ArticlePDF Available

Quantumlike statistics of deterministic wave-particle interactions in a circular cavity

Authors:

Abstract and Figures

A deterministic low-dimensional iterated map is proposed here to describe the interaction between a bouncing droplet and Faraday waves confined to a circular cavity. Its solutions are investigated theoretically and numerically. The horizontal trajectory of the droplet can be chaotic: it then corresponds to a random walk of average step size equal to half the Faraday wavelength. An analogy is made between the diffusion coefficient of this random walk and the action per unit mass h/m of a quantum particle. The statistics of droplet position and speed are shaped by the cavity eigenmodes, in remarkable agreement with the solution of Schrödinger equation for a quantum particle in a similar potential well.
Content may be subject to copyright.
PHYSICAL REVIEW E 93, 042202 (2016)
Quantumlike statistics of deterministic wave-particle interactions in a circular cavity
Tristan Gilet*
Microfluidics Lab, Department of Aerospace and Mechanics, University of Li`
ege, B-4000 Li`
ege, Belgium
(Received 21 January 2016; published 5 April 2016)
A deterministic low-dimensional iterated map is proposed here to describe the interaction between a bouncing
droplet and Faraday waves confined to a circular cavity. Its solutions are investigated theoretically and numerically.
The horizontal trajectory of the droplet can be chaotic: it then corresponds to a random walk of average step size
equal to half the Faraday wavelength. An analogy is made between the diffusion coefficient of this random walk
and the action per unit mass /m of a quantum particle. The statistics of droplet position and speed are shaped
by the cavity eigenmodes, in remarkable agreement with the solution of Schr¨
odinger equation for a quantum
particle in a similar potential well.
DOI: 10.1103/PhysRevE.93.042202
I. INTRODUCTION
The interaction between Faraday waves and millimeter-
sized bouncing droplets (Fig. 1) has attracted much atten-
tion over the past decade owing to its reminiscence of
quantum-particle behavior [1,2]. Faraday waves are stationary
capillary waves at the surface of a vertically vibrated liquid
bath [35]. They appear spontaneously and sustainably when
the vibration amplitude exceeds the Faraday threshold. At
smaller amplitude, they only appear in response to a finite
perturbation (e.g., the impact of a bouncing droplet) and
are then exponentially damped [6]. Close to threshold the
Faraday instability is often subharmonic, i.e., the Faraday wave
frequency is half the frequency of the external vibration [4].
The selected Faraday wavelength λFis approximately given
by the dispersion relation of water waves. The memory Mis
defined as the decay time of Faraday waves below threshold,
divided by the wave period; it increases and diverges as the
threshold is approached.
Liquid droplets are able to bounce several successive times
on liquid interfaces before merging, and this in a wide range
of conditions [79]. Rebounds can be sustained by vertically
vibrating the liquid interface [1014]. When the rebound
dynamics locks into a periodic state with one impact every two
forcing periods, the droplet becomes a synchronous emitter of
Faraday waves [15]. More exactly, the droplet creates a radially
propagating circular capillary wave that excites standing
Faraday waves in its wake [6]. The resulting wave field then
contains contributions from the last Mdroplet impacts. In
the walking state, horizontal momentum is transferred from
the wave field to the impacting droplet, proportionally to the
local wave slope at the droplet position. The waves only exist
in response to droplet impacts, and in turn they cause the
horizontal motion of the droplet. This coupled wave-particle
entity at the millimeter scale is called a walker.
The dynamics of individual walkers has been investigated
experimentally in several configurations where it has shown
properties reminiscent of quantum particles [1]. For example,
walkers tunnel through weak boundaries [16]. They diffract or
interfere when individually passing through one or two slits,
respectively [17] (although these experimental results have
*Tristan.Gilet@ulg.ac.be
recently been challenged by theoretical arguments [18,19]).
Walkers also experience quantized orbits in response to
confinement by either central forces [20], Coriolis forces
[2123], or geometry [2426]. In the latter case, the walker
evolves in a cavity of finite horizontal extent. Chaotic tra-
jectories are often observed in the high-memory limit, when
confinement compels the walker to cross its own path again
after less than Mimpacts. In response to central and Coriolis
forces, the walker oscillates intermittently between several
different quantized orbits, as if it were in a superposition
of trajectory eigenstates [22,27,28]. The spatial extent of
these eigenstates is always close to an integer multiple of
half the Faraday wavelength λF/2, which can then be seen
as the analog of the de Broglie wavelength for quantum
particles [21]. The probability to find the walker in a given
state is proportional to the relative amount of time spent in
this state. In cavity experiments [24], the probability to find
the walker at a given position is strongly shaped by the cavity
geometry, as would be the statistical behavior of a quantum
particle in a potential well.
Several theoretical models have already been proposed
that capture various aspects of walker dynamics [6,25,2831].
Most of them consider a stroboscopic point of view, where the
droplet is assumed to impact the bath perfectly periodically.
The wave field is then expressed as a sum of contributions from
each successive impact [6]. The walker horizontal trajectory
is deduced from Newton’s second law, where the driving force
is proportional to the local wave slope and the effective mass
includes both the real droplet mass and an added mass from the
wave [32]. The resulting discrete iterated map can be turned
into an integrodifferential equation for the walker trajectory in
the limit of small horizontal displacements between successive
impacts [29]. In an infinite space, the contribution from one
droplet impact to the wave field is assumed to be a Bessel
function of the first kind J0[kF(xxd)] centered on the droplet
position xd, and of wave number kF=2π/λF[6]. The recent
model of Milewski et al. [31] provides a more sophisticated
model of the wave fields through inclusion of weak viscous
effects. It also includes realistic models of vertical bouncing
dynamics [13,33]. At the other end of the scale of complexity,
a generic model [30] has been proposed that reproduces
some key features of confined wave-particle coupling and
walker dynamics within a minimal mathematical framework.
These features include the time decomposition of the chaotic
2470-0045/2016/93(4)/042202(15) 042202-1 ©2016 American Physical Society
TRISTAN GILET PHYSICAL REVIEW E 93, 042202 (2016)
FIG. 1. (top) A drop of silicone oil bounces periodically at the
surface of a small pool (diameter 1.5 mm) of the same liquid which
is vertically vibrated at 80 Hz. (bottom) Drops of appropriate size
(here, diameter of 0.74 mm) can generate and interact with underlying
Faraday waves. This wave-particle association is called a walker. The
bottom left (right) picture shows a walker at low (high) memory, i.e.,
far from (close to) the Faraday-instability threshold.
trajectory into eigenstates [28], and the particle statistics being
shaped by confinement [24].
These hydrodynamic experiments and models have already
revealed a strong analogy between the statistical behavior
of chaotic walkers and quantum particles, in many different
configurations. It certainly results from the deterministic chaos
inherent to such wave-particle coupling. Nevertheless, it is
still unclear which ingredients are actually necessary and
sufficient for these quantum behaviors to appear. Also, no
direct connection between the equations of motion of the
walker and the Schr¨
odinger equation has been made yet. What
would be the equivalent of Planck’s constant for walkers? On
which timescale should the walker dynamics be averaged to
recover quantumlike statistics? This work aims at answering
these questions through a theoretical investigation of walker
dynamics under confinement in a two-dimensional cavity
[24].
First, the model developed in Ref. [30] is particularized
to a walker in a circular cavity (Sec. II). In Sec. III the
walker dynamics is analyzed as a function of its memory,
and results are compared to the experiments of Harris
et al. [24]. In Sec. IV it is shown that the present model gives
theoretical access to the experimentally unachievable limit of
zero damping (infinite memory) and perfect mode selection.
Finally, the above questions are addressed through direct com-
parison with analytical predictions from quantum mechanics
(Sec. V).
