Specifically, the momentum distribution
d 2
5
This equilibrium state corresponds to a fully established
n(k) ⇠ k
⇠ k in d = 3 dimensions [81] provides
condensate superimposed with weak sound excitations.
the “smoking-gun” of the influence of the non-thermal
fixed point. Semiclassical simulations by Nowak, GasenIn summary, non-thermal fixed points are nonequilibzer and collaborators [30, 155–157] showed that this scalrium field configurations, exhibiting universal scaling in
ing is associated with the creation, dilution, coarsening
time and space, to which the system is attracted if suitand relaxation of a complex vortex tangle, as predicted on
ably forced — e.g., in the case of Bose condensation, fol-
12
lowing a sufficiently strong cooling quench. In the vicinity of such fixed points, the relaxation of the field is critically slowed down and the dynamics exhibit self-similar
time evolution, governed by new critical exponents and
scaling functions. The possibility of categorising systems
into generalised “universality classes” associated with the
new critical exponents is a fascinating prospect and the
subject of current research [23, 24].
Finally in this section we note that prethermalisation
or “pre-Gibbsianisation”, i.e., the approach of a state
characterised by a Generalised Gibbs ensemble, usually
in near-integrable systems (see the chapter by Schmiedmayer), represents a special case of a Gaussian nonthermal fixed point, meaning that the e↵ective coupling
of the prethermalised modes vanishes. It is expected that
the exponents ↵ and in such a situation can become
very small compared to unity. Only at very late times
the remaining e↵ects of interactions may eventually drive
the system away from the fixed point towards a thermal
state.
V.
CONCLUSIONS AND OUTLOOK
In this chapter we have provided a brief introduction
to the scenario of the formation of a Bose–Einstein con-
[1] N. P. Proukakis, S. A. Gardiner, M. J. Davis, and
M. Szymańska, eds., Quantum Gases: Finite Temperature and Non-Equilibrium Dynamics (Imperial College
Press, London, UK, 2013).
[2] A. Griffin, T. Nikuni, and E. Zaremba, Bose-Condensed
Gases at Finite Temperatures (Cambridge University
Press, Cambridge, UK, 2009).
[3] V. N. Popov, Theor. Math. Phys. 11, 565 (1972).
[4] V. N. Popov, Functional Integrals in Quantum Field
Theory and Statistical Physics (Reidel, Dordrecht,
1983).
[5] A. Inoue and E. Hanamura, J. Phys. Soc. Jpn. 41, 771
(1976).
[6] E. Levich and V. Yakhot, Physical Review B 15, 243
(1977).
[7] E. Levich and V. Yakhot, J. Phys. A 11, 2237 (1978).
[8] E. Levich and V. Yakhot, J. Low. Temp. Phys. 27, 107
(1978).
[9] Y. B. Zeldovich and E. V. Levich, [Zh. Eksp. Teor. Fiz.
55, 2423 (1968)] Sov. Phys. JETP 28, 1287 (1969).
[10] L. P. Pitaevskii and S. Stringari, Bose–Einstein Condensation (Clarendon Press, Oxford, UK, 2003).
[11] S. G. Tikhodeev, [Zh. Eksp. Teor. Fiz. 97, 681 (1990)]
Sov. Phys. JETP 70, 380 (1990).
[12] U. Eckern, J. Low. Temp. Phys. 54, 333 (1984).
[13] D. W. Snoke and J. P. Wolfe, Phys. Rev. B 39, 4030
(1989).
[14] H. T. C. Stoof, Phys. Rev. Lett. 66, 3148 (1991).
[15] H. T. C. Stoof, Phys. Rev. A 45, 8398 (1992).
[16] H. T. C. Stoof, Bose-Einstein Condensation (Cambridge University Press, 1995), chap. Condensate For-
densate, and physics related to the dynamics of the BEC
phase transition. We have given a fairly comprehensive
review of the experiments studying the formation of simple, single-component BECs in three-dimensional atomic
gases, with brief mentions of how such features are affected by reduced e↵ective dimensionality or in cases
where more than one condensate may co-exist. However, the underlying physics described here is relevant to
several other systems, most notably exciton-polaritons
confined in strictly two-dimensional geometries featuring
pumping and decay, where experiments on condensate
formation have also been performed [165–167].
