Control of anomalous diffusion of a Bose polaron
Christos Charalambous
1
, Miguel Ángel García-March
1,2
, Gorka Muñoz-Gil
1
, Przemysław Ryszard
Grzybowski
3
, and Maciej Lewenstein
1,4
1
ICFO Institut de Ciéncies Fotóniques, The Barcelona Institute of Science and Technology, 08860 Castelldefels (Barcelona), Spain
2
Instituto Universitario de Matemática Pura y Aplicada, Universitat Politècnica de València, E-46022 València, Spain
3
Faculty of Physics, Adam Mickiewicz University, Umultowska 85, 61-614 Poznań, Poland
4
ICREA, Lluis Companys 23, E-08010 Barcelona, Spain
February 18, 2020
We study the diffusive behavior of a Bose
polaron immersed in a coherently coupled two-
component Bose-Einstein Condensate (BEC).
We assume a uniform, one-dimensional BEC.
Polaron superdiffuses if it couples in the same
manner to both components, i.e. either at-
tractively or repulsively to both of them. This
is the same behavior as that of an impurity
immersed in a single BEC. Conversely, the po-
laron exhibits a transient nontrivial subdiffu-
sive behavior if it couples attractively to one of
the components and repulsively to the other.
The anomalous diffusion exponent and the du-
ration of the subdiffusive interval can be con-
trolled with the Rabi frequency of the coher-
ent coupling between the two components, and
with the coupling strength of the impurity to
the BEC.
1 Introduction
The phenomenon of anomalous diffusion attracts a
growing interest in classical and quantum physics, ap-
pearing in a plethora of various systems [1, 2]. In
classical systems, there has been a considerable effort
to elucidate the properties and conditions of anoma-
lous diffusive behavior, with a large emphasis given
to the question of how this anomalous diffusion could
potentially be controlled. In many models, the ap-
pearance of the anomalous diffusion is attributed to
some random component of the system-environment
setup, usually distributed with a power-law. Exam-
ples include continuous time random walks [3], dif-
fusion on a fractal lattice [4], diffusivity (i.e. diffu-
sion coefficient) that is inhomogeneous in time [5, 6],
or space [711] in a regular or random manner, the
patch model [12, 13], hunters model [14], etc. In quan-
tum systems, a paradigmatic instance of a highly con-
trolled system is that of a Bose Einstein Condensate
(BEC). It was shown that BEC with tunable inter-
actions, are promising systems to study a number of
diffusion-related phenomena, such as Anderson Local-
ization (AL) in disordered media [1517], the expan-
sion of 1D BEC in disordered speckle potentials [15
23], the subdiffusive behavior of the expansion of a
wave packet of a 1D quantum, chaotic and nonlinear
system [15, 22, 2433], the Brownian motion of soli-
tons in BEC [34], as well as the superdiffusive motion
of an impurity in a BEC studied in [3537].
In this work, we study how an impurity in a coher-
ently coupled two-component BEC shows a transient
anomalous diffusing behavior. We study this phe-
nomenon under experimentally relevant conditions, as
long as the BEC can be approximated as uniform and
one dimensional. We show that this transient anoma-
lous diffusing behavior can be controlled through the
strength of the interactions and the coherent coupling.
To this end, we treat the Bose Polaron problem within
an open quantum system framework. The open quan-
tum system approach has been used recently in the
context of ultracold quantum gases to study the diffu-
sion of an impurity and two impurities in a BEC [35
37], for the movement of a bright soliton in a super-
fluid in one dimension [38], see also[3941]). On the
other hand, the effect of contact interactions, dipole-
dipole interactions and disorder on the diffusion prop-
erties of 1D dipolar two-component condensates were
studied in [42], identifying again the conditions for
subdiffusion. The study of the diffusive behavior of
a 2D two-component BEC in a disordered potential
was undertaken in [43]. Finally, an important study
on an impurity immersed in a two-component BEC
was reported in [44].
The most important novel result of this work, is
that we show that under certain assumptions, one
can observe a transient subdiffusive behavior of the
immersed impurity. We study this for experimentally
feasible parameters, as long as the BEC can be ap-
proximated as uniform and one dimensional, and we
examine how the strength of the coherent coupling
and interactions modify this subdiffusive behavior.
To be more specific about the particularities of the
system considered in this work, we assume that an
external field drives the population transfer (spin-
flipping) between the two atomic levels. The popu-
lation transfer between the two levels turns out to be
described by Josephson dynamics, leading to what is
Accepted in Quantum 2019-12-28, click title to verify 1
arXiv:1910.01571v4 [quant-ph] 17 Feb 2020
known as internal Josephson effect (see e.g. [45]). This
internal Josephson interaction controls the many-
body physics of multicomponent phase coherent mat-
ter. Importantly, after diagonalizing the Hamiltonian
through a Bogoliubov transformation, one obtains a
spectrum that has two branches: the density mode,
ungapped and with a linear behavior at low momenta;
and the spin mode, gapped, with a parabolic behavior
even at low momenta.
From the technical point of view, we identify how
under suitable assumptions, starting from the Hamil-
tonian describing the aforementioned system of an im-
purity in a coherently coupled two component BEC,
one can equivalently describe the impurity as a Brow-
nian particle in a bath, where the role of the bath
is played by the Bogoliubov modes of the coherently
coupled two-component BEC. Furthermore, we show
that the two branches obtained after the Bogoliubov
transformation, the density mode and the spin mode,
mentioned above, result in two distinct spectral den-
sities, which we derive in Section 3. We consider two
scenarios: same coupling among the impurity and the
two bosonic components, and repulsive coupling to
one component and attractive to the other. We show
that these scenarios correspond to the impurity cou-
pling either to the density or to the spin mode of the
two-component BEC, respectively. For the coupling
to the density mode there is no qualitative difference
in comparison to the case where the impurity is em-
bedded in a single BEC [35]. For the coupling to the
spin mode, we find a different spectral density, namely
a gapped sub-ohmic spectral density. We derive and
solve the equations of motion of the impurity. These
are obtained through the corresponding Heisenberg
equations for the bath and impurity particles and they
have the form of Generalized Langevin equations with
memory effects. By solving numerically these equa-
tions we find the effect of the gapped sub-Ohmic spec-
tral density on the Mean Square Displacement (MSD)
of the impurity.
The paper is organized as follows. In Section 2
we introduce the model Hamiltonian and transform
it into the form of a Caldeira-Leggett one. In Sec-
tion 3 we derive the spectral densities for the cases
of coupling to the density or spin modes. In Section 4
we find and solve the Langevin equations and in Sec-
tion 5 we present the results. We end the paper with
the discussion and outlook presented in Section 6.
2 Hamiltonian
The dilute Bose-Einstein condensates created in
atomic gases [46] consist of bosons with internal de-
grees of freedom: the atoms can be trapped in differ-
ent atomic hyperfine states. Soon after the first obser-
vation of atom trap BECs, experimentalists succeeded
in trapping partly overlapping BECs of atoms in dif-
ferent hyperfine states that are (i) hyperfine split [47]
or (ii) nearly degenerate and correspond to different
orientations of the spin [48]. We consider a two-
component Bose gas with both one-body (field-field)
and two-body (density-density) couplings, composed
of such atoms in different hyperfine states. Further-
more, we assume that the two components are cou-
pled through a Josephson (one-body) type of cou-
pling. The two-body interaction results from short-
range particle-particle interactions between atoms in
different internal states, while the one-body interac-
tion can be implemented by two-photon Raman opti-
cal coupling, which transfers atoms from one internal
state to the other. In present-day BEC experiments,
the internal Josephson or Rabi interactions, intercon-
verting atoms of different internal states, consist in
two-photon transitions, induced by a laser field or a
combination of a laser field and oscillating magnetic
field. To gain a perspective on the experimental rele-
vance of our study, we refer the readers to the work of
Refs. [47, 4951]. Finally, we assume an impurity that
is immersed in the two-component BEC. This impu-
rity interacts with both components through contact
interactions. In Fig. 1 a sketch of the setup is shown.
The Hamiltonian of an impurity interacting with a
two-species bosonic mixture in one dimension reads
H = H
I
+ H
(1)
B
+ H
(2)
B
+ H
IB
+ H
(12)
B
, (1)
where the impurity of mass m
I
is described by H
I
=
p
2
2m
I
+ U(x), with U(x) being the trapping potential.
The interactions with the bosons are described by
H
IB
. We study here only the case of free impurities,
hence we assume U(x) = 0. The terms of the individ-
ual bosonic species, labeled with the index j = 1, 2,
are
H
(j)
B
=
Z
Ψ
j
(x)
"
p
2
j
2m
j
B
+ V
j
(x)
#
Ψ
j
(x) dx
+
g
j
2
Z
Ψ
j
(x) Ψ
j
(x) Ψ
j
(x) Ψ
j
(x) dx,
where the intra-species contact interactions have
a strength given by the coupling constant g
j
=
4π~
2
a
(j)
B
/m
j
B
, with a
(j)
B
the scattering length for the
j
th
species atoms, m
j
B
the mass of the atoms of these
species and the external potential for the atoms of the
j
th
species is denoted by V
j
(x). For simplicity, we as-
sume V
1
(x) = V
2
(x) = V (x) and m
1
B
= m
2
B
= m
B
.
Furthermore we focus on the idealized case of an un-
trapped bath i.e. V (x) = 0 which results in a homo-
geneous density for the BEC. This is not a physical
scenario, but in practice, for a BEC trapped in a large
box, the impurity in the middle of this box would in-
deed approximately interact with a bath of constant
density.
The coupling Hamiltonian between the two bosonic
species consists in inter-species contact interactions,
with coupling constant g
12
, and a Rabi coupling ,
Accepted in Quantum 2019-12-28, click title to verify 2
which exchanges atoms between components, i.e.,
H
(12)
B
= g
12
Z
Ψ
1
(x) Ψ
2
(x) Ψ
2
(x) Ψ
1
(x) dx
+ ~
Z
Ψ
1
(x) Ψ
2
(x) dx + H.c. (2)
where g
12
= 4π~
2
a
(12)
B
/m
B
, with a
(12)
B
the scatter-
ing length for the intraspecies interactions. Without
any loss of generality, we will only consider real
and positive. This is because even if a complex Rabi
frequency is assumed, this can always be cancelled
by introducing a counteracting phase for one of the
BECs, which can be shown to have no effect on the
energy spectrum of the bath. The latter part of the
Hamiltonian, referred to as an internal Josephson in-
teraction, is a two-photon transition that is induced
by a laser field or a combination of a laser field and
an oscillating (rf) magnetic field. This also introduces
an effective energy difference between the two inter-
nal states/species of the BEC, which, assuming a low
intensity driving field, is simply equal to the detuning
δ of the two-photon transition. This detuning does
not affect our studies however, so for sake of clarity
and simplicity, we will assume it to be zero. We also
consider here only repulsive two-body coupling, i.e.
g
12
> 0.
Figure 1: We consider a setup of a coherently coupled two-
component BEC in which an impurity is immersed. By g
1
and g
2
we denote the intraspecies contact interactions of the
atoms of the first and second species respectively. g
12
refers
to the coupling strength of the interspecies contact interac-
tion among atoms of the first and second species. By we
denote the Rabi frequency of the Raman coherent coupling
between the two species. Finally g
(1)
IB
and g
(2)
IB
indicate the
coupling of the impurity to the atoms of the first and second
species respectively.
In what concerns the impurity-bosons interaction
part of the Hamiltonian, we assume that the interac-
tion is between the impurity and the densities of the
bosons, i.e. it has the form of a contact interaction:
H
IB
=
X
j=1,2
g
(j)
IB
Ψ
j
(x
j
(x) + h.c.