II. WAVE-PARTICLE COUPLING IN A CONFINED
GEOMETRY
The generic model [30] of the interaction between a particle
and a stationary wave confined to a domain Sis first recalled,
here in a dimensional form. It is primarily based on the
decomposition of the standing Faraday wave field Hn(X)
resulting from the nfirst impacts into the discrete basis of
eigenmodes k(X) of the domain:
Hn(X)=
k
Wk,nk(X),W
k,n =S
Hn(X)
k(X)dS, (1)
where
kis the conjugate of k. These eigenmodes are
orthonormal on S. They are here chosen to satisfy the Neumann
condition n·k=0 at the boundary ∂S with normal
vector n.
We consider the stroboscopic approach where the droplet
impacts the liquid bath at regular time intervals. At rebound
n, it impacts at position Xnand creates a crater of shape Z=
F(R) in the wave field. In the limit of a droplet size much
smaller than λF, the crater F(R) can be approximated by a δ
function weighted by the volume of liquid displaced [30]:
F(R)=
k
k(Xn)k(R)=δ(RXn).(2)
The contribution 
k(Xn) of the impact to each wave
eigenmode depends on the droplet position.
Each mode k(X) is given a viscous damping factor μk
[0,1] that depends on the forcing amplitude. It is defined as
the amplitude of mode kright before impact n+1, divided by
its amplitude right after impact n. The associated memory Mk
is given by Mk=−1/ln μk. The Faraday threshold instability
corresponds to max(μk)=1. The amplitude of each mode
Wk,n then satisfies the recurrence relation
Wk,n+1=μk[
k(Xn)+Wk,n]
=
n
n=0
μn+1n
k
k(Xn).(3)
At each impact, the particle is shifted proportionally to the
gradient of the wave field at the impact position:
Xn+1Xn=−δ
k
Wk,n k]Xn
=−δ
k
n1
n=0
μnn
k]Xnk]Xn,(4)
where δ>0 is the proportionality constant between the local
wave slope and the particle displacement. Equations (3) and (4)
form an iterated map that describes the evolution of the particle
in a cavity of arbitrary shape and dimension.
A. The circular cavity
Focus is now made on a two-dimensional circular cavity
of radius Rc, as studied experimentally by Harris et al. [24].
Cavity eigenfunctions are
k =φk
Rc
k(r,θ )=ϕk(r)eikθ ,kZ,N0,
(5)
042202-2
QUANTUMLIKE STATISTICS OF DETERMINISTIC WAVE- . . . PHYSICAL REVIEW E 93, 042202 (2016)
TABLE I. Twelve dominant Neumann eigenmodes for a cavity
of radius 14.3 mm filled with 20 cS oil and forced at 83 Hz.
Mode k, ΛFμ M Mode k, ΛFμ M
2,6 0.31 0.999 738 9,3 0.33 0.983 60
0,7 0.32 0.998 460 18,1 0.34 0.974 39
4,5 0.34 0.993 141 10,3 0.32 0.968 32
13,2 0.33 0.990 99 12,2 0.34 0.962 26
17,1 0.33 0.989 90 6,4 0.34 0.951 20
7,4 0.32 0.987 77 5,5 0.32 0.933 15
where r=R/Rcis the dimensionless radial position, θis the
angular position, and φk are the dimensionless eigenfunctions.
The dimensionless Faraday wavelength is defined as F=
λF/Rc. The radial functions ϕk are expressed as
ϕk(r)=1
π1k2
z2
k Jk(zkr)
Jk(zk),(6)
where zk is the th zero of the derivative of the Bessel function
(first kind) of order k,soJ
k(zk)=0. The integer kcan take
both negative and positive values. The eigenfunctions satisfy
orthonormality conditions
2π
0
1
0
φkφ
krdr =δkkδ(7)
as well as the Neumann boundary condition rφk =0inr=
1. Some of these eigenmodes are illustrated in Table I.
The dimensionless Faraday wave field hn(r,θ )=HnR2
c/
can be decomposed as a Dini series in this basis:
hn(r,θ )=+∞
k=−∞
=1
wk,nφk(r,θ ),(8)
with
wk,n =Rc
Wk,n =2π
0
1
0
hn(r,θ )φ
krdr. (9)
The condition wk,n =w
k,n results from hn(r,θ ) being real-
valued. Therefore,
hn(r,θ )=+∞
k=0
=1
ξkRe[wk,nφk(r,θ )],(10)
where ξk=2 when k>0 and ξk=1 when k=0. For the sake
of readability, we remove the factor ξkand index from most
notations, keeping in mind that, when cylindrical harmonics
are involved, functions that depend on kalso depend on .
Specifically, the abbreviated sum over kwill always refer to
a double sum over kand ,from0toand from 1 to ,
respectively, with inclusion of the factor ξk.
The particle position is expressed in cylindrical coor-
dinates Xn/Rc=(rncos θn,rnsin θn). The wave recurrence
relation (3) then becomes
wk,n+1=μk[wk,n +ϕk(rn)eikθn].(11)
The trajectory equation (4) can be projected along Xnthen in
the direction perpendicular to Xn:
rn+1cos (θn+1θn)=rnC∂hn
∂r (rnn)
,
(12)
rn+1sin (θn+1θn)=−C
rn∂hn
∂θ (rnn)
,
where
C=δ
R4
c
(13)
is a dimensionless constant that characterizes the intensity of
the wave-particle coupling. After calculating the gradient of
hnat the impact point, the iterated map becomes
rn+1cos (θn+1θn)=rnC
k
ϕ
k(rn)Re[wk,neikθn],
rn+1sin (θn+1θn)=C
rn
k
k(rn)Im[wk,neikθn],(14)
wk,n+1=μk[wk,n +ϕk(rn)eikθn],
where ϕ(r)=dϕ/dr.
B. Model discussion
This model of walkers neglects traveling capillary waves,
similarly to most previous works [6]. Indeed, these waves are
not re-energized by the vertical forcing, and their initial energy
spreads in two dimensions. Their amplitude is then smaller
than Faraday waves, and their contribution is likely to have
only marginal influence on the walker’s long-term behavior.
Only when the walker comes close to the boundary could these
capillary waves significantly modify the local trajectory. But
this effect would then be localized in space and time, so we
assume that it does not strongly affect the walker statistics.
In contrast to previous models [29,31], the trajectory
equation (4) is here first order in time. It therefore assumes that
the walker completely forgets its past velocity at each impact,
i.e., horizontal momentum is entirely dissipated. Nevertheless,
an inertial effect is still present, because the walker keeps a
constant velocity between successive impacts. An additional
inertial term, e.g., corresponding to the hydrodynamic boost
factor [32], would increase the model complexity and the
number of parameters, so it is left to future work.
This model should be compared to the experimental results
of a walker in a circular cavity reported by Harris et al. [24]. A
necessary step to a quantitative comparison is the accurate
determination of each eigenmode kand its associated
damping factor μk. Unfortunately, these modes are not easily
characterized experimentally, since they cannot be excited
one at a time. Uncertainty in the boundary conditions further
complicates mode identification. Ideal Dirichlet boundary
conditions (zero velocity where the liquid-air interface meets
the solid walls) could have been expected if the contact line
was pinned [34]. But in experiments contact lines are avoided
042202-3
TRISTAN GILET PHYSICAL REVIEW E 93, 042202 (2016)
on purpose because the vibration of the associated meniscus
would be a source of parasitic capillary waves. Instead, the
solid obstacles that provide confinement are always slightly
submerged [16,17,24,26]. The walker cannot penetrate these
shallow regions since waves are strongly dissipated there [16].
Recent numerical simulations suggest that such boundary
condition cannot be strictly expressed as a Robin condition
[35]. From a practical point of view, when Dirichlet boundary
conditions are considered in Eq. (14), the particle tends to
leave the cavity as its radial displacement does not vanish at
the boundary. For this reason, Neumann boundary conditions
(zero wave slope) are adopted in this work.
Determining the damping factor associated with each mode
is also an issue. In the absence of horizontal confinement,
viscous wave damping can be estimated by spectral meth-
ods [5,31]. The theoretical prediction is then validated by
measuring the Faraday threshold amplitude as a function
of forcing frequency. Unfortunately, confinement with sub-
merged vertical walls yields additional viscous dissipation.