An interesting question is what are the similarities and
di↵erences between these systems, and others such as
BECs of photons [168] and magnons [169]. Furthermore,
what can phase transitions in quantum gases teach us
about phase transitions that cannot be accessed experimentally, such as inflationary scenarios of early-universe
evolution? This was one of the motivating questions in
the formulation of the Kibble–Zurek mechanism, as well
as in the development of the theory of non-thermal fixed
points. It remains to be seen what we can learn about
such matters as the formation of cosmological topological defects and (possibly) baryon asymmetry by studying
nanokelvin gases here on earth.
[17]
[18]
[19]
[20]
[21]
[22]
[23]
[24]
[25]
[26]
[27]
[28]
[29]
[30]
mation in a Bose Gas, p. 226.
H. T. C. Stoof, Phys. Rev. Lett. 78, 768 (1997).
H. T. C. Stoof, J. Low Temp. Phys. 114, 11 (1999).
B. V. Svistunov, J. Mosc. Phys. Soc. 1, 373 (1991).
Y. Kagan, B. V. Svistunov, and G. V. Shlyapnikov,
Zh. Éksp. Teor. Fiz. 101, 528 (1992), [Sov. Phys. JETP
75, 387 (1992)].
Y. Kagan and B. V. Svistunov, Zh. Éksp. Teor. Fiz.
105, 353 (1994), [Sov. Phys. JETP 78, 187 (1994)].
Y. Kagan, Bose-Einstein Condensation (Cambridge
University Press, 1995), chap. Kinetics of Bose-Einstein
Condensate Formation in an Interacting Bose Gas, p.
202.
L. M. Sieberer, S. D. Huber, E. Altman, and S. Diehl,
Phys. Rev. Lett. 110, 195301 (2013).
E. Altman, L. M. Sieberer, L. Chen, S. Diehl, and
J. Toner, Phys. Rev. X 5, 011017 (2015).
G. Dagvadorj, J. M. Fellows, S. Matyjaskiewicz, F. M.
Marchetti, I. Carusotto, and M. H. Szymanska, Phys.
Rev. X 5, 041028 (2015).
V. E. Zakharov, V. S. L’vov, and G. Falkovich, Kolmogorov Spectra of Turbulence I: Wave Turbulence
(Springer, Berlin, 1992).
D. V. Semikoz and I. I. Tkachev, Phys. Rev. Lett. 74,
3093 (1995).
D. V. Semikoz and I. I. Tkachev, Phys. Rev. D 55, 489
(1997).
N. G. Berlo↵ and B. V. Svistunov, Phys. Rev. A 66,
013603 (2002).
B. Nowak, J. Schole, and T. Gasenzer, New J. Phys. 16,
093052 (2014).
13
[31] K. W. Schwarz, Phys. Rev. B 38, 2398 (1988).
[32] E. Kozik and B. Svistunov, Phys. Rev. Lett. 92, 035301
(2004).
[33] E. Kozik and B. Svistunov, Phys. Rev. Lett. 94, 025301
(2005).
[34] E. Kozik and B. Svistunov, Phys. Rev. B 72, 172505
(2005).
[35] E. V. Kozik and B. V. Svistunov, J. Low Temp. Phys.
156, 215 (2009).
[36] M. H. Anderson, J. R. Ensher, M. R. Matthews, C. E.
Wieman, and E. A. Cornell, Science 269, 198 (1995).
[37] K. B. Davis, M. O. Mewes, M. R. Andrews, N. J. van
Druten, D. S. Durfee, D. M. Kurn, and W. Ketterle,
Phys. Rev. Lett. 75, 3969 (1995).
[38] W. Ketterle and N. J. Van Druten (Academic Press,
1996), vol. 37 of Advances In Atomic, Molecular, and
Optical Physics, p. 181.
[39] C. W. Gardiner and P. Zoller, Phys. Rev. A 55, 2902
(1997).
[40] C. W. Gardiner and P. Zoller, Phys. Rev. A 58, 536
(1998).