. (3)
with g
(j)
IB
= 2π~
2
a
(j)
IB
/m
R
, where a
(j)
IB
represents the
scattering length of the impurity with the bosons of
the j
th
BEC, and m
R
= m
B
m
I
/(m
B
+ m
I
) is the
relevant reduced mass. From this point onwards, we
assume that the BEC is one dimensional, which sim-
plifies the analytical part of our studies. Nevertheless,
the main result of our work, which is the control of the
dynamics of an impurity in a coherently coupled two-
component BEC by coupling it either to the density
mode of the BEC or to the spin mode of the BEC, will
remain irrespective of the dimension. The diffusing
behavior itself however might change, since the spec-
tral density depends on the dimension of the BEC.
It is worth noting here, that the 1D case in fact is
peculiar since, in principle, in the purely 1D scenario
(not in the confined 3D elongated/cigar shaped case)
the condensation is destroyed by the phase fluctua-
tions [52]. However, if the phase coherence length is
larger than the band-size, then one can speak about
"true" BEC. In other words, as long as the physics
of interest happens on the scales smaller than the
phase coherence length, it is legitimate to use BEC
and Bogolubov-de Gennes theory, as we will do in our
work.
Furthermore, we comment here that, had we con-
sidered the more realistic case of a harmonically
trapped BEC, assuming it to be described by a
Thomas-Fermi density profile, this would lead to a
discretized Bogoliubov spectrum, which would then
have to be treated under a similar approximation
as in the single BEC case in [36]. In principle, we
expect that the maim result of this work, i.e. be-
ing able to couple the impurity to two distinct types
of baths, should hold, but the related diffusing be-
haviours might change. We expect that when the gap
of the spin models is much bigger than the trap fre-
quency, the sub-diffusion, that we observe in Sec. 5,
should persist in the same form. The reader should
also have in mind that, as discussed in the appendix
of [36] the limit of the trapping frequency of the BEC
going to 0 would not lead to the homogeneous BEC
case. This scenario requires a more careful treatment,
that goes beyond the scope of our paper.
A further assumption we make in our studies is that
of a dilute gas of low depletion, since in this case, we
will be able to apply the Bogoliubov diagonalization
technique and obtain the energy spectrum of the BEC
bath. In the low-density sub-milikelvin temperature
regime of the atom trap experiments, we may assume
that the trapped atoms interact only through the par-
tial s-wave channel, and that the many-body proper-
ties are well described by assuming the particles to
interact as hard spheres. The radius of those spheres
is given by the scattering length a, which we assume
to be positive. We say that the system of particle
density n is dilute if the packing fraction of space oc-
cupied by the spheres na
3
1. The assumption of
low depletion means that almost all particles occupy,
on average, the single particle state associated with
the condensate (k = 0, where k is the momentum for
the particular case of homogeneous BEC that we will
be considering). This implies that the temperatures
to be considered should be smaller than the critical
temperature.
Accepted in Quantum 2019-12-28, click title to verify 3
For a single BEC, all the bosons condensate at the
same state. However, this will not be the case for the
two-component BEC and one has to determine the
fraction of particles in each component, which will
depend on the ground state of the system. This is
determined by the parameters of the system.
With the above considerations in mind, we assume
that the two bosonic gases condense. This means that
we can apply mean field theory and further assuming
that the ground state is coherent, the wavefunctions
Ψ
j
(x) , Ψ
j
(x) for a homogeneous BEC are given by
Ψ
j
(x) = Ψ
j,0
(x) + δΨ
j
(x), (4)
where Ψ
j,0
(x) = φ
0
(x)
p
N
j
e
j
, with θ
j
being the
phase of the coherent j
th
component, N
j
the num-
ber of bosons of the j
th
species and δΨ
j
(x) =
P
k6=0
φ
j,k
(x)a
j,k
with φ
k
(x) =
1
V
j
e
ikx/~
the plane
wave solutions, with V
j
the corresponding bath’s vol-
ume. From here onwards we assume for simplicity
that V
1
= V
2
= V , i.e. that the two baths have the
same volume, and that we are dealing with homo-
geneous BECs. Here a
j,k
and a
j,k
are bosonic anni-
hilation and creation operatos. To proceed further,
we write the Hamiltonian in terms of these operators.
The bosonic parts read
H
(j)
B
=
X
k
k
a
j,k
a
j,k
+
g
j
2
X
k,k
0
,q
a
j,k+q
a
j,k
0
q
a
j,k
a
j,k
0
,
H
(12)
B
= g
12
X
k,k
0
,q
a
1,k+q
a
2,k
0
q
a
2,k
a
1,k
0
+ ~
X
k
a
1,k
a
2,k
+H.c.,
with
k
= k
2
/2m
B
. The zeroth order expectation
value (or mean field value) of the Hamiltonian reads
as
H
0
=
X
j
g
j
2V
Ψ
4
j,0
+
g
12
V
Ψ
2
1,0
Ψ
2
2,0
+ ~
1,0
2,0
)
+
1,0
)
Ψ
2,0
) (5)
2.1 Generalized Bogoliubov transformation
We now perform a generalized Bogoliubov transfor-
mation to diagonalize the Bosonic part of the Hamil-
tonian and hence obtain the energy spectrum of the
bath. We follow closely the results of [5355] in
the rest of this section. This generalized Bogoliubov
transformation is understood to be composed of a ro-
tation, a scaling and one more rotation as in [53].
The derivation is based on a simple geometrical pic-
ture which results in a convenient parametrization of
the transformation. Following the generalized Bogoli-
ubov transformation, the initial bath operators are
transformed as
a
j,k
= Q
0
j,+,k
b
+,k
+ Q
1
j,+,k
b
+,k
+ (1)
δ
j,
Q
0
j,,k
b
,k
+ Q
1
j,,k
b
,k
a
j,k
= Q
1
j,+,k
b
+,k
+ Q
0
j,+,k
b
+,k
(6)
+ (1)
δ
j,
Q
1
j,,k
b
,k
+ Q
0
j,,k
b
,k
,
with b
+(),k
and b
+(),k
and the creation/annhilation
operators for the final spin (+) and density or phonon
() mode. In the latter, the total density fluctuates,
while in the spin mode (+) the unlike particle densi-
ties fluctuate out of phase. This, in the presence of
an internal Josephson interaction as in our case, is a
Josephson plasmon [56]. In Eq. (6) the parameters
are as follows: δ
1(2),(+)
= 1, δ
1(2),+()
= 0 and
Q
φ
j,s,k
= (7)
R
sj
ˆ
Γ
j,s,k
h
(1δ
j,s
) cos (γ
k
)+δ
j,s
sin (γ
k
)
i
cos θ
+R
sj
0
ˆ
Γ
j
0
,s,k
h
(1δ
j
0
,s
) cos (γ
k
)+δ
j
0
,s
sin (γ
k
)
i
sin θ,
where j
0
6= j, φ {0, 1}, s {+, −},
ˆ
Γ
j(j
0
),s,k
=
2
j(j
0
),s,k
+ (1)
φ
]/
j(j
0
),s,k
and R
1+
= R
2
=
(1, 1)
>
, R
2+
= R
1
= (1, 1)
>
. Also, in Eq. (7),
sin (γ
k
) = (8)
v
u
u
u
t
1
2
1
[ω
2
1,k
ω
2
2,k
]
q
(ω
2
1,k
ω
2
2,k
)
2
+ 16Λ
2
12
n
1
n
2
e
1,k
e
2,k
,
with cos (γ
k
) defined accordingly, and Γ
j,s,k
=
p
e
j,k
/E
s,k
where
E
±,k
:= ~
±,k
= (9)
P
j
ω
2
j,k
±
q
(ω
2
1,k
ω
2
2,k
)
2
+16Λ
2
12
n
1
n
2
e
1,k
e
2,k
2
1
2
,
with
e
j,k
=
k
(1)
j
1+ (1)
j
cos θ
12
~n
2n
1
n
2
ω
j,k
=
q
e
2
j,k
+
j
n
j
e
j,k
,
Λ
1
n
1
= g
1
n
1
cos
2
(θ)
+ g
2
n
2
sin
2
(θ)+g
12
sin (2θ) cos (θ
12
),
Λ
2
n
2
= g
1
n
1
sin
2
(θ)
+ g
2
n
2
cos
2
(θ)g
12
sin (2θ) cos (θ
12
),
Λ
12
n
1
n
2
=
g
2
n
2
g
1
n
1
2
sin(2θ)
+ g
12
n
1
n
2
cos (2θ)cos(θ
12
),
where n
j
=
N
j
V
is the particle density of the j
th
bath,
and θ is the free parameter (angle) to be determined
Accepted in Quantum 2019-12-28, click title to verify 4
by the minimisation of the total energy. In the above
expression, θ
12
= θ
1
θ
2
is the relative phase between
the two BEC. The minimization of the zeroth energy
with respect to the angle θ gives
tan(θ) =
q
n
1
n
2
, if θ
12
= π,
q
n
2
n
1
, if θ
12
= 0.
(10)
0.0 0.5 1.0 1.5 2.0
0.0
1.0
2.0
3.0
k
E
E
-
E
+
, =0
E
+
, =500
0.0 0.2
0.00
0.02
J
+
[ ]
0.01
0.0
0.2
0.000
0.001
0.002
ω
Ω
Ω
Figure 2: (a) Energy spectrum for a coherentely coupled
two component BEC. There are two branches in the spectrum
corresponding to the density (-) and spin modes (+). We plot
both for different values of the coherent coupling, . First,
this illustrates that the gap opens for the spin mode (+);
and second, it shows that, while for = 0 both branches
behave similarly, i.e., linearly for low k and quadratically for
large k, for finite the (+) mode behaves quadratically even
at low k. This has direct implications on the behavior of the
spectral density in case 2, plotted in (b). When the = 0
(blue line) the spectral density behaves as for the density
mode (i.e. with a w
3
-behavior). The red and green lines
(for Ω = 50π, 100πHz, respectively) show instead a different
behavior. The inset shows a zoom, where we checked that
it fits the simplified behavior in Eq. (50), i.e. has a lower
gap and behaves as
ω initially. In these plots we used
g = g
12
= 2.15 × 10
37
J · m , n = 7(µm)
1
, g
IB
=
0.5 × 10
37
J · m, with BEC and impurities made of Rb and
K atoms, respectively.