Indeed, modes φk of high kand small are never observed
in experiments right above the Faraday threshold (D. Harris,
private communication), while spectral methods would predict
them as highly unstable.
III. FINITE MEMORY IN A DAMPED WORLD
Harris and coworkers [24] considered a circular cavity
of radius Rc=14.3 mm filled with silicone oil of density
965 kg/m3, surface tension σ=20 mN/m, and kinematic
viscosity ν=20 cS. The forcing frequency f=70 Hz was
chosen in relation with the cavity size, such that the most
unstable mode at Faraday threshold is one of the radial
modes (k=0, = 0). In this work, damping factors are
approximately determined from a spectral method [5] for this
same cavity with Neumann boundary conditions and various
forcing frequencies. A frequency of 83 Hz is then selected
that again gives predominance to one of the purely radial
modes. Only modes of damping factor μk>0.01 are retained.
The twelve modes of highest damping factor are illustrated in
Table I. Their wavelength is always very close to the Faraday
wavelength that would result from this forcing frequency in
the absence of confinement.
The average walking speed vwincreases with the coupling
constant C. In experiments [24], vw=8.66 mm/s at a forcing
frequency of f=70 Hz and at 99% of the Faraday threshold.
This corresponds to an average step size 2vw/f =0.017Rc.
The same value is obtained when Eq. (14) is solved with
C=3×105. If we further assume that δ2vw/f ,the
volume of fluid displaced at each impact can be estimated
from Eq. (13)tobe5μL. This volume corresponds to
an interface deflection of a few hundred micrometers on a
horizontal length scale of a few millimeters. It is one order
of magnitude higher than the droplet volume, even when
corrected with the boost factor [32].
A. Circular orbits
The iterated map (14) admits periodic solutions (Fig. 2)
where the particle orbits at constant speed around the center of
the cavity: rn=r1,θn=αn, and wk,n =wk1eikαn (the wave
pattern also rotates). The walking velocity is r1α. Plugging
FIG. 2. Convergence to a stable circular orbit of type A (r1=
0.592, α=0.0135) at memory M=7.7andC=3×105.The
underlying wave-field corresponds to the time at which the walker is
at the position indicated by the black circle ().
this solution into the iterated map yields
wk1=μk
eikα μk
ϕk1,
r1(1cos α)=C
k
ϕk1ϕ
k1μk
cos ()μk
12μkcos ()+μ2
k
,
r2
1sin α=C
k
2
k1μk
sin ()
12μkcos ()+μ2
k
,(15)
where ϕk1=ϕk(r1) and ϕ
k1=[k/dr](r1). A detailed analy-
sis of these periodic solutions is given in Appendix A.Atlow
memory, the only solution to Eq. (15)isα=0 (fixed points).
The corresponding radii r1satisfy
(r1)=0,with (r)=
k
μk
1μk
ϕ2
k(r)
2.(16)
As memory increases (i.e., as all μkincrease), each of these
fixed points experiences a pitchfork bifurcation where two
other solutions ±α= 0 appear that correspond to clockwise
and counterclockwise orbits. This bifurcation is analogous to
the walking threshold observed and rationalized for unconfined
walkers [6,29]. These orbits emerge from the loss of stability
of their corresponding fixed point. The radial stability of
both fixed points and corresponding orbits is related to
(r), as already shown in Ref. [30] for the unimodal and
unidimensional version of Eqs. (3) and (4). They are usually
stable when (r1)>0 (orbits of type A) and unstable when
(r1)<0 (orbits of type B). For single-frequency forcing,
successive orbits of a same type are approximately separated
by F/2. The convergence of trajectories towards stable orbits
of type A involves wobbling, i.e., radial oscillations (Fig. 2).
042202-4
QUANTUMLIKE STATISTICS OF DETERMINISTIC WAVE- . . . PHYSICAL REVIEW E 93, 042202 (2016)
r
0 0.2 0.4 0.6 0.8 1
M
5
6
7
8
9
10
15
20
30
50
100
FIG. 3. Bifurcation diagram of the radial position ras a function
of memory Mat C=3×105. The probability distribution function
is represented in levels of gray; darker regions correspond to more
frequently visited radial positions. Thick solid lines and thin dashed
lines correspond to stable and unstable fixed points, respectively
[Eq. (16)]. Full and empty symbols represent stable and unstable
circular orbits, respectively [Eq. (15)]. Circles () indicate that
at least one pair of eigenvalues is complex-conjugated (type A)
while triangles () are used when all corresponding eigenvalues are
real-valued (type B).
B. Transition to chaos
As memory increases, each circular orbit of type A
destabilizes radially through a Neimark–Sacker bifurcation
(Fig. 3)[30]. Orbits of larger radius destabilize at higher
memory, since it takes more rebounds before the particle
revisits its past positions. Sustained wobbling orbits of finite
amplitude are then observed, where the radial position oscil-
lates periodically. As memory is increased further, wobbling
orbits progressively disappear. At high memory (M>50 in
Fig. 3), there are no stable periodic attractors left and the walker
chaotically explores the entire cavity. The chaotic nature of the
system is highlighted by the exponential growth of the distance
between two trajectories that were initially extremely close to
one other (positive Lyapunov exponent in Fig. 4).
A chaotic trajectory is represented in Fig. 5. Several
trajectory patterns are recurrent, including abrupt changes in
direction and off-centered loops of characteristic radius close
to half the Faraday wavelength F/2. Overall, the walker is
observed to spend significantly less time at radial positions
where orbits were stable at low memory (Fig. 3).
IV. ACCESSIBLE INFINITE MEMORY
Unlike its hydrodynamic analog, a quantum particle con-
fined in a cavity is known to be a Hamiltonian system for which
there is no dissipation. The behavior of individual walkers that
are reminiscent of quantum particles were always observed
at high memory, i.e., when dissipation is almost balanced
by external forcing. Indeed, in these conditions, the walker
gets a chance to walk onto the wave field left in its own
N
1000 1200 1400 1600 1800 2000
|r1-r2|
10-12
10-10
10-8
10-6
10-4
10-2
100
x
-0.5 0 0.5
y
-1.2
-1
-0.8
-0.6
-0.4
-0.2
FIG. 4. Exponential divergence of the radial distance |r1r2|be-
tween two neighboring trajectories, initially separated by |r1r2|=
1010 and placed on the same initial wave-field. Memory and coupling
constants are M=500 and C=3×105, respectively. The inset
shows the two trajectories represented with a solid line and dots,
respectively, for N[1400,1600].
wake [20,22,36]. For each mode k, this balance is theoretically
achieved when μk=1 (infinite memory). Memory above 100
is very challenging to achieve experimentally, owing to the
imperfect control of the forcing vibration [37]. Moreover, in
most geometries (including circular), cavities are such that
it is impossible to get several modes reaching μk=1for
the same forcing acceleration. This model does not suffer
from such limitations; it is possible to select a subset Sof
x
-1 -0.5 0 0.5 1
y
-1
-0.8
-0.6
-0.4
-0.2
0
0.2
0.4
0.6
0.8
1
FIG. 5. Chaotic trajectory at memory M=500 and C=3×
105. The solid line corresponds to 10 000 impacts, while the dots
emphasize 380 of these successive impacts.
042202-5
TRISTAN GILET PHYSICAL REVIEW E 93, 042202 (2016)
TABLE II. Subsets of selected modes S, corresponding range of
wavelength and coupling constant C, and associated symbols and
colors in subsequent figures.