[41] C. W. Gardiner and P. Zoller, Phys. Rev. A 61, 033601
(2000).
[42] C. W. Gardiner and P. Zoller, Quantum Noise
(Springer-Verlag, Berlin Heidelberg, 2004), 3rd ed.
[43] H.-J. Miesner, D. M. Stamper-Kurn, M. R. Andrews,
D. S. Durfee, S. Inouye, and W. Ketterle, Science 279,
1005 (1998).
[44] C. W. Gardiner, P. Zoller, R. J. Ballagh, and M. J.
Davis, Phys. Rev. Lett. 79, 1793 (1997).
[45] C. W. Gardiner, M. D. Lee, R. J. Ballagh, M. J. Davis,
and P. Zoller, Phys. Rev. Lett. 81, 5266 (1998).
[46] M. D. Lee and C. W. Gardiner, Phys. Rev. A 62, 033606
(2000).
[47] M. J. Davis, C. W. Gardiner, and R. J. Ballagh, Phys.
Rev. A 62, 063608 (2000).
[48] M. J. Bijlsma, E. Zaremba, and H. T. C. Stoof, Phys.
Rev. A 62, 063609 (2000).
[49] E. Zaremba, T. Nikuni, and A. Griffin, J. Low. Temp.
Phys. 116, 277 (1999).
[50] E. Zaremba, A. Griffin, and T. Nikuni, Phys. Rev. A
57, 4695 (1998).
[51] T. R. Kirkpatrick and J. R. Dorfman, Phys. Rev. A 28,
2576 (1983).
[52] T. R. Kirkpatrick and J. R. Dorfman, J. Low Temp.
Phys. 58, 399 (1985).
[53] T. R. Kirkpatrick and J. R. Dorfman, J. Low Temp.
Phys. 58, 301 (1985).
[54] T. R. Kirkpatrick and J. R. Dorfman, J. Low Temp.
Phys. 59, 1 (1985).
[55] O. J. Luiten, M. W. Reynolds, and J. T. M. Walraven,
Phys. Rev. A 53, 381 (1996).
[56] M. Köhl, M. J. Davis, C. W. Gardiner, T. W. Hänsch,
and T. W. Esslinger, Phys. Rev. Lett. 88, 080402
(2002).
[57] M. J. Davis and C. W. Gardiner, J. Phys. B: At. Mol.
Opt. Phys. 35, 733 (2002).
[58] P. W. H. Pinske, A. Mosk, M. Weidemüller, M. W.
Reynolds, T. W. Hijmans, and J. T. M. Walraven, Phys.
Rev. Lett. 78, 990 (1997).
[59] D. M. Stamper-Kurn, H.-J. Miesner, A. P. Chikkatur,
S. Inouye, J. Stenger, and W. Ketterle, Phys. Rev. Lett.
81, 2194 (1998).
[60] H. T. C. Stoof and M. J. Bijlsma, J. Low. Temp. Phys.
124, 431 (2001).
[61] N. P. Proukakis, J. Schmiedmayer, and H. T. C. Stoof,
Phys. Rev. A 73, 053603 (2006).
[62] M. C. Garrett, A. Ratnapala, E. D. van Ooijen, C. J.
Vale, K. Weegink, S. K. Schnelle, O. Vainio, N. R. Heckenberg, H. Rubinsztein-Dunlop, and M. J. Davis, Phys.
Rev. A 83, 013630 (2011).
[63] D. M. Harber, J. M. McGuirk, J. M. Obrecht, and E. A.
Cornell, J. Low Temp. Phys. 133, 229 (2003).
[64] A. L. Marchant, S. H andel, T. P. Wiles, S. A. Hopkins,
and S. L. Cornish, New J. Phys. 13, 125003 (2011).
[65] J. Märkle, A. J. Allen, P. Federsel, B. Jetter,
A. Günther, J. Fortágh, N. P. Proukakis, and T. E.
Judd, Phys. Rev. A 90, 023614 (2014).
[66] M. Imamovic-Tomasovic and A. Griffin, J. Low Temp.
Phys. 122, 616 (2001).