By minimizing with respect to the population im-
balance f =
N
1
N
2
N
, one can obtain the following con-
ditions on the parameters of the system in order to
have an extremum of the energy,
+ Af cos θ
12
f
(1 f
2
)
1/2
= 0, (11)
where A =
(g
1
+g
2
2g
12
)n
4~
is the mutual interaction
parameter, =
2δ+(g
1
g
2
)n
4~
is the effective detuning
parameter, and
A cos θ
12
1
(1 f
2
)
3/2
> 0, (12)
is the condition to have a minimum of the energy. In
[53], it was shown that to obtain the minimum energy
of the system, without imposing any condition on the
detuning δ, as is our case, then the relative phase
should be chosen to be θ
12
= π, referred to as the
π-state configuration. From here on we assume the
symmetric case, i.e, g
1
= g
2
= g as this will allow us
to obtain analytically the spectral density in section
3. In this case, the equilibrium condition Eq. (11)
reads as
g g
12
+
~
n
1
n
2
(n
1
n
2
) = 0, (13)
which has two solutions
n
1
n
2
= 0 (GS1) ,
n
1
n
2
= ±n
r
1
2~
(gg
12
)n
2
(GS2) ,
(14)
corresponding to neutral GS1 and polarized ground
states GS2. Here, we make the strong Josephson junc-
tion assumption
|A| < 1, (15)
which is also referred to as the miscibility condi-
tion. This implies that the minimum energy equi-
librium ground state has to be GS1 as is shown in
[53, 55]. Handable expressions for the spectral den-
sity obtained section 3 are possible over GS1. For the
regime in which ground state is GS2 we expect similar
qualitative behavior, but we did not obtained a form
for the spectral density which allows us to obtain the
diffusive behavior of the impurity. The study of the
impurity diffusion over GS2, and even at the phase
transition, falls out of the scope of this paper. Under
the miscibility condition (15), the energy spectrum
expressions simplifies into
E
,k
= (
k
(
k
+ (g + g
12
) n))
1
2
, (16)
E
+,k
= [
k
(
k
+ (g g
12
) n + 4~Ω)
+2~ [(g g
12
) n + 2~Ω]]
1
2
. (17)
In Fig. 2 we plot the energy spectra as a function
of k for specific parameters, to illustrate the spin and
density branches. Furthermore, note that the bogoli-
ubov transformation elements satisfy the well-known
relation
Q
0
k
(Q
0
k
)
T
Q
1
k
(Q
1
k
)
T
= 1, (18)
where Q
φ
k
=
Q
φ
1,+,k
Q
φ
1,,k
Q
φ
2,+,k
Q
φ
2,,k
!
with φ {0, 1},
that implies normalization. However, as is shown in
Accepted in Quantum 2019-12-28, click title to verify 5
[54], these Bogoliubov operators, do not fulfill the
bosonic commutations relations, which is understood
as a consequence of the fact that they are not orthogo-
nalized with respect to the quasicondensate functions
Ψ
j,0
=
p
N
j
e
j
. In [54] it is shown that to overcome
this problem one needs to define some new transfor-
mation with components
b
Q
0
j,s,k
,
b
Q
1
j,s,k
, that are re-
lated to the previous ones as
b
Q
φ
j,s,k
= Q
φ
j,s,k
Ψ
j,0
N
j
Q
φ
j,s,k
Ψ
j,0
. (19)
The elements of this transformation, these new Bo-
goliubov operators, are expressed in terms of the Bo-
goliubov wave functions of our system f
j,s,k
,
e
f
j,s,k
as
b
Q
φ
j,s,k
=
f
j,s,k
+ (1)
φ
e
f
j,s,k
2
, (20)
where
f
1,,k
= f
2,,k
=
k
2E
,k
1/2
,
e
f
1,,k
=
e
f
2,,k
=
E
,k
2
k
1/2
,
f
1,+,k
= f
2,+,k
=
k
+ ~
2E
+,k
1/2
,
e
f
1,+,k
=
e
f
2,+,k
=
E
+,k
2 (
k
+ ~Ω)
1/2
.
The spin mode branch is gapped while the density
mode branch is gapless. For the latter, at low values of
the momentum k the dispersion is linear, with a speed
of sound c
d
=
p
n (g + g
12
) /(2m
B
). On the contrary
for the gapped branch, the dispersion relation goes as
k
2
for low k, and at k = 0, it has a gap
E
gap
=
p
2~ [(g g
12
) n + 2~Ω]. (21)
This corresponds to the Josephson frequency for small
amplitude oscillations. As we will see the fact that
there are two branches in the spectrum will give rise
to two different noise sources.
Furthermore, one should note that, had we not in-
troduced the Rabi coupling term in the Hamiltonian,
the latter would commute with both n
1
and n
2
such
that we would have two broken continuous symme-
tries and both branches would be gapless (notice that
E
gap
0 when ~ 0). In this case, the low mo-
mentum excitations would be both phase-like, as it
has to be for Goldstone modes of the U (1)xU (1) bro-
ken symmetries [57]. Hence the introduction of the
Rabi coupling term, results in the system having only
one continuous broken symmetry, namely only n has
to be conserved now and not both n
1
and n
2
. The long
wavelength limit of the Goldstone mode corresponds
to a low-amplitude phonon fluctuation in which the
total density oscillates and the unlike atoms move in
unison (i.e., with the same superfluid velocity). In
contrast, in the long wavelength gap mode the un-
like atoms move in opposite directions, while their
center of mass remains at rest. This fluctuation is
then reminiscent of the motion of ions in an optical
phonon mode, which also exhibits a gapped disper-
sion. At zero momentum, the gap mode corresponds
to an infinitesimal Josephson-like oscillation of the
populations in the distinguishable internal states. In
the strong Josephson coupling regime we have closed
orbits around a fixed point for the Josephson Hamilto-
nian, with vanishing mean polarisation (or population
imbalance) and a phase difference around π if ~ 0,
giving rise to plasma-like oscillations.
2.2 Transfor med Impurity-Bath interaction
In terms of the original annihilation and creation op-
erators, the impurity-bath term reads as
H
IB
=
X
j
1
V
X
k,q
V
(j)
IB
(k) ρ
I
(q) a
j,kq
a
j,k
(22)
=
X
j
r
n
j
V
X
k6=0
ρ
I
(k) V
(j)
IB
a
j,k
+ a
j,k
,
with ρ
I
(q) =
R
−∞
e
iqx
0
δ (x
0
x) dx
0
, V
(j)
IB
(k) =
F
k
[g
(j)
IB
δ (x x
0
)] where F is the Fourier trasnform.
Furthermore, n
j
, with j = 1, 2, is the averaged den-
sity of the j
th
bath. The second line of Eq. (22) is a
consequence of the assumption that the bosons con-
dense. We will consider two cases for the coupling of
the impurity to the baths.
1. In the first scenario the impurity couples to the
two baths in the same way
g
(1)
IB
= g
(2)
IB
= g
IB
, (23)
2. while in the second scenario, the interactions are
attractive with one of the baths and repulsive
with the other the other,
g
(1)
IB
= g
(2)
IB
= g
IB
. (24)
After the Bogoliubov transformation the impurity-
bath term reads in both cases as
1. H
()
IB
=
h
n
V
i
1
2
X
j,k6=0
ρ
I
(k) g
IB
h
b
Q
0
j,,k
+
b
Q
1
j,,k
i
x
,k
,
2. H
(+)
IB
=
h
n
V
i
1
2
X
j,k6=0
ρ
I
(k) g
IB
h
b
Q
0
j,+,k
+
b
Q
1
j,+,k
i
x
+,k
,
where x
±,k
:=
b
±,k
+ b
±,k
. These equations show
that in case 1 the impurity only couples to the den-
sity () mode of the bosonic baths, while in case 2 it
Accepted in Quantum 2019-12-28, click title to verify 6
couples only to the spin (+) mode. For both cases,
we rewrite the impurity-bath terms as
H
s
IB
=
X
j,k6=0
s∈{+,−}
V
j,s,k
e
ikx
b
s,k
+ b
s,k
, (25)
where s = for case 1 and s = + for case 2, and
V
j,s,k
=
r
n
V
g
IB
b
Q
0
j,s,k
+
b
Q
1
j,s,k
. (26)
We note here that
b
Q
0
1,s,k
+
b
Q
1
1,s,k
=
b
Q
0
2,s,k
+
b
Q
1
2,s,k
,
such that V
1,s,k
= V
2,s,k
=
b
V
s,k
. We linearize the
interaction (see [35] for validity of this assumption)
to get
H
IB
=
X
k6=0
s∈{+,−}
V
s,k
(I + ikx)
b
s,k
+ b
s,k
. (27)
where V
s,k
= 2
b
V
s,k
. Thus, after a redefinition b
s,k
b
s,k
V
s,k
E
s,k
I, the final total Hamiltonian reads as
H = H
I
+
X
k6=0
s∈{+,−}
E
s,k
b
s,k
b
s,k
+
X
k6=0
s∈{+,−}
~g
s,k
π
s,k
,
(28)
with g
j,s,k
= kV
s,k
/~ and π
s,k
= i
b
k,s
b
k,s
the
momentum of the bath particles. We see that as in
[35], the coupling occurs between the position of the
impurity and the momentum of the bath particles.
However, in our work, the coupling can take place to
one of the two different quasiparticle branches, de-
pending on the form of the impurity-baths interac-
tions.
3 Spectral densities
The spectral densities can be obtained from the self-
correlation functions [35] for each environment (cor-
responding to cases 1 and 2). These read
C (t) =
X
k 6= 0
s {+, −}
~g
2
s,k
hπ
s,k
(t) π
s,k
(0)i/~. (29)
Using that the bath is composed of bosons for which
D
b
k,s
b
k,s
E
=
1
e
~ω
k
k
B
T
1
, (30)
we obtain
C (t)=
X
k6=0
s∈{+,−}
g
2
s,k
coth
~ω
k
2k
B
T
cos (ω
k
t)i sin (ω
k
t)
= ν (t) (t) , (31)
where
ν (t) =
Z
0
X
s∈{+,−}
J
D
(ω) coth
~ω
2k
B
T
cos (ωt) ,
λ (t) =
Z
0
X
s∈{+,−}
J
D
(ω) sin (ωt) . (32)
In these definitions we used the spectral density,
J
D
(ω) = ~
X
k6=0
(g
s,k
)
2
δ (ω ω
k
) . (33)
The spectral density is then evaluated in the contin-
uous frequency limit as
J
D
(ω) = (34)
4ng
2
IB
D
d
~ (2π)
d
Z
dkk
d+1
(U
s,k
+V
s,k
)
2
δ (kk
E
s
(ω))
k
E
s
(k)|
k=k
E
s
(ω)
,
where D
d
is the surface of the hypersphere in the mo-
mentum space with radius k in d-dimensions. In the
particular case of 1D becomes D
1
= 2.
To obtain the expression for the continuous fre-
quency case, the inverse of the dispersion relation
from Eq. (9) is needed. For this general energy spec-
trum, obtaining such inverse function is not easy.
However this is indeed possible for the simplified case,
Eqs. (16)(17). The inverse of the density (-) branch
which is the one to which the impurity couples for the
case 1 type of coupling, reads
k
E
(ω) = (35)
m
B
ng
g
12
g
1 +
s
1
2g
12
g
+
g
12
g
2
+
2ω
ng
2
1
2
.
With this, for the density () branch (case 1 type of
coupling), the spectral density is
J
(ω) = ˜τ
G
(ω)
3/2
p
F
(ω)
, (36)
with
F
(ω) = 1 +
ω
Λ
2
, (37)
G
(ω) = 1 +
p
F
(ω), (38)
˜τ
=
(2g
IB
)
2
nm
3/2
B
2
1/2
π
p
Λ
, (39)
and where Λ
= n (g + g
12
) /2~ is the cutoff fre-
quency, which resembles the one in [35] when g is
replaced by
g+g
12
2
. In the limit of ω Λ
, the spec-
tral density can be simplified to
J
(ω) = m
I
τ
ω
3
, (40)
where
τ
=
(2η)
2
2πm
I
m
B
n (g + g
12
)
1/3
!
3/2
, (41)
Accepted in Quantum 2019-12-28, click title to verify 7
with η
=
g
IB
g+g
12
. Thus, for the the π-state equilib-
rium configuration one obtains a cubic spectral den-
sity.
3.1 Spin branch coupling
For the spin (+) branch (case 2 type of coupling), the
inverse of the spectrum reads as
k
E
+
(ω) =
m
B
ng× (42)
g
12
g
1
4~
ng
+
s
1
2g
12
g
+
g
12
g
2
+
2ω
ng
2
1
2
.
