Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

Interaction of One-Dimensional Quantum Droplets with Potential Wells and Barriers

Argha Debnath,1 Ayan Khan,1, ∗ and Boris Malomed2, 3


1
Department of Physics, School of Engineering and Applied Sciences,
Bennett University, Greater Noida, UP-201310, India
2
Department of Physical Electronics, School of Electrical Engineering,
Faculty of Engineering, Tel Aviv University, Ramat Aviv 69978, Israel
3
Instituto de Alta Investigación, Universidad de Tarapacá, Casilla 7D, Arica, Chile
We address static and dynamical properties of one-dimensional (1D) quantum droplets (QDs)
under the action of local potentials in the form of narrow wells and barriers. The QDs are governed
arXiv:2302.13367v1 [cond-mat.quant-gas] 26 Feb 2023

by the 1D Gross-Pitaevskii equation including the mean-field cubic repulsive term and the beyond-
mean-field attractive quadratic one. In the case of the well represented by the delta-functional
potential, three exact stable solutions are found for localized states pinned to the well. The Thomas-
Fermi approximation for the well and the adiabatic approximation for the collision of the QD
with the barrier are developed too. Collisions of incident QDs with the wells and barriers are
analyzed in detail by means of systematic simulations. Outcomes, such as fission of the moving QD
into transmitted, reflected, and trapped fragments, are identified in relevant parameter planes. In
particular, a counter-intuitive effect of partial or full rebound of the incident QD from the potential
well is studied in detail and qualitatively explained.

I. INTRODUCTION moving QDs, which was realized experimentally (in the


3D geometry) in Ref. [6]. In the 1D setting, simulations
Quantum droplets (QDs) in binary Bose-Einstein of the corresponding GPE with the cubic-quadratic non-
condensates (BECs) originate from the interplay of linearity demonstrate that the collision between slowly
the mean-field (MF) interactions and Lee-Huang-Yang moving droplets with zero phase difference between them
(LHY) corrections to them, induced by quantum fluc- leads to their merger into a single one, in an excited
tuations [1]. As first demonstrated by Petrov [2], this (breathing) state, while the collision between QDs with
setting is modeled by the modified Gross-Pitaevskii equa- phase difference π ends up with quasi-elastic rebound
tion (GPE) which includes the MF self-attractive cubic [18]. Collisions with large velocities tend to exhibit quasi-
term and the repulsive LHY quartic one. This predic- elastic outcomes, i.e., passage of the QDs through each
tion was followed by the creation of QDs in binary ho- other, similar to collisions between usual matter-wave
moatomic [3–6] and heteroatomic [7] BECs, as well as solitons [20].
in dipolar condensates [8, 9]. Actually, QDs represent a Another physically relevant problem is the consider-
new quantum state of matter, see review [10, 11]. ation of collisions of an incident QD with a local at-
The dimensional reduction 3D → 2D and 3D → 1D tractive or repulsive defect, which may be represented
for BEC under the action of a tight confining potential by a localized potential well or barrier, respectively. In
essentially changes the form of the effective GPE [12–15]. the experiment, such defects can be created by narrow
In contrast with the 3D setting, in the 1D limit the LHY red- or blue-detuned laser beams illuminating the con-
term has the self-attraction sign, and is quadratic, rather densate [21]. For moving solitons this problem was con-
than quartic [12]. However, it is also possible to extend sidered, in various forms, in the framework of the nonlin-
the 3D framework in quasi-one-dimensional (Q1D) ge- ear Schrödinger equation (NLSE) [22–29]. In BEC, colli-
ometry by tuning the transverse confinement suitably. sions of matter-wave solitons with barriers were studied
In this situation, the repulsive quartic LHY contribution both theoretically [30–33] and experimentally [21]. In
prevails to support the droplet formation in both homo- particular, much interest was drawn to the use of nar-
geneous and inhomogeneous systems [16, 17]. row potential barriers as splitters in the design of soliton
On the contrary, in the 1D system the residual self- interferometers [34–39].
repulsive MF term competes with the quadratic self- In this work, we aim to study scattering of QDs on
attraction, making it possible to create broad QD states potential wells and barriers in the framework of the
with the equilibrium density in their inner quasi-flat parts one-dimensional LHY-corrected GPE with the cubic-
[18, 19]. The fact that the density cannot exceed the equi- quadratic nonlinearity. The corresponding model is for-
librium value implies that the quantum matter filling the mulated in Section II. Analytical results obtained for
QDs is an incompressible liquid, therefore localized states the model are collected in Section III. In particular, we
are named “droplets”. consider quiescent QDs trapped in potential wells. For
A natural extension to understand the droplets is to the well represented by the delta-functional potential we
study its dynamics and possibility of collisions between present three exact stable solutions for localized states
pinned to the narrow well. The Thomas-Fermi and adia-
batic approximations are considered too, for pinned and
∗ ayan.khan@bennett.edu.in moving QDs, respectively. Collisions of moving QDs with
2

the barrier or well or studied by means of systematic


simulations in Section IV. The outcome of the collisions
is quantified by computing shares of the initial norm in
the transmitted, reflected and trapped waves (trapping
takes place in the case of the potential well). We also
plot maps, in the plane of the norm and velocity of the
incident QD, for values of the maximum strength of the
potential well admitting full transmission, and minimum
strength required for complete reflection. The paper is
concluded by Section V.