SSelected mo des (k, )∈S ΛFΔΛFlog10 C
S1–(0,5); (2,4); (6,3); 0.473 4% -5
(9,2); (13,1)˝
S2S1–(3,4); (5,3); 0.473 8% -6.5
(7,3);(10,2);(14,1) ˝ -6
-5.5
-5
-4.5
-4
-3.5
-3
-2.5
S3S2–(1,5); (1,4); (4,4); 0.473 16% -5
(4,3); (8,3); (8,2);
(11,2); (12,1); (15,1) ˝
S4–(0,12); (2,11); (7,9); 0.178 1% -6
(10,8); (13,7); (17,6);
(26,3); (30,2); (35,1) ˝
S5–(0,31); (2,30) ; (11,26) ; 0.066 0.2% -7
(24,21); (27,20); (47,13);
(62,8); (69,6); (77,4) ˝
S6S1–(0,12); (2,11); 0.473 4% -5.3
(7,9); (13,7) ; (17,6) ˝ + 0.178 0.5%
modes for which the memory is infinite (μk=1,k S) while
all other modes are instantaneously damped (zero memory,
μk=0,k /S). In the following, several subsets of selected
modes Sare considered (Table II). Each subset includes all the
modes for which the characteristic wavelength lies in a narrow
range F[1 F,1+F], except the latest subset which
is a combination of two wavelengths. Such mode selection
could be analog to preparing quantum particles with localized
momentum. For subset S2the coupling constant Cis varied
over four orders of magnitude.
A. Radial statistics
Figure 6shows a superposition of half a million impact po-
sitions. Dark circular strips indicate more frequent impacts at
certain radial positions r, as observed in experiments [24]. The
corresponding probability distribution function ρ(r)showsa
series of extrema at radii that are largely independent of C
[Fig. 7(a)]. Chaotic trajectories in the infinite-memory limit
are qualitatively identical to those arising at finite memory.
The radial position oscillates intermittently between several
preferred radii that correspond to the maxima of ρ(r)[30]
[Fig. 7(b)].
The iterated map is checked to be ergodic to a good
approximation (within 1% in these simulations), so the time-
average of any property over a single trajectory coincides with
an ensemble average of this property over many independent
FIG. 6. Two-dimensional probability distribution function ob-
tained by superimposing half a million impact positions issued from
78 independent trajectories. The subset of selected modes is S2
(Table II), with C=105.
trajectories at a given time. In other words,
lim
N→∞
1
N
n+N
n=n
f(rnn)=2π
01
0
ρ(r)f(r,θ )rdrdθ, (17)
for any regular function f(rnn). It is rather counterintuitive
that the wave-field does not blow up with time, since contri-
butions from previous impacts keep adding up without any of
them being damped out. The resulting wave field amplitudes
satisfy
wk,n =
n1
n=0
ϕk(rn)eikθn.(18)
In the long term, ergodicity yields
lim
n→∞ wk,n =n2π
0
eikθ1
0
ρ(r)ϕk(r)rdr. (19)
The angular integral vanishes by symmetry, except when
k=0. In this latter case, the only way to keep a finite wave
amplitude is to satisfy
1
0
ρ(r)ϕ0(r)rdr =0.(20)
So ρ(r) must always be orthogonal to each of the radial
eigenmodes φ0(r). We have checked numerically that for every
subset of selected modes S(Table II), the component of ρ(r)
along φ0(r) is always at least two orders of magnitude smaller
than the largest other component.
042202-6
QUANTUMLIKE STATISTICS OF DETERMINISTIC WAVE- . . . PHYSICAL REVIEW E 93, 042202 (2016)
ρ(r)
0 0.2 0.4 0.6 0.8 1
r
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1
(a)
n
0 200 400 600 800 1000
rn
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1
(b)
FIG. 7. (a) Probability distribution function ρ(r) of radial position rfor mode subset S2and three different values of C(Table II). The solid
line corresponds to the quantum prediction given in Eq. (35). (b) An example of the time evolution of the particle’s radial position for mode
subset S2and for C=105.
B. Coherence and diffusion
The walker trajectory is usually smooth and regular at the
scale of a few impacts (Fig. 5). But such coherence is lost
on longer timescales, where the trajectory oscillates and loops
chaotically. This behavior is quantified through the average
distance d(n)traveledinnsteps, defined as
d(n)=||xn+nxn||2n,(21)
and represented in Fig. 8. For small nthis distance increases
linearly with nso the motion is ballistic (Fig. 8). As soon
as dF/2, dstarts increasing proportionally to n, like
in normal diffusion. It then saturates in d2 owing to the
n / Nc
10-2 10-1 100101102103
d / ΛF
10-2
10-1
100
101
n / Nc
0.01 1 100
d / ΛF
0.01
0.1
1
FIG. 8. Average distance dtraveled in nsuccessive impacts, for
subsets S1,S4,S5and S6. The solid line corresponds to the ballistic
regime d/F=an/Nc, while the dashed line corresponds to the
diffusive regime d/F=bn/Nc. (Inset) Same plot, for S1,S2and
S3. Symbols and colors are explained in Table II.
finite size of the cavity. This ballistic-to-diffusive transition
was already observed for the one-dimensional (1D) version
of this model [30]. This suggests that the walker trajectory is
similar to a random walk for which the elementary steps are
of the order of F/2.
The average number of impacts in one step should depend
on the walker velocity. It can be estimated by first looking
at the amplitude wk,n of each wave eigenmode (see inset of
Fig. 9). Their time evolution is a succession of coherent linear
r
0 0.2 0.4 0.6 0.8 1
<|wk,n+1-wk,n|>
0
0.5
1
1.5
2
2.5
3
n
0 100 200 300 400
Re(wk,n)
-40
-20
0
20
40
Nw
FIG. 9. Average increment of wave amplitude at each impact
|wk,n+1wk,n| as a function of particle position r, for modes of the
subset S2. The solid line is |ϕk(r)|. Modes (k,)=(0,5) and (6,3) are
represented in black and gray, respectively. Inset shows an example
of the time evolution of the wave amplitude Re(wk,n), for these same
modes. Nwis defined as the distance between two successive extrema
of wk,n.
042202-7
TRISTAN GILET PHYSICAL REVIEW E 93, 042202 (2016)
Nw
0 50 100 150 200
PDF
0
0.005
0.01
0.015
0.02
0.025
0.03
0.035
0.04
Nc
1 10 100 1000
<Nw>
1
10
100
1000
ΛF/2
0 0.1 0.2 0.3
Lw
0
0.1
0.2
0.3
FIG. 10. Probability distribution function of the coherence time
Nwfor the two modes (k,)=(0,5) and (6,3) of subset S2(in black
and gray, respectively). The vertical line corresponds to Nw=Ncas
defined in Eq. (23). Top inset shows average distance Lwtraveled
during a coherent segment Nwas a function of the half Faraday
wavelength F/2. Bottom inset shows average number of impacts
Nwof a coherent segment as a function of Nc. In both insets, data
from all subsets are represented with different symbols (Table II), and
the solid line corresponds to equal abscissa and ordinate.
segments of variable duration Nw. This constant growth rate
depends on the particle position; at every impact, each wk
increases by an average amount
|wk,n+1wk,n|n|ϕk(r)|,(22)
as confirmed in Fig. 9. The statistical distribution of Nw
peaks at approximately the same value for each mode of a
given selection, then it decreases exponentially for large Nw
(Fig. 10). So after the average coherence time Nweach
mode amplitude wkhas grown by about Nw|ϕk|and its
contribution to the total wave slope now scales as ξkwkϕk
2ξkNcϕ2
k/F. The average displacement per impact is then
x 2Nwϕ2
k/(F), where χ=kξk. The average
distance Lw=[||xn+Nwxn||2n]1/2traveled during one of
these coherent segments is observed to be close to F/2,
independently of Cor F(see upper inset in Fig. 10).
We deduce that Nw∼F(4Cχϕ2
k)1/2, which yields the
definition of the coherence timescale
NcF
=λFRc
δχ .(23)
The bottom inset of Fig. 10 validates the scaling law Nw
Ncover four decades of C,forseveralFand χ, with a
proportionality constant almost equal to unity. Even the subset
S6of mixed wavelengths () satisfies these scaling arguments.