In this case, the spectral density is
J
+
(ω) = m
I
eτ
+
G
+
(ω)
p
F
+
(ω)
, (43)
where
F
+
(ω) = 1 +
ω
Λ
+
2
,
G
+
(ω)=W (ω)
W (ω)+
1
2
1
2
s
1+
E
gap
Λ
+
2
1
2
,
W (ω) =
1
2
+
p
F
+
(ω)
1
2
s
1 +
E
gap
Λ
+
2
,
eτ
+
=
(2g
IB
)
2
nm
3/2
B
2
1/2
πm
I
p
Λ
+
, (44)
with Λ
+
= n (g g
12
) /2~. We note that to interpret
eτ
+
as a relaxation time, as is custom to do (see [35]),
one has to impose g g
12
to assure it remains a real
quantity. In other case, the spectral density will be
imaginary (note
G
+
(ω)
F
+
(ω)
is independent of the sign of
g g
12
).
Let us find how the spectral density in Eq. (43) sim-
plifies in two limiting cases. First, in the absence of
coherent coupling, Ω = 0, the gap vanishes, E
gap
= 0.
In this case, Eq. (43) is equal to that of the density
mode, upon the interchange Λ
Λ
+
. Therefore,
on the long time limit ω Λ
+
we obtain the same
cubic behavior of the spectral density. We illustrate
this case in Fig. 2. In panel (a) we show that the
two branches of the energy spectra have the same be-
havior, that is, linear at low k and parabolic for large
k.
Second, we consider the case of finite which im-
plies E
gap
> 0. A requirement which we impose on
the spectral density is that one cannot consider fre-
quencies lower than the gap energy E
gap
. Physically,
one can interpret this as follows: Since the energy
spectrum of the bath is gapped, with a gap given by
Eq. (21), the spectral density cannot assign a weight
at frequencies lower than this, because the bath can-
not excite the impurity with such frequencies since it
is not part of its spectrum. Then, we simplify the
spectral density as
b
J
+
(ω) = Θ (ω E
gap
) J
+
(ω) , (45)
with Θ(.) the step delta function.
Let us now comment on the frequency region right
above the energy gap of our system. To this end, we
replace ω = E
gap
+ , where > 0. We use as the
small value expansion parameter in our case, i.e. we
consider the limit E
gap
, such that ω E
gap
. We
furthermore introduce a cutoff Λ, for which it holds
that Λ E
gap
. The expressions in the spectral
density will now read as
F
+
() = 1 +
E
gap
Λ
+
2
, (46)
W () =
1
2
+
1
2
s
1+
E
gap
Λ
+
2
+
E
gap
Λ
+
E
2
gap
+ Λ
2
+
1
2
,
such that
G
+
() = (47)
1
2
+
1
2
s
1+
E
gap
Λ
+
2
E
gap
Λ
+
1
2
E
2
gap
+Λ
2
+
1
4
1
2
.
Hence the spectral density is
b
J
+
() = Θ () τ
+
1/2
, (48)
with
τ
+
= eτ
+
1
2
+
1
2
r
1 +
E
gap
Λ
+
2
!
(E
gap
)
1
2
1 +
E
gap
Λ
+
2
1/4
. (49)
The final form of the spectral density, after introduc-
ing a cutoff Λ, to avoid the related ultraviolet diver-
gencies mentioned above, is
b
J
+
(ω) = (50)
Θ (ω E
gap
) τ
+
[ω E
gap
]
1
2
Θ + E
gap
ω) .
We introduced a hard cutoff to our spectral density
as this better describes the physical system we study.
Such a spectral density, i.e. with an exponent on the
frequencies less than 1, is often associated to sub-
diffusive impurity dynamics. Note that in the re-
sults that we present below, we assume g g
12
as
in [58], which significantly simplifies the expression
for the coefficient of the spectral density τ
+
with-
out changing its behavior. In particular in this case
τ
+
= (2g
IB
)
2
nm
3/2
B
/2
1/2
πm
I
. In Fig. 2(b), we show
how the approximate spectral density in Eq. (50),
under the assumption E
gap
, compares to the
Accepted in Quantum 2019-12-28, click title to verify 8
original spectral density. For vanishingly small val-
ues of E
gap
the spectral density approaches the form
of that of the density mode, i.e. it goes as ω
3
, as
expected. In the limit we are interested, that is, for fi-
nite E
gap
, the spectral density behaves approximately
as in Eq. (50) (see inset in Fig. 2(b)).
Such gapped spectral densities as in Eq. (50), have
been studied extensively in the literature. In general,
they are usually related to semiconductors [5961] or
photonic crystals (PC) [62]. In particular, the simpli-
fied form of the spectral density, Eq. (50), is related in
particular with 3D PCs. The latter, are artificial ma-
terials engineered with periodic dielectric structures
[63]. If one considers an atom embedded in such a
material, it is known that if the resonant frequency of
the excited atom approaches the band gap edge of the
PC, strong localization of light, atom-photon bound
states, inhibition of spontaneous emission and frac-
tionalized steady-state inversion appear [6467]. The
rapidly varying distribution of field modes near the
band gap [62, 68] requires a non-Markovian descrip-
tion [69] of the reduced dynamics of quantum systems
coupled to the radiation field of a PC [66, 67, 7072].
This enhanced appearance of non-Markovian effects
is also confirmed by a recent study based on exact
diagonalization [73]. In this study it was observed
that for frequencies of the bath much larger than the
band gap, energy transfer between the system and
the bath is such that information and energy flow ir-
reversibly from system to bath leading to Markovian
dynamics. Conversely, at the edges of the gaps, one
observes the largest backflow of information where the
energy bounces between the system and bath lead-
ing to non-Markovian evolution of the system. Fur-
thermore, deep within the band gap, less excitations
and energy are exchanged between the system and
the bath, which is shown to lead to localized modes
[74], expressed as dissipationless oscillatory behavior,
plus non-exponential decays (such as for example frac-
tional relaxations [75]).
From the work in [74], a relationship is suggested
of such long-lived oscillations that appear in the dy-
namics of a system coupled to a bath with a gapped
spectral density, with the fact that the Hamiltonian of
the system might have thermodynamic and dynamic
instabilities. This is the case when the Hamiltonian
is unbounded from below, i.e. non-positive. Physi-
cally this happens when one deals with Hamiltonians
that do not conserve the particles number, and this is
indeed the case for the Hamiltonian of the free quan-
tum Brownian motion we study here. Hence, as a
result, this unbounded Hamiltonian induces dynami-
cal instabilities in the long-time regime, correspond-
ing to the limit ω Λ. In practice, as in the spirit
of [35], one can show that the effect of the bath on
the impurity is not only to dissipate its energy, but
as well to introduce an inverse parabolic potential in
which the impurity is diffusing (which would work as
a renormalization of the trapping frequency had we
considered a harmonic trapping potential). This in-
verse parabolic potential is understood to be a con-
sequence of the unboundedness of the Hamiltonian,
and is what is resulting in the dynamical instability
of the long time solution of the impurity dynamics.
Unfortunately, contrary to the case in [35], we will not
consider a harmonic trap for the impurity, and hence
the positivity of the Hamiltonian is violated irrespec-
tive of the strength of the coupling of the impurity to
the bath. In practice one would study the impurity
constrained in a box of a certain size, which if in-
cluded in the modelling of the system, would result in
a positively defined Hamiltonian, at the price of com-
plicating significantly the analytical solution for the
impurity’s dynamics. Hence, we assume here that we
are looking at timescales where the effect of the finite
sized box are not manifested. Theoretically, there are
also a number of other ways to circumvent this prob-
lem, even without referring to the presence of a box,
such as taking into account bilinear terms in the im-
purity’s or bath’s operators in the Hamiltonian as in
[7679]. In any case, we will show below that for the
regime of the transient effect that we are interested
in, this will not change our results.
Following the approach sketched above, we take ad-
vantage of the simplicity of the Fröhlich like Hamil-
tonian we are considering above. Furthermore, we
remind that we will look at the long time dynamics
of the impurity, i.e. ω Λ, as was implied by the
above study on the spectral density’s form. In addi-
tion, one should take into account the dissipationless
oscillatory behavior, which can also be expressed as
an incomplete decay of the Green function impurity
propagator which we will study below. In fact by
identifying the equivalence of the appearance of these
oscillations with the incomplete decay of the Green
function, this provides us with a very simple condition
upon which the long-lived oscillations appear, that is
that the Green function has at least one purely imagi-
nary pole, which can be shown to only be possible for
frequencies within the band gap [62]. We will study
this in the next section.
4 Heisenberg equations and their solu-
tion
In this section, we derive the equation of motion for
the impurity, which will allow us to study its diffusive
behavior under various scenarios. To do so, we be-
gin with the Heisenberg equations of motion for both
the impurity and the bath particles. The latter set
of equations can be solved, and we use this solution
to obtain a Langevin like equation of motion for the
impurity. The Heisenberg equations for the bath par-
Accepted in Quantum 2019-12-28, click title to verify 9
ticles are
db
s,k
(t)
dt
= i [H, b
s,k
(t)]
= i
s,k
b
s,k
(t) ~
2
X
j=1
g
(j)
s,k
x (t) ,
db
s,k
(t)
dt
=
i
~
h
H, b
s,k
(t)
i
=
i
~
s,k
b
s,k
(t) ~
2
X
j=1
g
(j)
s,k
x (t) , (51)
and for the central particle
dx (t)
dt
=
i
~
[H, x (t)] =
p (t)
m
I
, (52)
dp (t)
dt
=
i
~
[H, p (t)]
=
i
~
[U(x),x (t)]
X
k6=0
j={1,2}
s={+,−}
~g
(j)
s,k
π
s,k
(t) .
Substituting the solutions of the equations of motion
for the bath into that of the central particle, one gets
¨x (t) +
t
Z
Γ (t s) x (s) ds =
B (t)
m
I
, (53)
where
Γ (τ ) =
1
m
I
Z
0
P
j={1,2}
s={+,−}
J
(j)
s
(ω)
ω
cos (ωτ ) ,
B (t) = (54)
X
k6=0
i~
X
j={1,2}
s={+,−}
g
(j)
s,k
b
s,k
(t) e
k
t
b
s,k
(t) e
k
t
,
are the damping and noise terms, respectively. Note
that in Eq. (53) we neglected a term Γ (0) x (t).
This term may introduce dynamic instabilities in our
system in the long time regime. As in [35], we neglect
it as these instabilities are unphysical, that is, will
not occur in a physical realization of the system and
will only occur in the long time behavior. To be more
specific, for the coupling to the density mode this term
reads as,
Γ
(0) = τ
Λ
3
3
. (55)
For the coupling to the spin mode this term reads as
Γ
+
(0) = τ
+
"
πE
1/2
Gap
+ 2 + E
Gap
)
0.5
(56)
× F
2,1
1
2
,
1
2
;
1
2
;
E
Gap
Λ + E
Gap

.
As in [35], the solution of Eq. (53) takes the form
x(t) = (57)
G
1
(t)x(0) + G
2
(t) ˙x(0) +
1
m
I
Z
t
0
G
2
(t s)B(s)ds,
with the corresponding Green functions given by
L
z
[G
1
(t)] =
z
z
2
+ zL
z
[Γ(t)]
=
1
z + L
z
[Γ(t)]
, (58)
L
z
[G
2
(t)] =
1
z
2
+ zL
z
[Γ(t)]
, (59)
where L
z
[·] represents the Laplace transform. The
expressions for G
1
(t) and G
2
(t) depend on the spe-
cific type of bath we consider. For the first scenario
(coupling to the density mode), in [35] it was found
that Γ(t) is
Γ
(t) = (60)
τ
t
3
t cos t) 2
2 Λ
2
t
2
sin
t)
,
and under the assumption of z Λ
L
z
(t)] = τ
Λ
z + O
z
2
, (61)
which results in
L
z
[G
1
(t)] =
1
(1 + Λ
τ
) z
, (62)
and
L
z
[G
2
(t)] =
1
(1 + Λ
τ
) z
2
. (63)
Then, one obtains
G
1
(t) =
1
(1 + Λ
τ
)
, (64)
and
G
2
(t) =
t
(1 + Λ
τ
)
, (65)
where we see that the Green functions has an iden-
tical form to that of [35]. More importantly, G
2
(t)
diverges at t , a consequence of the fact that an
equilibrium state is not reached at this limit, as we
see in next section.