II. THE MODEL FIG. 1. Spatial density distributions, |ψ(x)|2 , as given by the
QD solution (7), for different values of norm (9), which are
indicated in the figure.
In this section we plan to summarize the theoretical
model in 1D. We start from the binary BEC with equal
self-repulsion coefficients in both components, g11 = Stationary solutions for QDs with chemical potential
g22 = g > 0, and equal numbers of atoms in them. This µ < 0, produced by Eq. (3), are looked for in the usual
setting makes it possible to consider the symmetric con- form,
figuration, with equal MF wave functions of the compo-
nents, ψ1 = ψ2 ≡ ψ. Then, defining a small difference be- ψ (x, t) = exp (−iµt) φ(x) , (5)
tween the inter-component attraction strength (g12 < 0)
where real stationary wave function φ(x) > 0 satisfies the
and intra-component self-repulsion, δg = g + g12 > 0,
equation
with δg  g, one can derive the effective 1D GPE in-
cluding the beyond-MF correction (the quadratic term) 1 d2 φ
as [12]: µφ = − + f (x)φ + φ3 − φ2 . (6)
2 dx2
√ In particular, the free-space version of Eq. (6), with
2 2
∂ψ h̄ ∂ ψ 2m 3/2 f (x) = 0, gives rise to the known family of exact QD
ih̄ =− + f (x)ψ + δg|ψ|2 ψ − g |ψ|ψ, solutions [12],
∂t 2m ∂x2 πh̄
(1)
where f (x) represents an external potential, and m is the 3µ
atomic mass. φ(x) = − p √ , (7)
Equation (1) determines characteristic units of length, 1+ 1 + 9µ/2 cosh −2µx
x0 , time, t0 , energy, E0 , and wave function, where µ < 0 is the chemical potential, which takes values
in a finite bandgap,

πh̄2 δg π 2 h̄3 δg 0 < −µ < 2/9. (8)
x0 = √ , t0 = ,
2mg 3/2 2mg 3
√ The norm of the soliton (7) is [12]
h̄2 h̄ 2m 3/2
E0 = 2 = , ψ0 = g , (2)
mx0 t0 πh̄δg +∞
" p ! #
3 −µ/2 + 1
Z
2 4 p
N= |ψ(x)| dx = ln p − 3 −µ/2 .
which suggests rescaling t ≡ t0 t̃, x ≡ x0 x̃, ψ ≡ ψ0 ψ̃, −∞ 3 1 + 9µ/2
f 0 (x) ≡ f˜(x̃)/E0 . In this notation, Eq. (1) is cast in the (9)
normalized form (where the tildes are omitted): Note that the dependence N (µ), as given by Eq. (9), sat-
isfies the celebrated Vakhitov-Kolokolov (VK) criterion,
∂ψ 1 ∂2ψ dN/dµ < 0, (10)
i =− + f (x)ψ + |ψ|2 ψ − |ψ|ψ, (3)
∂t 2 ∂x2
which is a necessary condition for the stability of localized
with Hamiltonian (energy), states supported by any self-attractive nonlinearity [40–
Z +∞ " 2 42].
1 ∂ψ In the limit of µ → −0, the QD state (7) is similar to
E= + f (x)|ψ(x)|2
−∞ 2 ∂x traditional solitons dominated by the quadratic nonlin-
2 1
 earity, such as ones produced by the Korteweg – de Vries
3 4
− |ψ(x)| + |ψ(x)| dx ≡ Ekin + Eint , (4) equation [43, 44],
3 2

which includes the kinetic, alias gradient (first), and in- φ(x) ≈ − p . (11)
2
teraction (potential) terms. cosh −µ/2x
3

In the opposite limit, µ → −2/9 [see Eq. (8)], Eq. (7)


produces quasi-flat droplets (“puddles”, in terms of Ref.
[18]), with the nearly constant inner density,

φ2 (x) ≤ nmax = 4/9, (12)

and a logarithmically large width,

L ≈ (3/2) ln [1/ (1 + 9µ/2)] . (13)

Exactly at the border of the bandgap (8), µ = −2/9, QD


(7) carries over into the front solutions,

2/3
φfront (x; µ = −2/9) = , (14)
1 + exp (±2x/3)

which connect the zero state and the constant-amplitude


(alias continuous-wave, CW) one,

ψCW = (2/3) exp((2/9)it). (15)

In fact, the front solutions (14) represent boundaries of


a very broad flat-top QD.
Density profiles of the QD solution (7) with different
values of N are plotted in Fig. 1(a). In the figure, the
profiles corresponding to N = 20, 30, and 40 clearly ex-
hibit the flat-top shape. In agreement with the prediction
of the VK criterion, all the QDs (7) are stable solutions
of Eq. (3).
The Galilean invariance of Eq. (3) [with f (x) = 0]
makes it possible to set the QD in motion by application
of the boost (kick) to them with arbitrary parameter k,
which determines the velocity of the moving mode:

ψ (x, t) = exp ikx − i k 2 /2 t ψ (x − kt, t) .


  
(16)

The effective mass of the moving QD coincides with its


norm N , i.e., its kinetic energy is

Ekin = N k 2 /2 .

(17)

This result is used below to predict the threshold of the FIG. 2. Comparison between typical examples of analytical
rebound of the moving QD from a potential barrier, see solutions (blue solid lines), given, respectively, by Eqs. (21),
Eq. (40). (22) and (29), (30), and their counterparts (red bar lines) pro-
duced by the numerical solution of Eq. (6) [with the Gaussian
potential (20) approximation the delta-function] is depicted in
III. ANALYTICAL RESULTS (a) and (b). The respective values of ε in (a) and (b) are 0.2
and 0.3. It is seen that the analytical and numerical solutions
are indistinguishable. (c) The continuous dependence Nε (µ),
A. Localized modes pinned to narrow potential
as produced by the analytical solutions given, severally, by
wells: exact solutions
Eqs. (21) and (22) at µ > −2/9, and by Eqs. (29) and (30)
at µ < −2/9, for ε = 0.2, 0.3 and 0.4 (red, green, and orange
The discussion on the homogeneous solution allows us lines, respectively). The vertical blue dashed line indicates the
to move forward and consider a situation where an ex- boundary between the two analytical solutions, µ = −2/9. At
ternal potential (f (x) 6= 0). Our primary objective is to µ < −2/9, the curves are extended very close to the right edge
understand what happens to a QD when it encounters of interval (31), −µ = 2/9 + ε2 /2, beyond which solution (29)
responds to a potential well/barrier as described in Fig. does not exist. Counterparts of these curves produced by the
numerical solution (not shown here) are virtually identical to
3.
the analytical ones.
However, if the width of the potential well is much
smaller than the size of the QD (which naturally happens
4