[67] L. P. Kadano↵ and G. Baym, Quantum Statistical Mechanics (W.A. Benjamin, Menlo Park, CA, USA, 1962).
[68] R. Walser, J. Williams, J. Cooper, and M. Holland,
Phys. Rev. A 59, 3878 (1999).
[69] R. Walser, J. Cooper, and M. Holland, Phys. Rev. A
63, 013607 (2001).
[70] J. Wachter, R. Walser, J. Cooper, and M. Holland,
Phys. Rev. A 64, 053612 (2001).
[71] N. P. Proukakis, J. Phys. B: At. Mol. Opt. Phys. 34,
4737 (2001).
[72] N. P. Proukakis and K. Burnett, J. Res. Natl. Inst.
Stand. Technol. 101, 457 (1996).
[73] N. P. Proukakis, K. Burnett, and H. T. C. Stoof, Phys.
Rev. A 57, 1230 (1998).
[74] H. Shi and A. Griffin, Physics Reports 304, 1 (1998),
ISSN 0370-1573.
[75] T. Gasenzer, J. Berges, M. G. Schmidt, and M. Seco,
Phys. Rev. A 72, 063604 (2005).
[76] J. Berges and T. Gasenzer, Phys. Rev. A 76, 033604
(2007).
[77] A. Branschädel and T. Gasenzer, J. Phys. B: At. Mol.
Opt. Phys. 41, 135302 (2008).
[78] C. Bodet, M. Kronenwett, B. Nowak, D. Sexty, and
T. Gasenzer, Quantum Gases: Finite Temperature and
Non-Equilibrium Dynamics (Imperial College Press,
London, 2012), chap. Non-equilibrium Quantum ManyBody Dynamics: Functional Integral Approaches.
[79] J. Berges, A. Rothkopf, and J. Schmidt, Phys. Rev.
Lett. 101, 041603 (2008).
[80] J. Berges and G. Ho↵meister, Nucl. Phys. B813, 383
(2009).
[81] C. Scheppach, J. Berges, and T. Gasenzer, Phys. Rev.
A 81, 033611 (2010).
[82] M. Kronenwett and T. Gasenzer, Appl. Phys. B 102,
469 (2011).
[83] M. Babadi, E. Demler, and M. Knap, Phys. Rev. X 5,
041005 (2015).
[84] C. J. Myatt, E. A. Burt, R. W. Ghrist, E. A. Cornell,
and C. E. Wieman, Phys. Rev. Lett. 78, 586 (1997).
[85] F. Schreck, G. Ferrari, K. L. Corwin, J. Cubizolles,
L. Khaykovich, M.-O. Mewes, and C. Salomon, Phys.
Rev. A 64, 011402 (2001).
[86] J. Catani, G. Barontini, G. Lamporesi, F. Rabatti,
G. Thalhammer, F. Minardi, S. Stringari, and M. Inguscio, Phys. Rev. Lett. 103, 140401 (2009).
[87] M. Erhard, H. Schmaljohann, J. Kronjäger, K. Bongs,
and K. Sengstock, Phys. Rev. A 70, 031602(R) (2004).
[88] Y. Shin, M. Saba, A. Schirotzek, T. A. Pasquini, A. E.
14
[89]
[90]
[91]
[92]
[93]
[94]
[95]
[96]
[97]
[98]
[99]
[100]
[101]
[102]
[103]
[104]
[105]
[106]
[107]
[108]
[109]
[110]
[111]
[112]
[113]
Leanhardt, D. E. Pritchard, and W. Ketterle, Phys.
Rev. Lett. 92, 150401 (2004).
S. Stellmer, B. Pasquiou, R. Grimm, and F. Schreck,
Phys. Rev. Lett. 110, 263003 (2013).
J. M. Kosterlitz and D. J. Thouless, J. Phys. C. 6, 1181
(1973).
Z. Hadzibabic, P. Kruger, M. Cheneau, B. Battelier,
and J. Dalibard, Nature 441, 1118 (2006).
V. Schweikhard, S. Tung, and E. A. Cornell, Phys. Rev.
Lett. 99, 030401 (2007).