For the second case (coupling to the spin mode),
as discussed in previous section the spectral density
is gapped and given by Eq. (50). To proceed, we
first need an expression for the Laplace transform
L
z
+
(t)] which can be shown to read as
L
z
+
(t)] =
Z
0
"
Z
0
b
J
+
(ω)
ω
cos (ωt)
#
e
zt
dt
= z
Z
0
b
J
+
(ω)
ω (ω
2
+ z
2
)
=
τ
+
z
E
3
gap
+E
gap
z
2
Λ
1.5
×
E
gap
3
(E
gap
+iz)F
2,1
1,
3
2
;
5
2
;
Λ
E
gap
iz
E
gap
3
(E
gap
iz) F
2,1
1,
3
2
;
5
2
;
Λ
E
gap
+ iz
+
2
3
E
2
gap
+ z
2
F
2,1
1,
3
2
;
5
2
;
Λ
E
gap

, (66)
Accepted in Quantum 2019-12-28, click title to verify 10
where F
2,1
(α, β; γ; z) is the hypergeometric function
F
2,1
(α, β; γ; z) =
X
n=0
(α)
n
(β)
n
(γ)
n
z
n
n!
, (67)
with (·)
n
being the Pochhammer symbol. Unfortu-
nately, to invert the Laplace transform in Eq. (59),
given Eq. (66), is rather complicated. For this rea-
son we restrain ourselves to study only the long-time
limit, determined by z Λ. In this case the inverse
Laplace transform of the Green’s function in Eq. (59),
reads as
G
2
(t) = At, (68)
A=E
5
gap
E
5
gap
+
2
3
E
2
gap
Λ
1.5
τ
+
F
2,1
1,
3
2
;
5
2
;
Λ
E
gap
4
5
E
gap
Λ
2.5
τ
+
F
2,1
2,
5
2
;
7
2
;
Λ
E
gap
+0.285714Λ
3.5
τ
+
F
2,1
3,
7
2
;
9
2
;
Λ
E
gap

1
,
which has the same time dependence as in case 1 (cou-
pling to the density mode). We also find G
1
(t) = A
which is again a constant and hence does not play a
role in the time dependence of the MSD which we are
interested in. Unfortunately this is as far as we can get
analytically, as contrary to the coupling to the den-
sity mode, even though we have the Green function
at hand, using it to obtain an analytic expression for
the MSD of the impurity which is our ultimate goal
is not possible.
Equation (65) as well as Eq. (66) have been both
obtained at the long time limit, which implied ex-
panding the Laplace transform of the damping kernel
L
z
(t)] , L
z
+
(t)] at the first order in z/Λ
, z/Λ.
In general, one could have considered higher orders
of the aforementioned expansion, but should then be
careful in inverting the Laplace transform to obtain
the Green function in defining the relevant Bromwich
integral in the complex plane in such a way as to not
include the roots which correspond to divergent run-
away solutions [35]. Even if one would do so, the
result for the Green function would not change much,
and this can be proven by considering a numerical in-
version of the Laplace transform for the Green func-
tion, where the long time limit assumption is not
made. For the coupling to the density mode this
was shown using the Zakian method in [35]. For the
spin mode, we checked this using the same method.
Moreover, we contrasted its results to two other meth-
ods for numerically inverting a Laplace transform, in
particular, the Fourier and the Stehfest methods [80].
The Zakian method, gives the inverse of the Laplace
transform of a function F (z) in the following form
f (t) =
2
t
N
X
j=1
Re
k
j
F
β
j
t

, (69)
where k
j
and β
j
are real and complex constants given
in [80]. With all of these methods, the Green function
behaves linearly with time for the range of parame-
ters we considered. In fact in the numerical results
presented in the next section, the Zakian method was
used to obtain the Green function, such that our re-
sults are not restricted just to the long time limit,
z Λ.
In addition, we are also now in a position to check
the presence of the long-lived oscillations in our sys-
tem. As was mentioned before, this can only be the
case if the Green function exhibits a purely imaginary
pole, which if it exists, should correspond to a fre-
quency within the bandgap. As is shown in [81], this
will be the case, for the frequency that is a solution
of
ω
2
+ Γ (0) (ω) = 0, (70)
where
(ω) := P
Z
0
b
J
+
(ω
0
)
ω ω
0
0
=
2τ
+
Λ
1.5
F
2,1
1,
3
2
;
5
2
;
Λ
E
gap
ω
3 (E
gap
ω)
, (71)
is the bath self energy correction, where P denotes the
principal value. One can show that the expression of
Eq. (71) is always non-positive for ω < E
gap
and hence
the condition in Eq. (70) is never satisfied, such that
we do not have to worry about these oscillations in
the transient dynamics that we will study in the next
section.
Finally, one can evaluate the validity of the linear-
ity assumption which allowed us to consider a lin-
ear coupling between the BEC and the impurity (see
Hamiltonian in Eq. (27)) in terms of the physical pa-
rameters of the system. This assumption reads as
kx 1. In [35] it was shown that, as a function
of the temperature, there exist a maximum time for
which the linear assumption holds. In the system dis-
cussed here, since an expression for the MSD cannot
be found, for each set of parameters one has to eval-
uate numerically the long-time behavior of the MDS
and determine the maximum time for which the as-
sumption holds. To this end, one has to note that, dif-
ferently to [35], for the coupling to the spin mode the
momenta grows parabolically with ω even for small k
and there is an energy gap. Then, to evaluate the cri-
teria (kx 1), one has to use the expression for the
energy, Eq. (17) together with the numerically evalu-
ated MSD. We checked this condition in the numerical
examples presented in next section. A final comment
is that, we also made sure that the assumption of a
purely 1D BEC was valid, by checking that the spa-
tial extend of the motion did not exceed the phase
coherence length [52].
Accepted in Quantum 2019-12-28, click title to verify 11
Log
10
[<x
2
(t)>/a
HO
]
*
*
0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6
- 0.5
0.0
0.5
1.0
+
=1.5
+
=2
+
=3
(b)
Figure 3: Mean square displacement vs time for the case
of coupling to the spin mode. A cutoff of Λ = 10
¯
was
used, where
¯
= 1000πHz. In (a) we plot it for different
coherent couplings and in (b) for different couplings to the
bath. The MSD shows three regimes, where it behaves ap-
proximately as MSD(t) t
α
, and therefore linearly in log-log
pots, with a different slope given by the anomalous exponent
α: (i) an initial short time behavior, where α 2; (ii) a
nontrivial transient subdiffusive behavior, where α < 1. We
plot a dashed orange line as a guide to the eye, to illustrate
the different slopes in this regime; (iii) a long time ballistic
regime, with α = 2. In (a) we show that, as is reduced,
the subdiffusive platteau enlarges and α gets smaller. In
(b) we show that increasing the couplings to the bath τ
+
,
also enlarges the plateau and reduces α. We consider Rb
and K atoms for BEC and impurities, respectively. We use
g = g
12
= 2.15 × 10
37
J · m, density n = 7(µm)
1
, and
impurity-BEC g
IB
= 0.5 × 10
37
J · m; We take τ
+
= 1 in
(a) and = 100πHz in (b). The BEC was assumed to be
in the low temperature regime, i.e. when coth
~ω
2k
B
T
1
holds.
5 Results: Mean square displacement
With the Green propagator and the spectral density
at hand, we are now in a position to evaluate the
MSD. This, as shown in [35], is evaluated in the long
time limit ω Λ
, Λ as
h[x (t) x (0)]
2
i = MSD(t) = G
2
2
(t)h˙x
2
(0)i
+
1
2
Z
t
0
ds
Z
t
0
G
2
(s)G
2
(σ)h{B(s), B(σ)}i
ρ
B
, (72)
where we assumed that the impurity-bath are initially
in a product state ρ (0) = ρ
B
ρ
S
(0), where ρ
B
is
the thermal Gibbs state for the bath at temperature
T . The initial conditions of the impurity and bath
oscillators are then uncorrelated. Then, averages of
the form h˙x (0) B(s)i vanish. To treat the second term
in Eq. (72), we note that
h{B(s), B(σ)}i
ρ
B
= 2~ν (s σ) , (73)
where ν (t) is defined as in Eq. (32).
In case 1 (coupling to density mode) the spectral
density reads as in Eq. (40). Then, the MSD behaves
the same way as in [35], with the only difference of
replacing g g + g
12
. Hence the impurity will again
superdiffuse as
h[x (t) x (0)]
2
i =
h˙x
2
(0)i +
τ
Λ
2
2
t
ζ
2
, (74)
where ζ = 1 + τ
Λ
. Note that the superdiffusive
behavior hx
2
(t)i t
2
appears for both low temper-
ature (coth (~ω/2k
B
T ) 1) and high temperature
(coth (~ω/2k
B
T ) 2k
B
T/~ω) limits. Hence from
Eq. (74) we see that, effectively, the contribution of
the Bogoliubov modes to the MSD behavior in this
case is just to modify the mass of the free particle.
In case 2 (coupling to the spin mode), analytical
expressions for the Eq. (72) cannot be found. We
remind again that we are interested in the transient
effects attributed to the bath frequencies right above
the band gap, after making the assumption for ω =
E
gap
+ that ω E
gap
i.e. E
gap
. In this case the
Green function reads as in Eq. (50), while the noise
kernel at low temperatures, where coth
~ω
2k
B
T
1,
can be shown to be equal to
ν (t) = τ
+
E
gap
)
1.5
(75)
×
2
3
cos (E
gap
t)F
1,2
3
4
;
1
2
,
7
4
;
1
4
t
2
(E
gap
Λ)
2
+
2
5
t (E
gap
Λ) cos (E
gap
t)
×F
1,2
5
4
;
3
2
,
9
4
;
1
4
t
2
(E
gap
Λ)
2

,
where
F
1,2
(α; β, γ; z) =
X
n=0
(α)
n
(β)
n
(γ)
n
z
n
n!
. (76)
In this case, we were not able to obtain an analytic
solution for the MSD. We evaluate it numerically, with
the results being valid for finite , as we used the
simplified version of the spectral density, Eq. (50). In
all calculations we checked that all assumptions made
are fulfilled.
In Fig. 3 we show the numerically evaluated MSD
as a function of time according to Eq. (72) and dif-
ferent Rabi frequencies and interaction strengths. We
remind that initially, h˙x
2
(0)i = 0. In Fig. 3 (a) we
show how decreasing both enlarges the duration of
Accepted in Quantum 2019-12-28, click title to verify 12
the subdiffusive plateau and reduces the anomalous
exponent α. We should note that the results are valid
only for finite : since we use the simplifies spectral
density, Eq. (50), we are never able to describe the
smooth transition to the cubic spectral density, which
will show an smooth change to ballistic behavior for
the whole range. Then, the effect of reducing is
merely to reduce the gap, not to change the form of
the spectral density. As a consequence, the plateau is
enlarged. In Fig. 3 (b) we show how the MSD varies
as a function of the coupling stength of the impurity
to the BECs. Here, we observe that increasing the
coupling strength results in more subdiffusive motion
and an increase in the duration of the subdiffusive
plateau. Furthermore, note that, in Fig. 3, time is
measured in units of the inverse of
¯
Ω = 1000πHz and
hence the transient subdiffusive phenomenon appears
in time of the order of ms.