which is reduced in comparison to the bandgap men-


tioned in Eq. (8). Thus it follows from Eq. (23) that the
pinned states of the QDs exist only if the delta-functional
pinning potential is not too strong, viz.,

ε < εmax = 2/3. (24)

A majority of numerical results are reported below for


ε ≡ −V0 < 2/3, i.e., condition (24) holds for them.
In the limit case of µ = −2/9, the exact solution for the
pinned QD, given by Eqs. (21) and (22) takes a simple
but nontrivial form:
FIG. 3. A schematic representation of a flat-top QD colliding
with a rectangular potential well with width 2a and depth V0 . 2/3
φε (x, µ = −2/9) = −1 . (25)
1 + ε (2/3 − ε) exp (2|x|/3)
in the case if the flat-top QDs), potential (41) may be Note that, while the exact solution in Eq. (7) with µ =
naturally approximated by the delta-function, −2/9 degenerates into the CW state as noted in Eq. (15)
with the divergent norm, solution (25) with the same µ
f (x) = fε (x) = −εδ(x), (18) remains a localized one, with a final norm [defined as in
Eq. (9)]
with strength
   
4 2 2
ε ≡ −2aV0 . (19) Nε (µ = −2/9) = ln −2 −ε . (26)
3 3ε 3
The same potential (18) with ε < 0 represents a nar-
row potential barrier, such as the one used as the soliton In the opposite limit corresponding to the left edge of
splitter in interferometers [34–38]. interval (23),
The exact analytical solutions for nonlinear modes
µ → µ0 ≡ −ε2 /2, (27)
pinned to the delta-functional potential well described in
Eq. (18), are produced below. We also substantiate these
the solution given by Eqs. (21) and (22) degenerates into
analytical results numerically as described in Fig. 2(a,b).
The computational counterparts are provided by the nu- φε (x; µ = −ε2 /2) = A exp (−ε|x|) , (28)
merical solution of Eq. (6) where the delta-functional is
approximated by means of the standard Gaussian, where A is an arbitrary infinitesimal amplitude. This
 2 limit form of the solution is tantamount to the commonly
1 x
δ̃(x) = √ exp − 2 , (20) known bound state maintained by the delta-functional
πσ σ potential well in the framework of the linear Schrödinger
equation.
with small width such that σ = 0.03 [45].
A typical example of the exact solution produced by
The exact solutions of Eq. (6) for QDs pinned to the
Eqs. (21) and (22) is displayed in Fig. 2(a), where it is
delta-functional potential well (18). First, in the interval
compared with its counterpart produced by a numerical
(8) of the values of µ, the pinned mode can be constructed
solution of Eq. (6) with the potential well taken as per
as a direct extension of its free-space counterpart, given
Eqs. (41) and (19). It is seen that the analytical and
by Eq. (7), viz.,
numerical solutions are virtually identical ones.
3µ The comparison of the value of the norm given by Eq.
φε (x, µ) = − p √ , (26) and the infinitesimal norm at µ = −ε2 /2 correspond-
1+ 1 + 9µ/2 cosh −2µ (|x| + ξ) ing to Eq. (28) suggests that the dependence Nε (µ) is
(21) monotonously decreasing. This feature, which is corrob-
where the positive “lost length” is orated by Fig. 2(c), suggests that the family of pinned
r "p # QDs is stable, according to the above-mentioned VK cri-
2 ε2 + (1 + 9µ/2) (−2µ − ε2 ) + ε terion (10). [40–42]. Their stability is indeed corrobo-
2ξ = − ln p √  .
µ 1 + 9µ/2 −2µ − ε rated by direct simulations of Eq. (3) (not shown here in
(22) detail, as results of the simulations are straightforward).
It is seen from Eq. (22) that the exact solutions for the In the presence of the attractive δ-functional potential,
pinned states exists in the following interval of values of it is possible to find another exact solution for pinned
the chemical potential, QDs in the region of µ < −2/9, where free-space solitons
(7) do not exist. In the case when ε satisfies the condition
ε2 /2 < −µ < 2/9, (23) (24), this region extends the existence interval (8). The
5

corresponding exact solution, which has no counterparts the norm decreases from a finite value, corresponding to
in the free space, is solution (34), at µ = −ε2 /2 to N = 0 at µ = −2/9. The
compliance of the present family with the VK criterion
3µ implies its stability.
φ̃ε (x) = − p √ 
1 + −(1 + 9µ/2) sinh −2µ (|x| + η)
(29)
[cf. expression (7) for the solution constructed above], B. The Thomas-Fermi (TF) approximation for
the positive “lost length” is pinned state
r "p #
2 ε2 − (1 + 9µ/2) (ε2 + 2µ) + ε A specific analytical approximation for solutions to Eq.
2η = − ln p √  , (30)
µ − (1 + 9µ/2) −2µ − ε (6) with the deep potential well (41), i.e., with large −V0
and µ < 0, is provided by the TF approximation, which
cf. Eq. (22). Expression (29) is relevant if it yields real neglects the derivative (kinetic-energy) term:
values of η. This condition holds in the following finite 
interval of values of the chemical potential,
q
 1 + 1 − (V0 − µ), at |x| < a,
2 4q
φTF (x) = (35)
2/9 ≤ −µ ≤ 2/9 + ε2 /2. (31) 1 1
+ 4 + µ, at |x| > a