P. Krüger, Z. Hadzibabic, and J. Dalibard, Phys. Rev.
Lett. 99, 040402 (2007).
P. Cladé, C. Ryu, A. Ramanathan, K. Helmerson, and
W. D. Phillips, Phys. Rev. Lett. 102, 170401 (2009).
S. Tung, G. Lamporesi, D. Lobser, L. Xia, and E. A.
Cornell, Phys. Rev. Lett. 105, 230408 (2010).
C.-L. Hung, X. Zhang, N. Gemekle, and C. Chin, Nature
470, 236 (2011).
N. Prokof’ev, O. Ruebenacker, and B. Svistunov, Phys.
Rev. Lett. 87, 270402 (2001).
T. P. Simula and P. B. Blakie, Phys. Rev. Lett. 96,
020404 (2006).
M. Holzmann and W. Krauth, Phys. Rev. Lett. 100,
190402 (2008).
R. N. Bisset, M. J. Davis, T. P. Simula, and P. B. Blakie,
Phys. Rev. A 79, 033626 (2009).
S. P. Cockburn and N. P. Proukakis, Phys. Rev. A 86,
033610 (2012).
Z. Hadzibabic and J. Dalibard, Nano optics and atomics: transport of light and matter waves (Rivista del
Nuovo Cimento, 2011), vol. 34 of Proceedings of the
International School of Physics ”Enrico Fermi”, Vol.
CLXXIII, chap. Two Dimensional Bose Fluids: An
Atomic Physics Perspective, p. 389.
G. Roumpos, M. Lohse, W. H. Nitsche, J. Keeling, M. H. Szymańska, P. B. Littlewood, A. Lö✏er,
S. Höfling, L. Worschech, A. Forchel, et al., Proceedings
of the National Academy of Sciences 109, 6467 (2012).
W. H. Nitsche, N. Y. Kim, G. Roumpos, C. Schneider,
M. Kamp, S. Höfling, A. Forchel, and Y. Yamamoto,
Phys. Rev. B 90, 205430 (2014).
D. S. Petrov, G. V. Shlyapnikov, and J. T. M. Walraven,
Phys. Rev. Lett. 87, 050404 (2001).
I. Shvarchuck, C. Buggle, D. S. Petrov, K. Dieckmann,
M. Zielonkowski, M. Kemmann, T. G. Tiecke, W. von
Klitzing, G. V. Shlyapnikov, and J. T. M. Walraven,
Phys. Rev. Lett. 89, 270404 (2002).
M. Hugbart, J. A. Retter, A. F. Varón, P. Bouyer,
A. Aspect, and M. J. Davis, Phys. Rev. A 75, 011602
(2007).
J. J. Binney, N. J. Dowrick, A. J. Fisher, and M. E. J.
Newman, The Theory of Critical Phenomena: An Introduction to the Renormalization Group (Oxford University Press, USA, 1992).
J. Zinn-Justin, Quantum Field Theory and Critical Phenomena (Clarendon Press, Oxford, 2002), 4th ed.
P. C. Hohenberg and B. I. Halperin, Rev. Mod. Phys.
49, 435 (1977).
G. Baym, J.-P. Blaizot, M. Holzmann, F. Laloë, and
D. Vautherin, Phys. Rev. Lett. 83, 1703 (1999).
V. A. Kashurnikov, N. V. Prokof’ev, and B. V. Svistunov, Phys. Rev. Lett. 87, 120402 (2001).
P. Arnold and G. Moore, Phys. Rev. Lett. 87, 120401
(2001).
[114] F. Gerbier, J. H. Thywissen, S. Richard, M. Hugbart,
P. Bouyer, and A. Aspect, Phys. Rev. Lett. 92, 030405
(2004).
[115] M. J. Davis and P. B. Blakie, Phys. Rev. Lett. 96,
060404 (2006).