In general, from the results presented in Fig. 3,
we numerically find three regimes of behavior for the
MSD, and in each regime it behaves as MSD(t) t
α
,
where α is different at each regime. The exponent
α is known as the anomalous exponent. In regime
(i), there is an initial short time behavior where, as
expected, the MSD grows more or less ballistically
with time. So here, α 2; In regime (ii), there is a
plateau where α < 1. This is a transient subdiffusive
behavior; Finally, in regime (iii), which is the long
time behavior, the impurity superdiffuses with α = 2.
We interpret this behavior as follows: the impurity
performs free motion initially. Then after interact-
ing with the large frequencies of the bath, the im-
purity begins to perform a subdiffusive motion since
it screens the part of the spectral density that de-
pends on the square root of the bath modes frequen-
cies. At long times, and after undergoing dissipation
for some time, the impurity again effectively only in-
teracts with the lower frequencies of the bath which
have zero effect on the motion of the impurity and
hence the impurity performs a ballistic motion.
Finally, we comment here that the results of the nu-
merical integrations undertaken to obtain the MSD in
this section, indeed appear to agree with the analyti-
cal calculations we performed in the previous section.
This is to be understood in the following sense. The
analytical results of the previous section, implied that
the coupling of the impurity to the spin mode, results
in the impurity interacting with a bath giving rise
to a different spectral density, which intuitively one
expects to give a different MSD behaviour. This is
indeed what we observe in the results presented in
Fig. 3.
6 Conclusions
In this work, we studied the diffusive behavior of
an impurity immersed in a coherently coupled two-
component BEC, that interacts with both of them
through contact interactions. We showed how start-
ing from the standard Hamiltonian that would de-
scribe such a scenario, one can recast the problem
into that of a quantum Brownian particle diffusing
in a bath composed of the Bogoliubov modes of the
two-component BEC. We discussed the under certain
assumptions and conditions required to obtain this
description.
We found that the main difference of this scenario
compared to that of the impurity being coupled to
a single BEC studied in [35], is that for the scenario
of the impurity being coupled differently to the two
BECs, namely coupled attractively to one of them and
repulsively to the other but with the same magnitude,
results in the impurity being coupled to the spin mode
of the coherently coupled two-component BEC. This
implies that its dynamics is determined by a quali-
tatively different spectral density. In particular this
new spectral density is gapped and subohmic close to
the gap. We demonstrate numerically, that such a
spectral density gives rise to a transient subdiffusive
behavior. Furthermore, we show that this transient
effect can be controlled by the magnitude of the Rabi
frequency, as well as by the strength with which the
impurity couples to the two BECs. These can control
the time duration for which this subdiffusive behavior
appears. A mechanism for inducing a transient con-
trolled subdiffusion in Brownian motion has been also
proposed in [82], but with a completely different way
for achieving it and most importantly not considering
the system from a microscopic perspective. Moreover,
we comment that the setup we studied, thanks to the
appearance of this gapped subohmic spectral density,
could also serve for simulating quantum-optical phe-
nomena, that could be seen for example in photonic
crystals, using instead cold atoms, as was proposed
also in [83] for the case of optical lattices. In addi-
tion, we note that our studies could be extended to
the scenario of having two impurities in the coherently
coupled two-component BEC, and study as in [37],
the effects that the coupling to the spin mode could
have on the bath-induced entanglement between the
two impurities. Finally, we could also study the effect
that this new gapped spectral density could have on
the functioning of the impurity as a probe to measure
the temperature of the two-component BEC, as in
[84]. Last but not least, it should be noted here that,
if one considers the scenario of attractive two-body
coupling, i.e. g
12
< 0, and includes the Lee-Huang-
Yang corrections to the Hamiltonian, one will obtain
the scenario of quantum droplets studied theoretically
in [85] and recently proven experimentally in [86] .
This we expect to lead in different interesting dynam-
ics for the immersed impurity in the two-component
coherently coupled BEC.
Accepted in Quantum 2019-12-28, click title to verify 13
Acknowledgments
We (M.L. group) acknowledge the Spanish Min-
istry MINECO (National Plan 15 Grant: FISI-
CATEAMO No. FIS2016-79508-P, FPI), the Ministry
of Education of Spain (FPI Grant BES-2015-071803),
EU FEDER, European Social Fund, FundaciÃş
Cellex, Generalitat de Catalunya (AGAUR Grant No.
2017 SGR 1341 and CERCA/Program), ERC AdG
OSYRIS and NOQIA, EU FETPRO QUIC, and the
National Science Centre, Poland-Symfonia Grant No.
2016/20/W/ST4/00314. MAGM acknowledges fund-
ing from the Spanish Ministry of Education and Voca-
tional Training (MEFP) through the Beatriz Galindo
program 2018 (BEAGAL18/00203).
References
[1] P. HÃďnggi and F. Marchesoni. Introduction:
100years of brownian motion. Chaos: An In-
terdisciplinary Journal of Nonlinear Science, 15
(2):026101, 2005. DOI: 10.1063/1.1895505. URL
https://doi.org/10.1063/1.1895505.
[2] I. M. Sokolov and J. Klafter. From diffusion
to anomalous diffusion: A century after ein-
steinâĂŹs brownian motion. Chaos: An Inter-
disciplinary Journal of Nonlinear Science, 15(2):
026103, 2005. DOI: 10.1063/1.1860472. URL
https://doi.org/10.1063/1.1860472.
[3] H. Scher and E.W. Montroll. Anomalous transit-
time dispersion in amorphous solids. Phys. Rev.
B, 12:2455–2477, Sep 1975. DOI: 10.1103/Phys-
RevB.12.2455. URL https://link.aps.org/
doi/10.1103/PhysRevB.12.2455.
[4] A. Bunde and S. Havlin. Fractals in science.
Springer-Verlag Berlin Heidelberg, 1994. DOI:
10.1007/978-3-662-11777-4.
[5] M J Saxton. Lateral diffusion in an
archipelago. single-particle diffusion. Biophys J,
64, 1993. DOI: 10.1016/S0006-3495(93)81548-
0. URL https://www.ncbi.nlm.nih.gov/
pubmed/8369407.
[6] M. J. Saxton. Single-particle tracking: the dis-
tribution of diffusion coefficients. Biophys J,
72, 1997. DOI: 10.1016/S0006-3495(97)78820-
9. URL https://www.ncbi.nlm.nih.gov/
pubmed/9083678.
[7] F. Leyvraz, J. Adler, A. Aharony, A. Bunde,
A. Coniglio, D.C. Hong, H.E. Stanley, and
D. Stauffer. The random normal supercon-
ductor mixture in one dimension. Journal
of Physics A: Mathematical and General, 19
(17):3683–3692, dec 1986. DOI: 10.1088/0305-
4470/19/17/030. URL https://doi.org/10.
1088%2F0305-4470%2F19%2F17%2F030.
[8] S. Hottovy, G. Volpe, and J. Wehr. Noise-
induced drift in stochastic differential equa-
tions with arbitrary friction and diffusion
in the smoluchowski-kramers limit. Jour-
nal of Statistical Physics, 146(4):762–773, Feb
2012. ISSN 1572-9613. DOI: 10.1007/s10955-
012-0418-9. URL https://doi.org/10.1007/
s10955-012-0418-9.
[9] A.G. Cherstvy and R. Metzler. Population
splitting, trapping, and non-ergodicity in het-
erogeneous diffusion processes. Phys. Chem.
Chem. Phys., 15:20220–20235, 2013. DOI:
10.1039/C3CP53056F. URL http://dx.doi.
org/10.1039/C3CP53056F.
[10] A.G. Cherstvy, A.V. Chechkin, and R. Met-
zler. Anomalous diffusion and ergodic-
ity breaking in heterogeneous diffusion pro-
cesses. New Journal of Physics, 15(8):
083039, aug 2013. DOI: 10.1088/1367-
2630/15/8/083039. URL https://doi.org/10.
1088%2F1367-2630%2F15%2F8%2F083039.
[11] A.G. Cherstvy, A.V. Chechkin, and R. Metzler.
Particle invasion, survival, and non-ergodicity in
2d diffusion processes with space-dependent dif-
fusivity. Soft Matter, 10:1591–1601, 2014. DOI:
10.1039/C3SM52846D. URL http://dx.doi.
org/10.1039/C3SM52846D.
[12] P. Massignan, C. Manzo, J. A. Torreno-Pina,
M. F. García-Parajo, M. Lewenstein, and G. J.
Lapeyre. Nonergodic subdiffusion from brownian
motion in an inhomogeneous medium. Phys. Rev.
Lett., 112:150603, Apr 2014. DOI: 10.1103/Phys-
RevLett.112.150603. URL https://link.aps.
org/doi/10.1103/PhysRevLett.112.150603.
[13] Carlo Manzo, Juan A. Torreno-Pina, Pietro
Massignan, Gerald J. Lapeyre, Maciej Lewen-
stein, and Maria F. Garcia Parajo. Weak er-
godicity breaking of receptor motion in living
cells stemming from random diffusivity. Phys.
Rev. X, 5:011021, Feb 2015. DOI: 10.1103/Phys-
RevX.5.011021. URL https://link.aps.org/
doi/10.1103/PhysRevX.5.011021.
[14] C. Charalambous, G. Muñoz Gil, A. Celi,
M. F. Garcia-Parajo, M. Lewenstein, C. Manzo,
and M. A. García-March. Nonergodic subd-
iffusion from transient interactions with het-
erogeneous partners. Phys. Rev. E, 95:
032403, Mar 2017. DOI: 10.1103/Phys-
RevE.95.032403. URL https://link.aps.org/
doi/10.1103/PhysRevE.95.032403.
[15] B. Min, T. Li, M. Rosenkranz, and W. Bao.
Subdiffusive spreading of a bose-einstein con-
densate in random potentials. Phys. Rev. A,
86:053612, Nov 2012. DOI: 10.1103/Phys-
RevA.86.053612. URL https://link.aps.org/
doi/10.1103/PhysRevA.86.053612.
[16] G. Roati, C. DâĂŹErrico, L. Fallani, M. Fattori,
C. Fort, M. Zaccanti, G. Modugno, M. Mod-
ugno, and M. Inguscio. Anderson localization of a
non-interacting boseâĂŞeinstein condensate. Na-
Accepted in Quantum 2019-12-28, click title to verify 14
ture, 453, 2008. DOI: 10.1038/nature07071. URL
https://doi.org/10.1038/nature07071.
[17] F. Jendrzejewski, A. Bernard, K. MÃijller,
P. Cheinet, V. Josse, M. Piraud, L. Pez-
zÃľ, L. Sanchez-Palencia, A. Aspect, and
P. Bouyer. Three-dimensional localization of
ultracold atoms in an optical disordered po-
tential. Nature Physics, 8, 2012. DOI:
10.1038/nphys2256. URL https://doi.org/
10.1038/nphys2256.
[18] L. Sanchez-Palencia and M. Lewenstein. Dis-
ordered quantum gases under control. Nature
Physics, 6, 2010. DOI: 10.1038/nphys1507. URL
https://doi.org/10.1038/nphys1507.
[19] G. Modugno. Anderson localization in
bose–einstein condensates. Reports on Progress
in Physics, 73(10):102401, sep 2010. DOI:
10.1088/0034-4885/73/10/102401. URL
https://doi.org/10.1088%2F0034-4885%
2F73%2F10%2F102401.
[20] J. Billy, V. Josse, Z. Zuo, A. Bernard, B. Ham-
brecht, P. Lugan, D. ClÃľment, L. Sanchez-
Palencia, P. Bouyer, and A. Aspect. Direct obser-
vation of anderson localization of matter waves in
a controlled disorder. Nature, 453, 2008. DOI:
10.1038/nature07000. URL https://doi.org/
10.1038/nature07000.