2
It is a straightforward annex to interval (23) in which the
above solution (21) exists. As well as the latter solution, [in fact, the solution (35) at |x| > a is relevant for the
the one given by Eqs. (29) and (30) satisfies the VK flat-top QD, with µ ≈ −2/9, the nearly constant solution
stability criterion (10), as shown in Fig. 2(c). Note that, around the potential well being close to the CW back-
in the limit of µ + 2/9 → −0, solution (29) takes the ground (15)]. Then, for a small perturbation δφ added
same form (25) to which the above solution (21) amounts to φTF (x), the linearization of the nonlinear term in Eq.
at µ + 2/9 → +0, hence solution (29) provides smooth (6) produces, in the lowest approximation with respect
continuation of (21) across the point of µ + 2/9 = 0 [see to the fact that −V0 > 0 is a large parameter, an effec-
an illustration in Fig. 2(c)]. tive term in the GPE, +3εδ(x) · δφ, where ε is defined as
The exact solution (29) can also be constructed in the per Eq. (19), and δ(x) is introduced in the same approx-
region of imation as in Eq. (18). Then, the combination of this
term and one corresponding to Eq. (18) gives rise to an
ε > 2/3, (32) effective repulsive potential,

i.e., outside of the region (24). In this case, expression feffective (x) = +2ε(x). (36)
(30) is replaced by
" p # This crude analysis predicts, in a quantitative form, that
1 ε − ε2 − (1 + 9µ/2) (ε2 + 2µ) the collision of a flat-top QD with the deep potential well
η=√ ln p √ , (33) may lead, instead of the naturally expected passage, to
−2µ

− (1 + 9µ/2) −2µ − ε
the counterintuitive outcome in the form of rebound. This
and the solution exists in the same interval of the chem- possibility is indeed demonstrated by numerical results
ical potential as given by Eq. (31), while it has no coun- reported below. It is relevant to mention that reflection
terpart at −µ < 2/9. Indeed, in the limit of µ+2/9 → −0 of incident wave packets from a potential well is a well-
Eqs. (29) and (33) demonstrate that the solution degen- known quantum effect, which was also elaborated in the
erates into φ = 0 [on the contrary to the nonzero solution framework of the MF theory [24, 25].
(25), into which solution (21) carries over in the limit of
µ + 2/9 → +0], thus providing the natural continuity
with the absence of the pinned solution at µ > −2/9 in C. The interaction of the moving QD with a
the region (32). On the other hand, point (27), where narrow potential barrier
solution (21) vanishes, as shown above [see Eq. (28)], is
now an internal point of interval (31), due to condition As mentioned above, collisions of moving matter-wave
(32). At this point, the present solution takes a relatively packets with narrow potential barriers is a problem of
simple form, high relevance for the design of soliton interferometers
[34–38]. In the present context, treating the repulsive
φε (x; µ = −ε2 /2)
delta-functional potential (18) as a perturbation [46]
3ε2 /2 makes it possible to predict the threshold value kthr of ve-
= p .
1 + 9ε2 /4 − 1 sinh [ε|x| + (1/2) ln (9ε2 /4 − 1)] locity k [see Eq. (16)] which separates the rebound and
(34) passage of the incoming QD. In the adiabatic approxi-
mation, which neglects deformation of the QD under the
The family of the exact solutions produced by Eqs. action of the barrier, an effective potential of the inter-
(29) and (33) also satisfies the VK criterion; in particular, action of the QD with the barrier can be found, in the
6

the QD due to its interaction with the local potential) as


Z +∞
2
Ueff (Ξ) = −ε δ(x) |ψ (x − Ξ)| dx
−∞
9εµ2
= −h i2 ,
p √
1 + 1 + 9µ/2 cosh −2µΞ
(37)
where Ξ(t) is the coordinate of the center of the mov-
ing QD. Obviously, the height of the effective potential
barrier (37) is

9εµ2
Umax =  p 2 , (38)
1 + 1 + 9µ/2
FIG. 4. The plot of the spatiotemporal evolution of the den-
sity, |ψ (x, t)|2 , presents an example of the full passage of the hence the comparison of this expression with the QD’s
incident QD through the potential well (41), for parameters kinetic energy (17) predicts the threshold value as
indicated in the figure.
2 18εµ2
kthr =  p 2 , (39)
N (µ) 1 + 1 + 9µ/2

where N (µ) is to be taken as per Eq. (9). The incident


QD with k < kthr or k > kthr is expected, respectively,
to bounce back or pass over the barrier.
In the limit case of flat-top QDs, which correspond to
µ close to −2/9 [see Eq. (8)], Eqs. (38) and (39) take a
simplified form:
4 2 8ε
Umax ≈ ε, kthr ≈ . (40)
9 9N
It may also be relevant to consider shuttle motion of the
QD in a cavity formed by two narrow barriers, sepa-
rated by a large distance L. The period of the motion
with velocity k is T = 2L/k, hence the largest shut-
tle frequency which can be maintained by the cavity is
ωmax = πkthr /L.

IV. NUMERICAL ANALYSIS

In this section we aim to study the dynamics of QDs


while subjected to a potential well/barrier. For this pur-
pose we perform the numerical simulation of Eq. (3) us-
ing the split-step method based on the fast-Fourier trans-
form. In the numerical simulations, the local potential in
Eq. (3) was taken as a rectangular one as described in
FIG. 5. (a) An example of the partial transmission of the
incident QD through the potential well (41) with depth V0 =
Fig.3,
−0.14. A very small share of the wave function is trapped 
in the well. (b) The temporal evolution of the transmitted,
V0 , for − a < x < +a,
f (x) = (41)
trapped and reflected shares, defined as per Eqs. (43). 0, for |x| > a.