The curves of average distance vs time perfectly collapse
on a single curve independent of F(or χ) and C, when n
and dare normalized by Ncand F, respectively (see inset
of Fig. 8). Only subsets corresponding to Nc<4 (i.e., the
two largest values of Cfor S2) fail to collapse perfectly. The
ballistic regime is described by
d
F=an
Nc
,(24)
where a0.57. Similarly the diffusive regime satisfies
d
F=bn
Nc
,(25)
where b0.62 based on the data of subset S5, for which the
diffusive region is the largest since Fis the smallest (Fig. 8).
If the diffusive behavior is attributed to a two-dimensional (2D)
random walk, then the corresponding diffusion coefficient is
defined as
˜
Dd2
4n=b2
4
2
F
Nc
.(26)
The crossover between both regimes occurs in n/Nc=
b2/a21.2, so in d/F=b2/a 0.67. It can be seen as the
elementary step of this random walk, and it is slightly larger
than half the Faraday wavelength. In dimensional terms, the
diffusion coefficient is
D=R2
cf˜
D=b2
4
λ2
Ff
Nc0.096 λ2
Ff
Nc
,(27)
where fis the impact frequency.
V. COMPARISON WITH PREDICTIONS OF QUANTUM
MECHANICS
In this section, the analogy between walkers at infinite
memory and quantum particles is further developed. More
exactly, the solutions of the map (14) are compared to the
predictions of the Schr¨
odinger equation for a quantum particle
subject to the same confinement.
A. Correspondence of timescales
A single quantum particle is statistically described by a
wave function (x,t): The probability to find the particle
at position xis given by ρ(x,t)=||2. For nonrelativistic
free particles, the wave function evolves according to the
Schr¨
odinger equation
it+
2
2m2=0,(28)
where is the reduced Planck constant and mis the rest mass
of the particle. Special relativity can be taken into account by
instead considering the relativistic Klein–Gordon equation for
the function ψ(x,t), which applies for spinless particles:
1
c2ttψ+m2c2
2ψ−∇2ψ=0.(29)
In his pilot-wave theory, de Broglie [38] hypothesized
that quantum particles would vibrate at the Compton fre-
quency ωc=mc2/(Zitterbewegung). The nonrelativistic
Schr¨
odinger equation can be retrieved by expressing ψas the
modulation of a nonrelativistic, slowly varying wave function
(x,t) by this vibration at the Compton frequency:
ψ(x,t)=eimc2
t(x,t).(30)
042202-8
QUANTUMLIKE STATISTICS OF DETERMINISTIC WAVE- . . . PHYSICAL REVIEW E 93, 042202 (2016)
Substitution in Eq. (29) indeed yields
it+
2
2m2=
2
2mc2tt. (31)
The left-hand side is identical to Eq. (28), so the right-hand
side is a relativistic correction. The characteristic frequency of
(x,t)isω=k2/(2m). Therefore, the relativistic correction
is negligible when ω/k c, or equivalently when c1,
where λc=2π/(mc) is the Compton wavelength.
Couder and coworkers have already observed in many dif-
ferent configurations [17,20,21] that the Faraday wavelength
λFwas the walker equivalent of the de Broglie wavelength
λdB =2π/k of quantum particles (Analogy No. 1). In a recent
review paper, Bush [1] pushed the analogy one step further and
hypothesized that the bouncing motion of walkers could be the
equivalent of this Zitterbewegung, so the bouncing frequency
fwould correspond to the Compton frequency mc2/(2π)
(Analogy No. 2).
The diffusive behavior of the walker on the long term
suggests a third analogy (No. 3), between the walker’s
diffusion coefficient Dand the coefficient of Schr¨
odinger
equation /m. It is equivalent to say that the Schr¨
odinger
timescale 2π/ω =4πm/(k2) is analogous to the dimensional
coherence time of the walker λ2
F/(πD)[4/(πb2)]Ncf1
3.3Ncf1. De Broglie’s momentum relation states that the
speed of a nonrelativistic particle is v=k/m, which is
then equivalent to [πb2/2]λFfN1
c0.60λFfN1
cfor the
walker. This value is remarkably close to the ballistic speed
Ff/Ncwith a=0.57 as observed in Fig. 8. The analog of
the speed of light cis [bπ/2]λFfN1/2
c=0.78λFfN1/2
c,
which does not seem to be a fundamental constant in the
walker’s world. Nevertheless, the relativistic limit vc
corresponds to Ncπb2/20.60 for the walker, i.e., the
coherence time becomes of the order of the bouncing period.
Saying that a quantum particle cannot go faster than light
is then analogous to saying that the coherence time of a
walker trajectory should be at least one rebound time. When
Nc=πb2/20.60, the equivalent speed of light becomes
λFf, which is the phase speed of capillary waves. Finally,
in quantum mechanics the correspondence principle states
that classical mechanics is recovered in the limit 0. This
translates into the walker behavior being fully ballistic when
Dgoes to zero, i.e., when the coherence timescale Ncgoes
to infinity. The three equivalences and their implications are
summarized in Table III.
B. Probability density
In a 2D circular infinite potential well of radius Rc,thewave
function can be decomposed into a discrete basis of cylindrical
harmonics φk(r) defined in Eq. (5):
=1
Rc
k
ckφk(r)ekt,(32)
where (R2
c2+z2
k)φk=0 and
ωk=
2mR2
c
z2
k(33)
in order to satisfy the Schr¨
odinger equation. The probability
density function (PDF) is then
ρ(r,t)=
j,k
c
jckφ
j(r)φk(r)ei(ωkωj)t.(34)
Coefficients ckmust satisfy SρdS =k|ck|2=1. Since all
zkare distinct (there is no degeneracy), the time-averaged
probability density is
ρ(r)=
k|ckφk(r)|2=
k|ck|2ϕk(r)2,(35)
which is necessarily axisymmetric. Although all zeros of
Bessel functions (and derivatives) are strictly distinct, it is
always theoretically possible to find and select two eigenmodes
(j,k)forwhich|zjzk|is arbitrarily small. The corre-
sponding imaginary exponential in Eq. (34) would generate
some beating at the extremely low frequency (ωjωk) that
would challenge any practical computation of a time average.
Nevertheless, as far as the modes selected in this work are
concerned (Table II), |zjzk|>0.01 for any j= k.
How well are the walker statistics described by these
quantum predictions? Following de Broglie’s pilot-wave the-
ory [38,39], we assume an analogy between the quantum wave
function and the classical wave field of the walker, averaged
over the coherent timescale. In Sec. IV, the walker dynamics
was considered in configurations where a small number of
modes were equally excited while others were strictly not.
This suggests the identification of quantum coefficients ckas
ck=ξk
kξ2
k
,(36)
where again ξk=1ifk=0 and ξk=2 otherwise. As seen in
Fig. 7(a), the corresponding quantum prediction of the average
probability density [Eq. (35)] is reminiscent but not identical
to the one obtained from walker simulations. Nevertheless,
their extrema of density coincide almost perfectly.
C. Average kinetic energy
In quantum mechanics, the kinetic energy operator ˆ
T=
2
2m2is Hermitian, and its expected value
T=|ˆ
T|=−
2
2mS
2dS =
2
2mR2
c
k|ck|2z2
k
(37)
is time-independent in a infinite potential well. A dimension-
less kinetic energy can then be defined based on the cavity
radius Rcand Compton frequency ωc/(2π):
T=2
mT4π2
R2
cω2
c=B2
k|ck|2z2
k,(38)
where B=2π/(mR2
cωc) is a dimensionless coefficient.
According to Table III,Bis equivalent to the dimensionless
diffusion coefficient ˜
Dof the walker, defined in Eqs. (26)
and (27). This quantum prediction is in remarkable agreement
with the observed average kinetic energy of simulated walkers,
for all considered mode subsets (Fig. 11). Only simulations at
large Cfail to match the prediction, possibly because they
042202-9
TRISTAN GILET PHYSICAL REVIEW E 93, 042202 (2016)
TABLE III. Correspondence of variables between walkers and quantum particles (b0.62).