[21] B. Deissler, M. Zaccanti, G. Roati, C. DâĂŹEr-
rico, M. Fattori, M. Modugno, G. Modugno,
and M. Inguscio. Delocalization of a disordered
bosonic system by repulsive interactions. Nature
Physics, 6, 2010. DOI: 10.1038/nphys1635. URL
https://doi.org/10.1038/nphys1635.
[22] E. Lucioni, B. Deissler, L. Tanzi, G. Roati,
M. Zaccanti, M. Modugno, M. Larcher, F. Dal-
fovo, M. Inguscio, and G. Modugno. Ob-
servation of subdiffusion in a disordered in-
teracting system. Phys. Rev. Lett., 106:
230403, Jun 2011. DOI: 10.1103/Phys-
RevLett.106.230403. URL https://link.aps.
org/doi/10.1103/PhysRevLett.106.230403.
[23] Stefan Donsa, Harald Hofstätter, Othmar
Koch, Joachim Burgdörfer, and Iva Březi-
nová. Long-time expansion of a bose-
einstein condensate: Observability of an-
derson localization. Phys. Rev. A, 96:
043630, Oct 2017. DOI: 10.1103/Phys-
RevA.96.043630. URL https://link.aps.org/
doi/10.1103/PhysRevA.96.043630.
[24] D. L. Shepelyansky. Delocalization of quantum
chaos by weak nonlinearity. Phys. Rev. Lett.,
70:1787–1790, Mar 1993. DOI: 10.1103/Phys-
RevLett.70.1787. URL https://link.aps.org/
doi/10.1103/PhysRevLett.70.1787.
[25] G. Kopidakis, S. Komineas, S. Flach, and
S. Aubry. Absence of wave packet diffusion
in disordered nonlinear systems. Phys. Rev.
Lett., 100:084103, Feb 2008. DOI: 10.1103/Phys-
RevLett.100.084103. URL https://link.aps.
org/doi/10.1103/PhysRevLett.100.084103.
[26] A. S. Pikovsky and D. L. Shepelyansky.
Destruction of anderson localization by a
weak nonlinearity. Phys. Rev. Lett., 100:
094101, Mar 2008. DOI: 10.1103/Phys-
RevLett.100.094101. URL https://link.aps.
org/doi/10.1103/PhysRevLett.100.094101.
[27] S. Flach, D. O. Krimer, and Ch. Skokos.
Universal spreading of wave packets in disor-
dered nonlinear systems. Phys. Rev. Lett.,
102:024101, Jan 2009. DOI: 10.1103/Phys-
RevLett.102.024101. URL https://link.aps.
org/doi/10.1103/PhysRevLett.102.024101.
[28] Ch. Skokos, D. O. Krimer, S. Komineas, and
S. Flach. Delocalization of wave packets in
disordered nonlinear chains. Phys. Rev. E,
79:056211, May 2009. DOI: 10.1103/Phys-
RevE.79.056211. URL https://link.aps.org/
doi/10.1103/PhysRevE.79.056211.
[29] Hagar Veksler, Yevgeny Krivolapov, and Shmuel
Fishman. Spreading for the generalized nonlin-
ear schrödinger equation with disorder. Phys.
Rev. E, 80:037201, Sep 2009. DOI: 10.1103/Phys-
RevE.80.037201. URL https://link.aps.org/
doi/10.1103/PhysRevE.80.037201.
[30] M. Mulansky and A. Pikovsky. Spread-
ing in disordered lattices with different non-
linearities. EPL (Europhysics Letters), 90
(1):10015, apr 2010. DOI: 10.1209/0295-
5075/90/10015. URL https://doi.org/10.
1209%2F0295-5075%2F90%2F10015.
[31] T. V. Laptyeva, J. D. Bodyfelt, D. O. Krimer,
Ch. Skokos, and S. Flach. The crossover from
strong to weak chaos for nonlinear waves in dis-
ordered systems. EPL (Europhysics Letters),
91(3):30001, aug 2010. DOI: 10.1209/0295-
5075/91/30001. URL https://doi.org/10.
1209%2F0295-5075%2F91%2F30001.
[32] A. Iomin. Subdiffusion in the nonlinear
schrödinger equation with disorder. Phys. Rev.
E, 81:017601, Jan 2010. DOI: 10.1103/Phys-
RevE.81.017601. URL https://link.aps.org/
doi/10.1103/PhysRevE.81.017601.
[33] M. Larcher, F. Dalfovo, and M. Mod-
ugno. Effects of interaction on the diffu-
sion of atomic matter waves in one-dimensional
quasiperiodic potentials. Phys. Rev. A, 80:
053606, Nov 2009. DOI: 10.1103/Phys-
RevA.80.053606. URL https://link.aps.org/
doi/10.1103/PhysRevA.80.053606.
[34] L.M. Aycock, H.M. Hurst, D.K. Efimkin,
D. Genkina, H.-I. Lu, V.M. Galitski, and
I. B. Spielman. Brownian motion of soli-
tons in a bose–einstein condensate. Proceed-
ings of the National Academy of Sciences, 114
(10):2503–2508, 2017. ISSN 0027-8424. DOI:
Accepted in Quantum 2019-12-28, click title to verify 15
10.1073/pnas.1615004114. URL https://www.
pnas.org/content/114/10/2503.
[35] A. Lampo, S.H. Lim, M.A. García-March, and
M. Lewenstein. Bose polaron as an instance
of quantum Brownian motion. Quantum, 1:
30, September 2017. ISSN 2521-327X. DOI:
10.22331/q-2017-09-27-30. URL https://doi.
org/10.22331/q-2017-09-27-30.
[36] A. Lampo, C. Charalambous, M.A. García-
March, and M. Lewenstein. Non-markovian
polaron dynamics in a trapped bose-einstein
condensate. Phys. Rev. A, 98:063630, Dec
2018. DOI: 10.1103/PhysRevA.98.063630.
URL https://link.aps.org/doi/10.1103/
PhysRevA.98.063630.
[37] C. Charalambous, M.A. Garcia-March,
A. Lampo, M. Mehboudi, and M. Lewenstein.
Two distinguishable impurities in BEC: squeez-
ing and entanglement of two Bose polarons.
SciPost Phys., 6:10, 2019. DOI: 10.21468/Sci-
PostPhys.6.1.010. URL https://scipost.org/
10.21468/SciPostPhys.6.1.010.
[38] D.K. Efimkin, J. Hofmann, and V. Galit-
ski. Non-markovian quantum friction of bright
solitons in superfluids. Phys. Rev. Lett.,
116:225301, May 2016. DOI: 10.1103/Phys-
RevLett.116.225301. URL https://link.aps.
org/doi/10.1103/PhysRevLett.116.225301.
[39] H.M. Hurst, D.K. Efimkin, I. B. Spielman,
and V. Galitski. Kinetic theory of dark soli-
tons with tunable friction. Phys. Rev. A,
95:053604, May 2017. DOI: 10.1103/Phys-
RevA.95.053604. URL https://link.aps.org/
doi/10.1103/PhysRevA.95.053604.
[40] A. Cem Keser and V. Galitski. Analogue stochas-
tic gravity in strongly-interacting boseâĂŞe-
instein condensates. Annals of Physics, 395:
84 111, 2018. ISSN 0003-4916. DOI:
https://doi.org/10.1016/j.aop.2018.05.009. URL
http://www.sciencedirect.com/science/
article/pii/S0003491618301453.
[41] Julius Bonart and Leticia F. Cugliandolo. From
nonequilibrium quantum brownian motion to
impurity dynamics in one-dimensional quan-
tum liquids. Phys. Rev. A, 86:023636, Aug
2012. DOI: 10.1103/PhysRevA.86.023636.
URL https://link.aps.org/doi/10.1103/
PhysRevA.86.023636.
[42] X.-D. Bai and J.-K. Xue. Subdiffusion
of dipolar gas in one-dimensional quasiperi-
odic potentials. Chinese Physics Letters, 32
(1):010302, jan 2015. DOI: 10.1088/0256-
307x/32/1/010302. URL https://doi.org/10.
1088%2F0256-307x%2F32%2F1%2F010302.
[43] K.-T. Xi, J. Li, and D.-N. Shi. Localization
of a two-component boseâĂŞeinstein conden-
sate in a two-dimensional bichromatic optical
lattice. Physica B: Condensed Matter, 436:
149 156, 2014. ISSN 0921-4526. DOI:
https://doi.org/10.1016/j.physb.2013.12.010.
URL http://www.sciencedirect.com/
science/article/pii/S0921452613007837.
[44] Y. Ashida, R. Schmidt, L. Tarruell, and
E. Demler. Many-body interferometry of
magnetic polaron dynamics. Phys. Rev. B,
97:060302, Feb 2018. DOI: 10.1103/Phys-
RevB.97.060302. URL https://link.aps.org/
doi/10.1103/PhysRevB.97.060302.
[45] A.J. Leggett. Bose-einstein condensation
in the alkali gases: Some fundamental con-
cepts. Rev. Mod. Phys., 73:307–356, Apr
2001. DOI: 10.1103/RevModPhys.73.307.
URL https://link.aps.org/doi/10.1103/
RevModPhys.73.307.
[46] K. B. Davis, M. O. Mewes, M. R. Andrews,
N. J. van Druten, D. S. Durfee, D. M. Kurn,
and W. Ketterle. Bose-einstein condensation
in a gas of sodium atoms. Phys. Rev. Lett.,
75:3969–3973, Nov 1995. DOI: 10.1103/Phys-
RevLett.75.3969. URL https://link.aps.org/
doi/10.1103/PhysRevLett.75.3969.
[47] C. J. Myatt, E. A. Burt, R. W. Ghrist,
E. A. Cornell, and C. E. Wieman. Produc-
tion of two overlapping bose-einstein conden-
sates by sympathetic cooling. Phys. Rev. Lett.,
78:586–589, Jan 1997. DOI: 10.1103/Phys-
RevLett.78.586. URL https://link.aps.org/
doi/10.1103/PhysRevLett.78.586.
[48] D. M. Stamper-Kurn, M. R. Andrews, A. P.
Chikkatur, S. Inouye, H.-J. Miesner, J. Stenger,
and W. Ketterle. Optical confinement of a
bose-einstein condensate. Phys. Rev. Lett., 80:
2027–2030, Mar 1998. DOI: 10.1103/Phys-
RevLett.80.2027. URL https://link.aps.org/
doi/10.1103/PhysRevLett.80.2027.
[49] H.-J. Miesner, D. M. Stamper-Kurn, J. Stenger,
S. Inouye, A. P. Chikkatur, and W. Ketterle.
Observation of metastable states in spinor bose-
einstein condensates. Phys. Rev. Lett., 82:
2228–2231, Mar 1999. DOI: 10.1103/Phys-
RevLett.82.2228. URL https://link.aps.org/
doi/10.1103/PhysRevLett.82.2228.
[50] J. Stenger, S. Inouye, D. M. Stamper-Kurn,
H.-J. Miesner, A. P. Chikkatur, and W. Ket-
terle. Spin domains in ground-state boseâĂŞe-
instein condensates. Nature, 396:345–348, 1998.
DOI: 10.1038/24567. URL https://doi.org/
10.1038/24567.
[51] M. R. Matthews, D. S. Hall, D. S. Jin, J. R.
Ensher, C. E. Wieman, E. A. Cornell, F. Dal-
fovo, C. Minniti, and S. Stringari. Dynamical
response of a bose-einstein condensate to a dis-
continuous change in internal state. Phys. Rev.
Lett., 81:243–247, Jul 1998. DOI: 10.1103/Phys-
RevLett.81.243. URL https://link.aps.org/
doi/10.1103/PhysRevLett.81.243.
Accepted in Quantum 2019-12-28, click title to verify 16
[52] D. S. Petrov, G. V. Shlyapnikov, and J. T. M.