The potential landscape can be both well or barrier based


on the sign of V0 . V0 < 0 describes potential well while
V0 > 0 is a barrier.
The numerical results are reported for the width of
adiabatic approximation (which neglects deformation of the potential well a = 0.5, which is small in comparison
7

with other length scales, such as the width of the incident


QDs, and justifies the use of the approximation (18). In
this case, Eq. (19) yields ε = −V0 . Here, we apply zero
boundary condition corresponding to an infinitely deep
potential box while occupying a broad spatial domain,
|x| < L = 200. The initial wave function was taken as

ψ(x, t = 0) = exp(ikx)φ(x − xinit ), (42)

where φ(x) is the stationary solution (7) for the QD, k


is the kick which sets the QD in motion, and xinit is its
initial position.

A. Moving droplets

It is natural to expect that the travelling droplets can


manifest different characteristics while subjected to an
inhomogeneous potential landscape. These droplets can
be transmitted or reflected or even can be trapped based
on the initial kick (k) or the strength of the potential.
The transmitted, trapped, and reflected components
of the wave function, can be noted as,
Z L
Ntrans = dx|ψ(x, t)|2 ,
a FIG. 6. (a) An example of the partial transmission and partial
Z +a
reflection of the incident QD colliding with the potential well
Ntrap = dx|ψ(x, t)|2 , (41), whose depth is V0 = −0.248. (b) The temporal evolution
−a
Z−a
of the transmitted, trapped and reflected norms as defined in
Nref = dx|ψ(x, t)|2 , (43) Eq. (43).
−L

which are subject to the obvious constraint, Ntrans +


Ntrap + Nref = N throughout the dynamics. in qualitative agreement with the above prediction, see
At this stage we check the effect of the potential Eq. (36). The obvious fact that the trapped share re-
strength with fixed k and N . However we will report the mains very small, as seen, e.g., in Fig. 6 at V0 = −0.248
role of the initial velocity and the norm in the later stages. and in other figures displayed above, has a simple ex-
We start our numerical calculation with xinit = 65, planation. Indeed, the trapped mode maintained by the
moving with velocity k towards the local potential bar- narrow potential well is narrow itself in comparison to
rier/well defined by Eq. (41) with the width scaled to be the incident flat-top QD, its norm also being relatively
a = 0.5. small. A characteristic example of this feature is pre-
At the beginning, we present characteristic results ob- sented by solution (25) with norm (26) for µ = −2/9,
tained for the total norm N = 40 and velocity k = 0.1. while the QD’s norm diverges at µ = −2/9.
This relatively low collision velocity is appropriate for Further increase of the depth of the potential well leads
producing nontrivial results, while the fast moving QD to a regime in which the incident QD splits in three dif-
simply passes the barrier/well, or elastically bounces ferent parts, viz., the transmitted, reflected, and (small)
back from a very tall barrier. Conclusions valid for trapped wave packets. Figures 7 and 9 display examples
generic values of the N and k are formulated below. of this outcome of the collisions for depths V0 = −0.30
Figure 4 presents an example of the total transmission and V0 = −0.31, respectively. The former case is addi-
of the QD through the potential well (41) with depth tionally illustrated in Fig. 8 by a set of snapshots of the
V0 = −0.13. Yellow lines in the density plot indicate density profiles. It is seen that all the three packets are
excitations inside of the QD generated by the collision produced by the collision is in excited states, and a small
with the well (intrinsic excitation modes in quasi-1D QDs change of V0 , from −0.30 to −0.31, leads to a conspicu-
were studied in Ref. [19]). ous change of the dynamical picture. In terms of energy,
A trapped fraction of the wave field appears for a larger these collisions are considered in the next subsection.
depth of the potential well. As an example, Fig. 5 shows In accordance with the above qualitative analysis,
the transmission with partial capture of the wave func- based on the TF approximation and Eq. (36), nearly
tion at V0 = −0.14. Further increase of −V0 causes par- full reflection of the incident QD is observed at larger
tial reflection of the QD from the potential well, which is values of the well’s depth −V0 , as illustrated in Figs. 10
8

FIG. 8. Snapshots of the QD colliding with the potential


FIG. 7. (a) The splitting of the QD impinging onto poten- well of depth V0 = −0.3 at (a) t = 100, (b) t = 300, (c)
tial well (41), with depth V0 = −0.3, into the transmitted, t = 800. Over the time the incident QD splits into a larger
trapped and reflected parts are depicted. (b) The time evo- reflected part and a smaller transmitted one, along with a
lution of the norms of the transmitted, trapped and reflected narrow trapped mode [as described in (c)].
waves are plotted. (c) The figure magnifies the the evolution
of the norm corresponding to the trapped wave [described by
the orange line in (b)]. (recall it corresponds to the collision of the QD with the
potential well (41) of depth −V0 = 0.3. The results will
help to understand the situation in the general case as
for V0 = −0.55. A very small trapped component is ob- well.
served too, while transmission ceases to exist. In Fig. 11 we have plotted the variation of kinetic
Besides the collision, splitting of the 1D QD into two and interaction energies in this case, which naturally
or several fragments can also be caused by imprinting demonstrates, for the flat-top QD, the domination of
sufficiently small intrinsic excitation onto it [18]. the negative potential energy, which saturates around
−8.6, over the kinetic energy, which takes values around
0.20. Further, Fig. 12 demonstrates that kinetic energies
B. Analysis of energy in the trapping process of the transmitted part and reflected wave packets are
nearly equal, while their velocities and masses (norms)
To consider the role of energy in the trapping dynam- are widely different, as seen in Fig. 7. Simultaneously,
ics for the attractive potential, we use Eq. (4) for the the kinetic energy of the trapped packet oscillates around
system’s Hamiltonian, and address the case when the in- 0.006, indicating the presence of low-frequency excita-
cident QD splits in three fragments, as shown in Figs. 7 tions in it. On the other hand, the absolute values of the
9