Analogy Walker Quantum particle
No. 1 Faraday wavelength λFde Broglie wavelength 2π/k
No. 2 Bouncing frequency fZitterbewegung frequency mc2/(2π)
No. 3 Diffusion coefficient D=[b2/4]λ2
FfN1
cSchr¨
odinger diffusion /m
Derived Coherence time [4/(πb2)]Ncf1Schr¨
odinger timescale 4πm/(k2)
Ballistic speed v=[πb2/2]λFfN1
cQuantum speed v=k/m
Maximum speed [bπ/2]λFfN1/2
cSpeed of light c
Relativistic limit Ncb
2/2 Relativistic limit v/c < 1
Ballistic limit Nc→∞ Classical limit 0
would be in the relativistic regime (Nc<5). The energy of
the mixed-mode subset S6is also overestimated. The kinetic
energy of two interacting walkers was shown to be somehow
equivalent to the energy stored in the corresponding wave
field [40]. Nevertheless, as detailed in Appendix B,itishere
unclear if both energies are still equivalent for a single walker
confined in a cavity.
D. Position-dependent statistics
Radial- and azimuthal-velocity operators are defined by
projecting the momentum operator ˆ
P=−ialong the radial
and azimuthal directions, respectively:
ˆ
Vr=1
mer·ˆ
P=−i
mRc
r,
ˆ
Vθ=1
meθ·ˆ
P=−i
mRc
1
rθ.(39)
On the one hand, the radial-velocity operator is not Hermitian,
so it should not be observable. On the other hand, the
tangential-velocity operator has a trivial expected value of
zero, by symmetry. The Heisenberg uncertainty principle states
that one cannot measure accurately and simultaneously both
T
q
uantum
10-6 10-4 10-2 100
Twalker
10-6
10-4
10-2
100
FIG. 11. Average dimensionless kinetic energy
Tof the walker,
vs quantum prediction, with the equivalence B˜
D. Symbols
correspond to different mode subsets (Table II). The solid line is
the quantum prediction [Eq. (38)].
the position and momentum of a quantum particle. However,
it is possible to “weakly” measure a quantum particle, gaining
some information about its momentum without appreciably
disturbing it, so its position can be “strongly” measured
directly after [41]. The information obtained from individual
measurements is limited. But one can perform many trials,
then postselect particles that were observed at a given position
and calculate their associated average momentum. We here
propose to extend this concept of weak measurement to a
particle in a 2D circular cavity. We define the Hermitian
operators
ˆ
V2
r=ˆ
V
r
ˆ
δ(rR)
2πR
ˆ
Vrand ˆ
V2
θ=ˆ
V
θ
ˆ
δ(rR)
2πR
ˆ
Vθ,(40)
which are aimed to represent the squared radial and azimuthal
velocities at a given radial position R, respectively. Their
expected values
|ˆ
V2
r|=
mRc2S
rrδ(rR)
2πR dS,
(41)
|ˆ
V2
θ|=
mRc2S
θθδ(rR)
2πR r2dS,
correspond to the variance of each velocity component. They
are time-averaged,
|ˆ
V2
r|=
mRc2
k|ck|2[rϕk]2
r=R,(42)
|ˆ
V2
θ|=
mRc2
k|ck|2k2[ϕk]2
r=R,(43)
and made dimensionless,
v2
r= |ˆ
V2
r|4π2
R2
cω2
c=B2
k|ck|2[rϕk]2
r=R,
(44)
v2
θ= |ˆ
V2
θ|4π2
R2
cω2
c=B2
k|ck|2k2[ϕk]2
r=R.
Again, the equivalence between Band ˜
Dallows for a
direct comparison between walkers and quantum predictions.
The evolution of
v2
rwith Ris very similar in both worlds,
although the proportionality coefficient between both (best fit)
varies from one subset to another (Fig. 12). The agreement
is even better for
v2
θ, where the walker variance is almost
exactly twice the quantum prediction, at any radial position,
for most values of C(Fig. 13). This remarkable similarity
042202-10
QUANTUMLIKE STATISTICS OF DETERMINISTIC WAVE- . . . PHYSICAL REVIEW E 93, 042202 (2016)
R
0 0.2 0.4 0.6 0.8 1
v2
r/˜
D2
0
100
200
300
400
500
600
FIG. 12. Average variance of the radial velocity
v2
rof the walker at
a given radial position r, normalized by ˜
D2, for subset S2(Table II).
The solid line corresponds to 3.16 times the quantum prediction
[Eq. (44); best fit].
holds for any of the markedly different functions
v2
θ(R)at
each mode subset considered (Fig. 14). However, the quantum
calculation strongly underestimates both walker variances at
large C(relativistic regime).
These results demonstrate how much the velocity statistics
of the walker are shaped by the wave function in almost the
same way as the statistics of a quantum particle would be.
Moreover, they lead to an interpretation of the Heisenberg
uncertainty principle for walkers (slightly different from the
one proposed in Ref. [17]). The uncertainty in position can be
related to the coherence length and it is then of the order of
λF/2 (equivalently λdB/2). The uncertainty in speed is directly
R
0 0.2 0.4 0.6 0.8 1
v2
θ/˜
D2
0
20
40
60
80
100
120
140
160
180
200
FIG. 13. Average variance of the tangential velocity
v2
θof the
walker at a given radial position r, normalized by ˜
D2, for subset
S2(Table II). The solid line corresponds to 2.0 times the quantum
prediction [Eq. (44); best fit].
R
0 0.2 0.4 0.6 0.8 1
0
100
200
300
0
5000
10000
0
500
1000
0
50
100
150
FIG. 14. Variance of the tangential velocity
v2
θof the walker at
a given radial position r, normalized by ˜
D2, for subsets S1,S4,S5,
and S6, from top to bottom (Table II). The solid lines correspond
respectively to 1.0, 2.1, 2.2, and 2.0 times the quantum prediction
associated with each subset [Eq. (44); best fit].
given by the dimensional version of the standard deviation v,
which scales as D/λF[equivalently /(dB)]. The product
of these uncertainties then scales as the diffusion coefficient
D, which is equivalent to /m. So the uncertainty principle is
recovered for walkers provided their dynamics is analyzed at
the scale of diffusion.
VI. DISCUSSION
The theoretical framework introduced in Sec. II is one
of the simplest mathematical representations of a particle
042202-11
TRISTAN GILET PHYSICAL REVIEW E 93, 042202 (2016)
coupled to a wave. It qualitatively captures the key features
observed in the experiments on confined walkers performed
by Harris et al. [24]. It reproduces both the circular orbits at
low memory and the chaotic trajectories at high memory. The
current limitation for a quantitative comparison originates in
the lack of experimental information about the damping rate
of each eigenmode. Nevertheless, this model allows for the
exploration of regimes that are not accessible experimentally.
Of particular interest is the possibility to set the memory
of each mode to either infinity (no damping) or zero (full
damping). Quantumlike behavior of individual walkers were
observed at high-memory, when the system was as close to
conservative as it can be. In this model, the mode selection
around one given wavelength can be seen as analogous to the
preparation of a quantum state with a more-or-less defined
momentum. Nevertheless, it must be noted that this walker
model is still highly dissipative since the nonselected modes
(the ones that do not resonate with the forcing) are immediately
damped.
The chaotic trajectories of confined walkers are ballistic
on the short term and diffusive on the long term. The
coherence distance, beyond which the ballistic behavior is
lost, corresponds to half the Faraday wavelength, as a careful
qualitative look at the experimental data [24] also confirms.
The Faraday wavelength is at the heart of most quantumlike
behaviors of walkers; it is identified as equivalent to the de
Broglie wavelength for a quantum particle. Our analysis of
the diffusive motion in a circular corral has suggested another
equivalence, between the diffusion coefficient Dand the factor
/m in the Schr¨
odinger equation. The walker behavior thus
becomes apparently random only when it is analyzed at a
length scale larger than λF/2. This analogy is confirmed by
the observed ballistic speed of the walker, which corresponds
closely to the de Broglie speed k/m. Similarly, the average
kinetic energy of the walker matches Schr¨
odinger’s prediction
over several orders of magnitude.