[116] R. P. Smith, R. L. D. Campbell, N. Tammuz, and
Z. Hadzibabic, Phys. Rev. Lett. 106, 250403 (2011).
[117] S. Ritter, A. Öttl, T. Donner, T. Bourdel, M. Köhl, and
T. Esslinger, Phys. Rev. Lett. 98, 090402 (2007).
[118] T. Donner, S. Ritter, T. Bourdel, A. Öttl, M. Köhl, and
T. Esslinger, Science 315, 1556 (2007).
[119] A. Bezett and P. B. Blakie, Phys. Rev. A 79, 033611
(2009).
[120] N. Navon, A. L. Gaunt, R. P. Smith, and Z. Hadzibabic,
Science 347, 167 (2015).
[121] T. W. B. Kibble, J. Phys. A: Math. Gen. 9, 1387 (1976).
[122] L. D. Landau and E. M. Lifshitz, Statistical Physics,
Part 1 (Butterworth–Heinemann, Oxford, UK, 1980),
3rd ed.
[123] W. H. Zurek, Nature 317, 505 (1985).
[124] W. H. Zurek, Physics Reports 276, 177 (1996).
[125] J. R. Anglin and W. H. Zurek, Phys. Rev. Lett 83, 1707
(1999).
[126] C. N. Weiler, T. W. Neely, D. R. Scherer, A. S. Bradley,
M. J. Davis, and B. P. Anderson, Nature 455, 948
(2008).
[127] D. V. Freilich, D. M. Bianchi, A. M. Kaufman, T. K.
Langin, and D. S. Hall, Science 329, 1182 (2010).
[128] C. W. Gardiner and M. J. Davis, J. Phys. B: At. Mol.
Opt. Phys. 36, 4731 (2003).
[129] P. B. Blakie, A. S. Bradley, M. J. Davis, R. J. Ballagh, and C. W. Gardiner, Advances in Physics 57, 363
(2008).
[130] N. P. Proukakis and B. Jackson, J. Phys. B: At. Mol.
Opt. 41, 203002 (2008).
[131] S. P. Cockburn and N. P. Proukakis, Laser Phys. 19,
558 (2009).
[132] M. J. Steel, M. K. Olsen, L. I. Plimak, P. D. Drummond,
S. M. Tan, M. J. Collett, D. F. Walls, and R. Graham,
Phys. Rev. A 58, 4824 (1998).
[133] P. D. Drummond and J. F. Corney, Phys. Rev. A. 60,
R2661 (1999).
[134] J. Chang, C. Hamner, and P. Engels (2009), 40th Annual Meeting of the APS Division of Atomic, Molecular
and Optical Physics.
[135] W. H. Zurek, Phys. Rev. Lett. 102, 105702 (2009).
[136] B. Damski and W. H. Zurek, Phys. Rev. Lett. 104,
160404 (2010).
[137] E. Witkowska, P. Deuar, M. Gajda, and K. Rza̧żewski,
Phys. Rev. Lett. 106, 135301 (2011).
[138] A. del Campo, A. Retzker, and M. B. Plenio, New J.
Phys. 13, 083022 (2011).
[139] G. Lamporesi, S. Donadello, S. Serafini, F. Dalfovo, and
G. Ferrari, Nat. Phys. 9, 656 (2013).
[140] S. Donadello, S. Serafini, M. Tylutki, L. P. Pitaevskii,
F. Dalfovo, G. Lamporesi, and G. Ferrari, Phys. Rev.
Lett. 113, 065302 (2014).
[141] L. Chomaz, L. Corman, T. Bienaimé, R. Desbuquois,
C. Weitenberg, S. N. amd Jérôme Beugnon, and J. Dalibard, Nat. Comm. 6, 6162 (2015).
[142] L. Corman, L. Chomaz, T. Bienaimé, R. Desbuquois,
C. Weitenberg, S. Nascimbène, J. Dalibard, and
J. Beugnon, Phys. Rev. Lett. 113, 135302 (2014).
15
[143] A. Das, J. Sabbatini, and W. H. Zurek, Sci. Rep. 2, 352
(2012).
[144] L. E. Sadler, J. M. Higbie, S. R. Leslie, M. Vengalattore,
and D. M. Stamper-Kurn, Nature 443, 312 (2006).
[145] S. De, D. L. Campbell, R. M. Price, A. Putra, B. M.
Anderson, and I. B. Spielman, Phys. Rev. A 89, 033631
(2014).