FIG. 9. (a) The partial reflection at V0 = −0.31, with a


very small packet trapped in the well. (b) The variation of
the transmitted, trapped and reflected norms are noted which
clearly indicates a predominantly reflection of the waves.

negative interaction (potential) energy in the reflected


wave function is larger by a factor ' 3 than in the trans-
mitted one, due to the obvious fact that the norm of the
reflected packet is essentially larger in Fig. 7.
Due to the small values of the kinetic and interaction
energies in the trapped wave are difficult to discern in
Figs. 11 and 12, they are displayed in detail in Fig. 13.
As said above, persistent oscillations of the energies in-
dicate that the trapped wave packet was created in the
excited state. The ratio of the absolute values of the
interaction and kinetic energies for this state is ' 30,
which is comparable to the same ratios for the reflected
and transmitted packets in Figs. 11 and 12.

C. Effects of the norm (number of particles) and


velocity on scattering of the quantum droplets
FIG. 10. (a) A reasonably deep potential well with depth
As mentioned earlier, the collision of the QD with a V0 = −0.55 leads to nearly total reflection of the QD. (b)-
sufficiently deep potential well tends to make the reflec- (d): Snapshots of the density profile at different times which
tion (rather than the naively expected transmission) a clearly substantiates the observation of (a). However, a very
dominant feature of the interaction [as predicted, in par- small fraction of trapped states can also be viewed from (a)
ticular, by Eq. (36)]. For this reason, it is interesting to and (d).
identify the largest well’s depth, −V0 , which still admits
the full transmission (passage) of the incident QD. Figure
14(a) presents a heatmap for such values of −V0 as a func-
tion of the droplet’s norm (scaled number of particles),
10

FIG. 11. The time dependence of (a) the kinetic (gradient) FIG. 13. The time dependence of (a) the kinetic (gradient)
and (b) interaction (potential) energies in the case corre- and (b) interaction (potential) energies of the trapped state
sponding to Fig. 7. The widely different vertical scales in created by the collison of the incident QD with the narrow
both the panels are noteworthy (though expected). potential well, in the case shown in Fig. 7.

N , and speed, k. It is clearly seen that a maximum value


of the so defined depth is achieved for moderate values of
the norm, around N ' 25. This fact may be understood
because the norm should be relatively large to consider
the interaction with the potential well as a relatively weak
perturbation, which cannot produce a conspicuous effect
(such as the reflection) and, on the other hand, if the
norm is too large, it will switch on the TF effect which
gives rise to the effective repulsive potential, as per Eq.
(36). As for the effect of the velocity k, it may be ex-
plained in a coarse approximation, by noting that, unless
the collision is very slow, it excites the above-mentioned
intrinsic oscillations in the trapped wave packet, which
impedes the straightforward transmission.
The above-mentioned counter-intuitive result, namely,
the full rebound of the incident QD from the deep narrow
potential well, deserves a detailed consideration too. The
result is summarized in the Fig. 14(b), which presents, in
the plane of (N, k), a heatmap of the minimum (thresh-
old) value of the well’s depth −V0 above which the full
rebound takes place. The asymptotic independence of
the threshold on N is explained by the fact that the out-
come of the collision of the incident flat-top QD with the
FIG. 12. The time dependence of (a) the kinetic (gradient) narrow potential well is determined by the interaction
and (b) interaction (potential) energies of the transmitted, of the well with the front (14) separating the zero and
trapped and reflected wave packets corresponding to Fig. 7. flat-top [near-CW, see Eq. (15)] levels of the density,
The green, blue and red lines describes the reflected, trans- irrespective of the width of the QD pattern.
mitted and trapped waves respectively. Finally, we have performed the numerical analysis of
11

FIG. 14. (a) The pseudo-color map (heatmap) of the max- FIG. 15. (a) The variation of the maximum total transmission
imum value of the potential-well’s depth −V0 which admits potential strength with the change of k and N for a potential
the full transmission (passage) of the incident QD, as a func- barrier is depicted here. (b) The figure described the corre-
tion of the QD’s norm N and velocity k. (b) The heatmap sponding variation of the minimum total reflection potential
of the minimum (threshold) value of −V0 which provides the strength with the change of k and N .
counter-intuitive outcome of the collision, viz., full rebound of
the QD from the narrow potential well.
V. CONCLUSION