The bouncing dynamics of the walkers was previously
hypothesized as reminiscent of the Zitterbewegung of quantum
particles [1]. From there, we found an analog for the speed of
light in the walker’s world, and we then identified the condition
for observing relativistic effects on walkers: they have to
travel a distance comparable to half the Faraday wavelength
at every rebound. Our mathematical framework allows for
a future investigation of this regime, which is unfortunately
not attainable in experiments where the walking steps are
usually limited to around λF/20 [6,24]. Equivalence relations
of Table III do not explicitly depend on the dispersion relation
of the waves. Nevertheless, the variables therein (such as the
coherence time or the elementary diffusion coefficient) do
depend on the considered wave-particle interaction, e.g., here
through the coupling constant C[Eq. (23)]. This might be the
reason why the analogs for the speed of light and the Planck
constant are not constant in the walker’s world.
Harris et al. [24] observed that the statistics of confined
walkers can be shaped by the cavity eigenmodes in the
high-memory limit. We have here calculated the position-
dependent variance of walker velocity in the limit of some
modes having an infinite memory. Their dependence on radial
position is remarkably close to the predictions obtained from
the quantum formalism (linear Schr¨
odinger equation) with
Hermitian observables, even when complex combinations of
modes are considered.
This model of walkers is reminiscent of the pilot-wave
theories of de Broglie and Bohm, although there are some
significant differences [1]. In the Bohmian mechanical de-
scription of quantum mechanics, particles are guided by
a pilot-wave prescribed by Schr¨
odinger’s equation. It thus
evolves at the Schr¨
odinger timescale. The particles do not
exert any direct individual feedback on this wave; only their
statistics shapes the wave. By contrast, individual bouncing
walkers locally excite the wave field that sets them into motion.
The double-solution theory of de Broglie [38]involvesan
additional pilot-wave centered on the particle, whose timescale
would be the Zitterbewegung period. This second wave could
be the analog of the real Faraday wave that couples with
individual walkers. It was already shown that the Schr¨
odinger
equation can be retrieved from a random walk of diffusion
coefficient /(2m)[42]. This work suggests that such a random
walk can originate from the chaos of a deterministic map that
describes the coupling of a wave and a particle. In other words,
the solution of Schr¨
odinger equation for a particle in a cavity
can be obtained from a purely deterministic mechanism that
does not involve any stochastic element.
Walkers have now been investigated for a decade. They have
shown many behaviors reminiscent of quantum particles. We
have shown here that, when they are confined in cavities, their
statistics closely approaches the solution of the Schr¨
odinger
equation. Future work is still required to identify the exact
limits of this analogy.
ACKNOWLEDGMENTS
This research was performed in the framework of the
Quandrops project, financially supported by the Actions de
Recherches Concertees (ARC) of the Federation Wallonie-
Bruxelles under Contract No. 12-17/02. This research has also
been funded by the Interuniversity Attraction Poles Program
(IAP 7/38 MicroMAST) initiated by the Belgian Science
Policy Office. T.G. thanks J. W. M. Bush, D. Harris, A. Oza, F.
Blanchette, R. Rosales, M. Biamonte, M. Labousse, S. Perrard,
E. Fort, Y. Couder, N. Sampara, L. Tadrist, W. Struyve, R.
Dubertrand, J.-B. Shim, M. Hubert, and P. Schlagheck for
fruitful discussions.
APPENDIX A: STABILITY OF FIXED POINTS
AND ORBITS
1. Finite memory
Fixed points of the iterated map (14) satisfy rn=r0and
wk,n =wk0. Because of axisymmetry, θncan take any constant
value, so the locus of these fixed points is a series of concentric
circles, referred to here as fixed lines. The iterated map then
becomes
wk0=μk
1μk
ϕk0R,
(A1)
k
μk
1μk
ϕk0ϕ
k0=0,
042202-12
QUANTUMLIKE STATISTICS OF DETERMINISTIC WAVE- . . . PHYSICAL REVIEW E 93, 042202 (2016)
where ϕk0=ϕk(r0) and ϕ
k0=[k/dr]r0. This second condi-
tion can be written (r0)=0 with
(r)=
k
μk
1μk
ϕ2
k(r)
2.(A2)
The stability of these solutions can be inferred from a
linearized version of the map for small perturbations: rn=
r0+˜
rn,θn=˜
θn,wkn =wk0+˜wkn, where ˜
xn1, ˜
θn1,
and ˜wkn 1. We also decompose ˜wkn =˜
ukn +i˜vkn .The
linearized map is then
˜
rn+1=˜
rnC
k
[ϕ
k0wk0˜
rn+ϕ
k0˜
ukn],
˜
θn+1=˜
θn+C
r2
0
k
k0[kwk0˜
θn+˜vkn ],
(A3)
˜
uk,n+1=μk[˜
ukn +ϕ
k0˜
rn],
˜vk,n+1=μkvkn k0˜
θn].
Perturbations (˜
r,˜
uk) are decoupled from ( ˜
θ,˜vk) and can be
analyzed independently.
Radial perturbations ˜
rn=˜
r0znand ˜
ukn =˜
u0znmust satisfy
˜
uk0=μkϕ
k0/(zμk)˜
r0as well as
1z=C
k
μkϕk0ϕ
k0
1μk+ϕ2
k0
zμk.(A4)
When C1, all the solutions zshould be in the neighborhood
of z=1. If z=1, then
=C
1Ck
μk
(1μk)2ϕ2
k0
(r0),(A5)
where
(r0)=
k
μk
1μkϕk0ϕ
k0+ϕ2
k0.(A6)
Therefore, in the limit of small Cand finite damping factors
μk<1, radially stable (unstable) fixed points are found where
(r) is minimum (maximum).
Azimuthal perturbations ˜
θn=˜
θ0znand ˜vkn =˜vk0znmust
satisfy ˜vk0=−μkk0/(zμk)˜
θ0as well as
(z1)1C
r2
0
k
k2ϕ2
k0μk
(1μk)(zμk)=0.(A7)
Since z=1 is always a solution, these azimuthal perturbations
are never more than marginally stable. The other solution z
increases from zero as damping factors μkare increased (i.e.,
as the forcing amplitude is increased). Therefore, for each
fixed point r0, there is a finite threshold in forcing amplitude
for which the damping factors μksatisfy
k
k2ϕ2
k0μk
(1μk)2=r2
0
C.(A8)
Above this threshold, there is at least one solution zlarger than
unity, so the corresponding fixed point becomes azimuthally
unstable. This corresponds to the walking threshold.
This azimuthal destabilization gives rise to periodic solu-
tions of the map (14) where the particle orbits at constant
speed around the center of the cavity: rn=r1,θn=αn and
wk,n =wk1eikαn (the wave pattern also rotates). The walking
velocity is then r1α. Plugging this solution in the iterated map
yields
wk1=μkϕk1
eikα μk
,
r1(1cos α)=C
k
ϕk1ϕ
k1μk
cos ()μk
12μkcos ()+μ2
k
,
r2
1sin α=C
k
2
k1μk
sin ()
12μkcos ()+μ2
k
,(A9)
where ϕk1=ϕk(r1) and ϕ
k1=[k/dr]r1. This system of
equations for (r1,α,wk1) can be solved numerically.
Slightly above the azimuthal destabilization threshold, the
orbital radius r1can be assumed to be as close to the fixed point
radius r0,sor1=r0+ε, with ε1. The angular velocity α
then satisfies
r0+C
k
μk(1+μk)
(1μk)3k2ϕk0ϕ
k0α2
2=C(r0)ε. (A10)
Since ε0 when α0, orbital solutions do originate from
the azimuthal destabilization of fixed points, here through a
pitchfork bifurcation. Each orbit then directly inherits from
the radial stability of its corresponding fixed point.