Walraven. Regimes of quantum degeneracy
in trapped 1d gases. Phys. Rev. Lett., 85:
3745–3749, Oct 2000. DOI: 10.1103/Phys-
RevLett.85.3745. URL https://link.aps.org/
doi/10.1103/PhysRevLett.85.3745.
[53] P. Tommasini, E. J. V. de Passos, A. F. R.
de Toledo Piza, M. S. Hussein, and E. Tim-
mermans. Bogoliubov theory for mutu-
ally coherent condensates. Phys. Rev. A,
67:023606, Feb 2003. DOI: 10.1103/Phys-
RevA.67.023606. URL https://link.aps.org/
doi/10.1103/PhysRevA.67.023606.
[54] S. Lellouch, T.-L. Dao, T. Koffel, and L. Sanchez-
Palencia. Two-component bose gases with one-
body and two-body couplings. Phys. Rev.
A, 88:063646, Dec 2013. DOI: 10.1103/Phys-
RevA.88.063646. URL https://link.aps.org/
doi/10.1103/PhysRevA.88.063646.
[55] M. Abad and A. Recati. A study of coher-
ently coupled two-component bose-einstein con-
densates. The European Physical Journal D,
67(7):148, Jul 2013. ISSN 1434-6079. DOI:
10.1140/epjd/e2013-40053-2. URL https://
doi.org/10.1140/epjd/e2013-40053-2.
[56] G.-S. Paraoanu, S. Kohler, F. Sols, and A.J.
Leggett. The josephson plasmon as a bo-
goliubov quasiparticle. Journal of Physics B:
Atomic, Molecular and Optical Physics, 34(23):
4689–4696, nov 2001. DOI: 10.1088/0953-
4075/34/23/313. URL https://doi.org/10.
1088%2F0953-4075%2F34%2F23%2F313.
[57] A. Recati and F. Piazza. Breaking of
goldstone modes in a two-component bose-
einstein condensate. Phys. Rev. B, 99:
064505, Feb 2019. DOI: 10.1103/Phys-
RevB.99.064505. URL https://link.aps.org/
doi/10.1103/PhysRevB.99.064505.
[58] E. Nicklas. A new tool for miscibility control:
Linear coupling. 01 2013.
[59] S. John and T. Quang. Spontaneous emission
near the edge of a photonic band gap. Phys. Rev.
A, 50:1764–1769, Aug 1994. DOI: 10.1103/Phys-
RevA.50.1764. URL https://link.aps.org/
doi/10.1103/PhysRevA.50.1764.
[60] H.-T. Tan, W.-M. Zhang, and G.-x. Li.
Entangling two distant nanocavities via a
waveguide. Phys. Rev. A, 83:062310, Jun
2011. DOI: 10.1103/PhysRevA.83.062310.
URL https://link.aps.org/doi/10.1103/
PhysRevA.83.062310.
[61] J. Prior, I. de Vega, A.W. Chin, S.F.
Huelga, and M.B. Plenio. Quantum dy-
namics in photonic crystals. Phys. Rev. A,
87:013428, Jan 2013. DOI: 10.1103/Phys-
RevA.87.013428. URL https://link.aps.org/
doi/10.1103/PhysRevA.87.013428.
[62] A.G. Kofman, G. Kurizki, and B. Sher-
man. Spontaneous and induced atomic de-
cay in photonic band structures. Journal of
Modern Optics, 41(2):353–384, 1994. DOI:
10.1080/09500349414550381. URL https://
doi.org/10.1080/09500349414550381.
[63] V. P Bykov. Spontaneous emission in a periodic
structure. Journal of Experimental and Theoret-
ical Physics, 35:269, 01 1972.
[64] E. Yablonovitch. Inhibited spontaneous
emission in solid-state physics and electron-
ics. Phys. Rev. Lett., 58:2059–2062, May
1987. DOI: 10.1103/PhysRevLett.58.2059.
URL https://link.aps.org/doi/10.1103/
PhysRevLett.58.2059.
[65] P. Lambropoulos, G.M. Nikolopoulos, T.R.
Nielsen, and S. Bay. Fundamental quantum
optics in structured reservoirs. Reports on
Progress in Physics, 63(4):455–503, mar 2000.
DOI: 10.1088/0034-4885/63/4/201. URL
https://doi.org/10.1088%2F0034-4885%
2F63%2F4%2F201.
[66] M. Woldeyohannes and S. John. Coher-
ent control of spontaneous emission near a
photonic band edge. Journal of Optics
B: Quantum and Semiclassical Optics, 5(2):
R43–R82, feb 2003. DOI: 10.1088/1464-
4266/5/2/201. URL https://doi.org/10.
1088%2F1464-4266%2F5%2F2%2F201.
[67] T. Quang, M. Woldeyohannes, S. John, and
G.S. Agarwal. Coherent control of spontaneous
emission near a photonic band edge: A single-
atom optical memory device. Phys. Rev. Lett.,
79:5238–5241, Dec 1997. DOI: 10.1103/Phys-
RevLett.79.5238. URL https://link.aps.org/
doi/10.1103/PhysRevLett.79.5238.
[68] A. G. Kofman and G. Kurizki. Unified the-
ory of dynamically suppressed qubit decoher-
ence in thermal baths. Phys. Rev. Lett.,
93:130406, Sep 2004. DOI: 10.1103/Phys-
RevLett.93.130406. URL https://link.aps.
org/doi/10.1103/PhysRevLett.93.130406.
[69] H.P. Breuer and F. Petruccione. The Theory of
Open Quantum Systems. OUP Oxford, 2007.
ISBN 9780199213900. URL https://books.
google.es/books?id=DkcJPwAACAAJ.
[70] A. Rivas, A. Douglas K. Plato, S.F. Huelga,
and M.B. Plenio. Markovian master equations:
a critical study. New Journal of Physics, 12
(11):113032, nov 2010. DOI: 10.1088/1367-
2630/12/11/113032. URL https://doi.org/
10.1088%2F1367-2630%2F12%2F11%2F113032.
[71] I. de Vega, D. Alonso, and P. Gaspard. Two-
level system immersed in a photonic band-
gap material: A non-markovian stochastic
schrödinger-equation approach. Phys. Rev. A,
71:023812, Feb 2005. DOI: 10.1103/Phys-
RevA.71.023812. URL https://link.aps.org/
doi/10.1103/PhysRevA.71.023812.
Accepted in Quantum 2019-12-28, click title to verify 17
[72] I. de Vega, D. Porras, and I.J. Cirac. Matter-
wave emission in optical lattices: Single par-
ticle and collective effects. Phys. Rev. Lett.,
101:260404, Dec 2008. DOI: 10.1103/Phys-
RevLett.101.260404. URL https://link.aps.
org/doi/10.1103/PhysRevLett.101.260404.
[73] R. Vasile, F. Galve, and R. Zambrini. Spec-
tral origin of non-markovian open-system dy-
namics: A finite harmonic model with-
out approximations. Phys. Rev. A, 89:
022109, Feb 2014. DOI: 10.1103/Phys-
RevA.89.022109. URL https://link.aps.org/
doi/10.1103/PhysRevA.89.022109.
[74] W.-M. Zhang, P.-Y. Lo, H.-N. Xiong, M. W.-
Y. Tu, and F. Nori. General non-markovian dy-
namics of open quantum systems. Phys. Rev.
Lett., 109:170402, Oct 2012. DOI: 10.1103/Phys-
RevLett.109.170402. URL https://link.aps.
org/doi/10.1103/PhysRevLett.109.170402.
[75] F. Giraldi and F. Petruccione. Fractional re-
laxations in photonic crystals. Journal of
Physics A: Mathematical and Theoretical, 47
(39):395304, sep 2014. DOI: 10.1088/1751-
8113/47/39/395304. URL https://doi.org/
10.1088%2F1751-8113%2F47%2F39%2F395304.
[76] M. Bruderer, A. Klein, S.R. Clark, and
D. Jaksch. Polaron physics in optical
lattices. Phys. Rev. A, 76:011605, Jul
2007. DOI: 10.1103/PhysRevA.76.011605.
URL https://link.aps.org/doi/10.1103/
PhysRevA.76.011605.
[77] S. Patrick Rath and R. Schmidt. Field-
theoretical study of the bose polaron. Phys. Rev.
A, 88:053632, Nov 2013. DOI: 10.1103/Phys-
RevA.88.053632. URL https://link.aps.org/
doi/10.1103/PhysRevA.88.053632.
[78] R.S. Christensen, J. Levinsen, and G.M. Bruun.
Quasiparticle properties of a mobile impurity in
a bose-einstein condensate. Phys. Rev. Lett.,
115:160401, Oct 2015. DOI: 10.1103/Phys-
RevLett.115.160401. URL https://link.aps.
org/doi/10.1103/PhysRevLett.115.160401.
[79] Y.E. Shchadilova, R. Schmidt, F. Grusdt,
and E. Demler. Quantum dynamics of ul-
tracold bose polarons. Phys. Rev. Lett.,
117:113002, Sep 2016. DOI: 10.1103/Phys-
RevLett.117.113002. URL https://link.aps.
org/doi/10.1103/PhysRevLett.117.113002.
[80] Q. Wang and H. Zhan. On different nu-
merical inverse laplace methods for solute
transport problems. Advances in Water Re-
sources, 75:80 92, 2015. ISSN 0309-1708. DOI:
https://doi.org/10.1016/j.advwatres.2014.11.001.
URL http://www.sciencedirect.com/
science/article/pii/S0309170814002152.
[81] P.-Y. Lo, H.-N. Xiong, and W.-M. Zhang. Break-
down of bose-einstein distribution in photonic
crystals. Scientific Reports, 5, 2015. DOI:
10.1038/srep09423. URL https://doi.org/10.
1038/srep09423.
[82] J. Spiechowicz, J. ÅĄuczka, and P. HÃďnggi.
Transient anomalous diffusion in periodic sys-
tems: ergodicity, symmetry breaking and veloc-
ity relaxation. Scientific Reports, 6, 2016. DOI:
10.1038/srep30948. URL https://doi.org/10.
1038/srep30948.
[83] C. Navarrete-Benlloch, I. de Vega, D. Por-
ras, and J.I. Cirac. Simulating quantum-
optical phenomena with cold atoms in op-
tical lattices. New Journal of Physics, 13
(2):023024, feb 2011. DOI: 10.1088/1367-
2630/13/2/023024. URL https://doi.org/10.
1088%2F1367-2630%2F13%2F2%2F023024.
[84] M. Mehboudi, A. Lampo, C. Charalam-
bous, L.A. Correa, M.Á. García-March, and
M. Lewenstein. Using polarons for sub-
nk quantum nondemolition thermometry in a
bose-einstein condensate. Phys. Rev. Lett.,
122:030403, Jan 2019. DOI: 10.1103/Phys-
RevLett.122.030403. URL https://link.aps.
org/doi/10.1103/PhysRevLett.122.030403.
[85] D. S. Petrov. Quantum mechanical stabilization
of a collapsing bose-bose mixture. Phys. Rev.
Lett., 115:155302, Oct 2015. DOI: 10.1103/Phys-
RevLett.115.155302. URL https://link.aps.
org/doi/10.1103/PhysRevLett.115.155302.
[86] C. R. Cabrera, L. Tanzi, J. Sanz, B. Naylor,
P. Thomas, P. Cheiney, and L. Tarruell. Quan-
tum liquid droplets in a mixture of bose-einstein
condensates. Science, 359(6373):301–304, 2018.
ISSN 0036-8075. DOI: 10.1126/science.aao5686.
URL https://science.sciencemag.org/
content/359/6373/301.
Accepted in Quantum 2019-12-28, click title to verify 18