Using the known one-dimensional GPE (Gross-


Pitaevskii equation) with the cubic self-repulsion and
the interaction of the incident QDs with the narrow bar- quadratic attraction induced by quantum fluctuations,
rier, which is represented by the narrow potential (41) we have studied in detail the static and dynamic solutions
with V0 > 0. The results are summarized in Fig. 15, for QDs (quantum droplets) interacting with narrow po-
in which panel (a) and (b) display, respectively, the tential wells and barriers. In the case of the potential
heatmaps, in the (N, k) plane, for the maximum value well represented by the delta-functional potential, three
of V0 which admits the full transmission of the incident different stable solutions for QDs pinned by the delta-
QD, and the minimum (threshold) one above which the functional potential are found in the exact form. The
total reflection occurs. In the above-mentioned adiabatic TF (Thomas-Fermi) approximation for the narrow rect-
approximation, which neglects deformation of the QD angular well, and the adiabatic approximation for the
colliding with the potential barrier [46], both values of V0 collision of the QD with the potential barrier are elabo-
would be equal. The strong difference between the left rated too. Collisions of moving QDs with the well or bar-
and right panels of Fig. 15 implies that the deformation, rier are systematically studied by means of direct GPE
including fission of the incident QD into the transmitted simulations. The results are reported by means of dia-
and reflected fragments, is a conspicuous effect. Never- grams which represent outcomes of the collisions, includ-
theless, Eqs. (39) and (40) make it possible to explain ing the splitting of the incident QD into transmitted and
qualitative features observed in the right panel of Fig. 15. reflected fragments, as well as a small trapped one, are
In particular, this is the growth of the threshold value of identified in the plane of the QD’s norm and velocity.
V0 with the increase of k at fixed N . In particular, a counter-intuitive outcome in the form of
12

the reflection of the QD by the potential well is identi- text of matter wave interferometry.
fied and qualitatively explained. In the general case, the
transmitted, reflected and trapped wave packets emerge
in excited states, featuring intrinsic oscillations.
As mentioned earlier, the collisional dynamics of mat-
ACKNOWLEDGMENT
ter wave solitons are well studied in literature. After the
experimental advent of the QDs it is natural curiosity to
investigate the response of these QDs against an obstacle. The work of B.A.M. was supported, in part, by the Is-
Here, we have tried to fill this void through this system- rael Science Foundation through grant No. 1695/22. AK
atic study of the QDs against potential well/barrier. We thanks Department of Science and Technology (DST), In-
expect our theoretical observations will complement the dia for the support provided through the project number
future experiments and open up new avenues in the con- CRG/2019/000108.

[1] T. D. Lee, K. Huang, C. N. Yang, Eigenvalues and eigen- (2016) 100401.


functions of a bose system of hard spheres and its low- [13] T. Ilg, J. Kumlin, L. Santos, D. S. Petrov, H. P. Büchler,
temperature properties, Physical Review 106 (6) (1957) Dimensional crossover for the beyond-mean-field correc-
1135. tion in bose gases, Physical Review A 98 (5) (2018)
[2] D. Petrov, Quantum mechanical stabilization of a 051604.
collapsing bose-bose mixture, Physical review letters [14] P. Zin, M. Pylak, M. Gajda, Revisiting a stability prob-
115 (15) (2015) 155302. lem of two-component quantum droplets, Physical Re-
[3] C. Cabrera, L. Tanzi, J. Sanz, B. Naylor, P. Thomas, view A 103 (1) (2021) 013312.
P. Cheiney, L. Tarruell, Quantum liquid droplets in a [15] M. Edmonds, T. Bland, N. Parker, Quantum droplets of
mixture of bose-einstein condensates, Science 359 (6373) quasi-one-dimensional dipolar bose–einstein condensates,
(2018) 301–304. Journal of Physics Communications 4 (12) (2020) 125008.
[4] P. Cheiney, C. Cabrera, J. Sanz, B. Naylor, L. Tanzi, [16] A. Debnath, A. Khan, Investigation of quantum droplets:
L. Tarruell, Bright soliton to quantum droplet transition An analytical approach, Ann. Phys. (Berlin) 533 (2021)
in a mixture of bose-einstein condensates, Physical re- 2000549.
view letters 120 (13) (2018) 135301. [17] A. Debnath, A. Khan, S. Basu, Dropleton-soliton
[5] G. Semeghini, G. Ferioli, L. Masi, C. Mazzinghi, L. Wol- crossover mediated via trap modulation, Physics Letters
swijk, F. Minardi, M. Modugno, G. Modugno, M. Ingus- A 439 (2022) 128137.
cio, M. Fattori, Self-bound quantum droplets of atomic [18] G. Astrakharchik, B. A. Malomed, Dynamics of one-
mixtures in free space, Physical review letters 120 (23) dimensional quantum droplets, Physical Review A 98 (1)
(2018) 235301. (2018) 013631.
[6] G. Ferioli, G. Semeghini, L. Masi, G. Giusti, G. Mod- [19] M. Tylutki, G. E. Astrakharchik, B. A. Malomed, D. S.
ugno, M. Inguscio, A. Gallemı́, A. Recati, M. Fattori, Petrov, Collective excitations of a one-dimensional quan-
Collisions of self-bound quantum droplets, Physical re- tum droplet, Physical Review A 101 (5) (2020) 051601.
view letters 122 (9) (2019) 090401. [20] N. Parker, A. Martin, S. Cornish, C. Adams, Collisions
[7] C. D’Errico, A. Burchianti, M. Prevedelli, L. Salas- of bright solitary matter waves, Journal of Physics B:
nich, F. Ancilotto, M. Modugno, F. Minardi, C. Fort, Atomic, Molecular and Optical Physics 41 (4) (2008)
Observation of quantum droplets in a heteronuclear 045303.
bosonic mixture, Phys. Rev. Res. 1 (2019) 033155. [21] A. Marchant, T. Billam, M. Yu, A. Rakonjac, J. Helm,
doi:10.1103/PhysRevResearch.1.033155. J. Polo, C. Weiss, S. Gardiner, S. Cornish, Quantum re-
URL https://link.aps.org/doi/10.1103/ flection of bright solitary matter waves from a narrow
PhysRevResearch.1.033155 attractive potential, Physical Review A 93 (2) (2016)
[8] M. Schmitt, M. Wenzel, F. Böttcher, I. Ferrier-Barbut, 021604.
T. Pfau, Self-bound droplets of a dilute magnetic quan- [22] H. Sakaguchi, M. Tamura, Scattering and trapping of
tum liquid, Nature 539 (7628) (2016) 259–262. nonlinear schrödinger solitons in external potentials,
[9] L. Chomaz, S. Baier, D. Petter, M. Mark, F. Wächtler, Journal of the Physical Society of Japan 73 (3) (2004)
L. Santos, F. Ferlaino, Quantum-fluctuation-driven 503–506.
crossover from a dilute bose-einstein condensate to a [23] R. H. Goodman, P. J. Holmes, M. I. Weinstein, Strong
macrodroplet in a dipolar quantum fluid, Physical Re- nls soliton–defect interactions, Physica D: Nonlinear Phe-
view X 6 (4) (2016) 041039. nomena 192 (3-4) (2004) 215–248.
[10] Z.-H. Luo, W. Pang, B. Liu, Y.-Y. Li, B. A. Malomed, A [24] C. Lee, J. Brand, Enhanced quantum reflection of
new form of liquid matter: Quantum droplets, Frontiers matter-wave solitons, Europhysics Letters 73 (3) (2005)
of Physics 16 (2021) 1–21. 321.
[11] A. Khan, A. Debnath, Quantum droplet in lower dimen- [25] T. Ernst, J. Brand, Resonant trapping in the transport of
sions, Frontiers in Physics (2022) 534. a matter-wave soliton through a quantum well, Physical
[12] D. Petrov, G. Astrakharchik, Ultradilute low- Review A 81 (3) (2010) 033614.
dimensional liquids, Physical review letters 117 (10)
13