[146] S. B. Papp, J. M. Pino, and C. E. Wieman, Phys. Rev.
Lett. 101, 040402 (2008).
[147] D. J. McCarron, H. W. Cho, D. L. Jenkin, M. P.
Köppinger, and S. L. Cornish, Phys. Rev. A 84, 011603
(2011).
[148] I.-K. Liu, R. W. Pattinson, T. P. Billam, S. A. Gardiner,
S. L. Cornish, H. T.-M., W.-W. Lin, S.-C. Gou, N. G.
Parker, and N. P. Proukakis, arXiv:1408.0891v3 (2015).
[149] J. Sabbatini, W. H. Zurek, and M. J. Davis, Phys. Rev.
Lett. 107, 230402 (2011).
[150] T. Swislocki, E. Witkowska, J. Dziarmaga, and M. Matuszewski, Phys. Rev. Lett. 110, 045303 (2013).
[151] J. Hofmann, S. S. Natu, and S. Das Sarma, Phys. Rev.
Lett. 113, 095702 (2014).
[152] S. Mathey, T. Gasenzer, and J. M. Pawlowski, Phys.
Rev. A 92, 023635 (2015).
[153] A. Piñeiro Orioli, K. Boguslavski, and J. Berges, Phys.
Rev. D 92, 025041 (2015).
[154] J. Berges and D. Sexty, Phys. Rev. Lett. 108, 161601
(2012).
[155] B. Nowak, D. Sexty, and T. Gasenzer, Phys. Rev. B 84,
020506(R) (2011).
[156] B. Nowak, J. Schole, D. Sexty, and T. Gasenzer, Phys.
Rev. A 85, 043627 (2012).
[157] B. Nowak, S. Erne, M. Karl, J. Schole, D. Sexty,
and T. Gasenzer, in Proc. Int. School on Strongly In-
[158]
[159]
[160]
[161]
[162]
[163]
[164]
[165]
[166]
[167]
[168]
[169]
teracting Quantum Systems Out of Equilibrium, Les
Houches, 2012 (to appear) (2013), arXiv:1302.1448
[cond-mat.quant-gas].
J. Schole, B. Nowak, and T. Gasenzer, Phys. Rev. A 86,
013624 (2012).
M. Schmidt, S. Erne, B. Nowak, D. Sexty, and T. Gasenzer, New J. Phys. 14, 075005 (2012).
M. Karl, B. Nowak, and T. Gasenzer, Sci. Rept. 3, 2394
(2013).
M. Karl, B. Nowak, and T. Gasenzer, Phys. Rev. A 88,
063615 (2013).
T. Gasenzer, L. McLerran, J. M. Pawlowski, and
D. Sexty, Nucl. Phys. A 930, 163 (2014).
K. Damle, S. N. Majumdar, and S. Sachdev, Phys. Rev.
A 54, 5037 (1996).
C. Connaughton, C. Josserand, A. Picozzi, Y. Pomeau,
and S. Rica, Phys. Rev. Lett. 95, 263901 (2005).
G. Nardin, K. G. Lagoudakis, M. Wouters, M. Richard,
A. Baas, R. André, L. S. Dang, B. Pietka, and
B. Deveaud-Plédran, Phys. Rev. Lett. 103, 256402
(2009).
V. V. Belykh, N. N. Sibeldin, V. D. Kulakovskii,
M. M. Glazov, M. A. Semina, C. Schneider, S. Höfling,
M. Kamp, and A. Forchel, Phys. Rev. Lett. 110, 137402
(2013).
K. G. Lagoudakis, F. Manni, B. Pietka, M. Wouters,
T. C. H. Liew, V. Savona, A. V. Kavokin, R. André,
and B. Deveaud-Plédran, Phys. Rev. Lett. 106, 115301
(2011).
J. Klaers, J. Schmitt, F. Vewinger, and M. Weitz, Nature 468, 545 (2010).
S. O. Demokritov, O. D. V. E. Demidov an, G. A.
Melkov, A. A. Serga, B. Hillebrands, and A. N. Slavin,
Nature 443, 430 (2006).