[26] L. D. Carr, R. R. Miller, D. R. Bolton, S. A. Strong, Non- interferometry, Physical Review A 89 (3) (2014) 033610.
linear scattering of a bose-einstein condensate on a rect- [37] O. J. Wales, A. Rakonjac, T. P. Billam, J. L. Helm,
angular barrier, Physical Review A 86 (2) (2012) 023621. S. A. Gardiner, S. L. Cornish, Splitting and recombi-
[27] C.-H. Wang, T.-M. Hong, R.-K. Lee, D.-W. Wang, nation of bright-solitary-matter waves, Communications
Particle-wave duality in quantum tunneling of a bright Physics 3 (1) (2020) 51.
soliton, Optics Express 20 (20) (2012) 22675–22682. [38] C. L. Grimshaw, T. P. Billam, S. A. Gardiner, Soliton
[28] A. Harel, B. A. Malomed, Interactions of spatial solitons interferometry with very narrow barriers obtained from
with fused couplers, Physical Review A 89 (4) (2014) spatially dependent dressed states, Physical Review Let-
043809. ters 129 (4) (2022) 040401.
[29] L. Al Sakkaf, U. Al Khawaja, Reflectionless potentials [39] H. Sakaguchi, B. A. Malomed, Matter-wave soliton inter-
and resonant scattering of flat-top and thin-top soli- ferometer based on a nonlinear splitter, New Journal of
tons, Phys. Rev. E 107 (2023) 014202. doi:10.1103/ Physics 18 (2) (2016) 025020.
PhysRevE.107.014202. [40] N. Vakhitov, A. Kolokolov, Stability criterion for solitary
URL https://link.aps.org/doi/10.1103/PhysRevE. waves, Izv Vyssh Uchebn Zaved Radiofiz 16 (2) (1973)
107.014202 1020–1026.
[30] J. L. Helm, T. P. Billam, S. A. Gardiner, Bright matter- [41] L. Bergé, Wave collapse in physics: principles and appli-
wave soliton collisions at narrow barriers, Physical Re- cations to light and plasma waves, Physics reports 303 (5-
view A 85 (5) (2012) 053621. 6) (1998) 259–370.
[31] B. Gertjerenken, T. P. Billam, L. Khaykovich, C. Weiss, [42] G. Fibich, The nonlinear Schrödinger equation: singular
Scattering bright solitons: Quantum versus mean-field solutions and optical collapse, Vol. 192, Springer, 2015.
behavior, Physical Review A 86 (3) (2012) 033608. [43] D. J. Korteweg, G. De Vries, Xli. on the change of form
[32] S.-C. Li, F.-Q. Dou, Matter-wave interactions in two- of long waves advancing in a rectangular canal, and on
component bose-einstein condensates, Europhysics Let- a new type of long stationary waves, The London, Edin-
ters 111 (3) (2015) 30005. burgh, and Dublin Philosophical Magazine and Journal
[33] C. L. Grimshaw, S. A. Gardiner, B. A. Malomed, Split- of Science 39 (240) (1895) 422–443.
ting of two-component solitary waves from collisions with [44] S. Novikov, S. V. Manakov, L. P. Pitaevskii, V. E.
narrow potential barriers, Physical Review A 101 (4) Zakharov, Theory of solitons: the inverse scattering
(2020) 043623. method, Springer Science & Business Media, 1984.
[34] A. Martin, J. Ruostekoski, Quantum dynamics of atomic [45] L. Wang, B. A. Malomed, Z. Yan, Attraction centers and
bright solitons under splitting and recollision, and impli- parity-time-symmetric delta-functional dipoles in critical
cations for interferometry, New Journal of physics 14 (4) and supercritical self-focusing media, Physical Review E
(2012) 043040. 99 (5) (2019) 052206.
[35] J. Polo, V. Ahufinger, Soliton-based matter-wave inter- [46] Y. S. Kivshar, B. A. Malomed, Dynamics of solitons in
ferometer, Physical Review A 88 (5) (2013) 053628. nearly integrable systems, Reviews of Modern Physics
[36] J. Helm, S. Rooney, C. Weiss, S. Gardiner, Splitting 61 (4) (1989) 763.
bright matter-wave solitons on narrow potential barri-
ers: Quantum to classical transition and applications to

You might also like