Orbit stability can be inferred from a perturbation anal-
ysis rn=r1+˜
rn,θn=+˜
θn,wk,n =(wk1+˜wk,n)eikαn,
where ˜
rn1, ˜
θn1, ˜wk,n 1, and wk1=uk1+ivk1.The
linearized map is then
˜
rn+1cos αr1sin α(˜
θn+1˜
θn)=˜
rnC
k
[ϕ
k1uk1˜
rn
k1vk1˜
θn+ϕ
k1˜
uk,n],
˜
rn+1sin α+r1cos α(˜
θn+1˜
θn)=C
r1
k
kϕ
k1vk1ϕk1vk1
r1˜
rn+k1uk1˜
θn+ϕk1˜vk,n,
˜
un+1,k =μk[cos ()(˜
uk,n +ϕ
k1˜
rn)sin ()vk,n k1˜
θn)],
˜vn+1,k =μk[cos ()vk,n k1˜
θn)+sin ()(˜
uk,n +ϕ
k1˜
rn)].
Perturbations along different directions are now coupled, and eigenvalues can be found numerically.
042202-13
TRISTAN GILET PHYSICAL REVIEW E 93, 042202 (2016)
2. Fixed points and orbits in the case of infinite memory
Fixed points do not exist anymore at infinite memory, at
least as soon as more than one mode is selected. Indeed,
they would require ϕk(r0)=0 simultaneously for all selected
modes k. We can then look for circular orbits rn=r1,θn=αn,
wkn =wk1eikαn. The wave equation imposes
wk1=ϕk1
eikα 1,(A11)
which is finite provided that k>0 (i.e., that no purely radial
mode is selected). Then
uk1=−ϕk1
2,v
k1=ϕk1
2
sin ()
1cos (),
r1(1cos α)=−C
2
k
ϕk1ϕ
k1,(A12)
r2
1sin α=C
2
k
2
k1
sin ()
1cos ().
In the limit ()21,
uk1=−ϕk1
2,v
k1=ϕk1
,
(A13)
21+
1r1=0,r
2
1α2=C1,
with
(r)=
k
ϕk(r)2,
(r1)=1,d
dr r=r1=
1.
When the mode selection includes is at least one radial mode
(k=0), the orbit solution here above is not valid anymore,
because vk1blows up when k=0. In the iterated map, the
radial wave mode satisfies
w0,n+1=w0,n +ϕ0(rn),(A14)
so the only way to avoid having this component blow up is
to impose ϕ0(r1)=ϕ01 =0, which selects the orbital radii but
leaves u01 and v01 undetermined. Moreover, if several radial
modes coexist, there should not be any orbital solution.
The other wave components (k>0) still satisfy
uk1=−ϕk1
2,v
k1=ϕk1
2
sin ()
1cos ().(A15)
Then the particle equations become
r1(1cos α)=Cϕ
01u01 +2
k>0
ϕ
k1uk1,
r2
1sin α=2C
k>0
k1vk1.(A16)
Again, these equations can be solved numerically to find the
orbital radii r1and their corresponding α.
APPENDIX B: WAVE ENERGY
In a recent paper, Borghesi et al. [40] observed an
equivalence between the kinetic energy of the walker and the
energy stored in the wave field, although the former was one
order of magnitude smaller than the latter. The dimensional
ballistic speed of the walker is here given by vw0.6λFf/Nc,
from which we infer the kinetic energy of the walker:
T=2π
3R3
dv2
w0.75χρ R3
dδ
R2
c
f2,(B1)
where Rdis the droplet radius. The time-averaged energy of a
mono-frequency wave field is given by
Ewave πρf 2λSH2
ndS =2π2ρf 2λ2
R2
c1
0h2
nrdr,
(B2)
where we check numerically for each subset Sthat
2π1
0h2
nrdr =
k|wk|2χ2
F
8C.(B3)
Therefore, the ratio between both energies is
T
Ewave 6
π
R3
dδ2
λ3
FR2
c
.(B4)
In this model, δis not directly expressed as a function of
other parameters, so it is unfortunately hard to conclude here
if both energies are equivalent. The estimation of δfrom more
advanced bouncing and walking models [33] is left to future
work.
[1] J.W.M.Bush,Annu. Rev. Fluid Mech. 47,269 (2015).
[2] J.W.M.Bush,Phys. Today 68(8), 47 (2015).
[3] M. Faraday, Philos. Trans. R. Soc. London 121, 319 (1831).
[4] T. Benjamin and F. Ursell, Proc. R. Soc. London, Ser. A 225,
505 (1954).
[5] K. Kumar and L. Tuckerman, J. Fluid Mech. 279,49 (1994).
[6] A. Eddi, E. Sultan, J. Moukhtar, E. Fort, M. Rossi, and Y. Couder,
J. Fluid Mech. 674,433 (2011).
[7] O. Jayaratne and B. Mason, Proc. R. Soc. London, Ser. A 280,
545 (1964).
[8] T. Gilet and J. W. M. Bush, J. Fluid Mech. 625,167 (2009).
[9] T. Gilet and J. W. M. Bush, Phys. Fluids 24,122103 (2012).
[10] Y. Couder, E. Fort, C.-H. Gautier, and A. Boudaoud, Phys. Rev.
Lett. 94,177801 (2005).
[11] T. Gilet, D. Terwagne, N. Vandewalle, and S. Dorbolo, Phys.
Rev. Lett. 100,167802 (2008).
[12] T. Gilet and J. W. M. Bush, Phys.Rev.Lett.102,014501 (2009).
[13] J. Molacek and J. W. M. Bush, J. Fluid Mech. 727,582 (2013).
[14] M. Hubert, D. Robert, H. Caps, S. Dorbolo, and N. Vandewalle,
Phys.Rev.E91,023017 (2015).
[15] Y. Couder, S. Protiere, E. Fort, and A. Boudaoud, Nature
(London) 437,208 (2005).
[16] A. Eddi, E. Fort, F. Moisy, and Y. Couder, Phys.Rev.Lett.102,
240401 (2009).
042202-14
QUANTUMLIKE STATISTICS OF DETERMINISTIC WAVE- . . . PHYSICAL REVIEW E 93, 042202 (2016)
[17] Y. Couder and E. Fort, Phys. Rev. Lett. 97,154101
(2006).
[18] C. Richardson, P. Schlagheck, J. Martin, N. Vandewalle, and
T. Bastin, arXiv:1410.1373.
[19] A. Andersen, J. Madsen, C. Reichelt, S. Rosenlund Ahl,
B. Lautrup, C. Ellegaard, M. T. Levinsen, and T. Bohr, Phys.
Rev. E 92,013006 (2015).
[20] S. Perrard, M. Labousse, M. Miskin, E. Fort, and Y. Couder,
Nat. Commun. 5,3219 (2014).
[21] E. Fort, A. Eddi, A. Boudaoud, J. Moukhtar, and Coude, Proc.
Natl. Acad. Sci. USA 107,17515 (2010).
[22] D. M. Harris and J. W. M. Bush, J. Fluid Mech. 739,444 (2014).
[23] A. U. Oza, D. M. Harris, R. R. Rosales, and J. W. M. Bush, J.
Fluid Mech. 744,404 (2014).
[24] D. M. Harris, J. Moukhtar, E. Fort, Y. Couder, and J. W. M.
Bush, Phys. Rev. E 88,011001(R) (2013).
[25] D. Shirokoff, Chaos 23,013115 (2013).
[26] B. Filoux, M. Hubert, and N. Vandewalle, Phys.Rev.E92,
041004(R) (2015).
[27] S. Perrard, M. Labousse, E. Fort, and Y. Couder, Phys. Rev. Lett.
113,104101 (2014).
[28] M. Labousse, S. Perrard, Y. Couder, and E. Fort, New J. Phys.
16,113027 (2014).
[29] A. U. Oza, R. R. Rosales, and J. W. M. Bush, J. Fluid Mech.
737,552 (2013).
[30] T. Gilet, Phys. Rev. E 90,052917 (2014).
[31] P. A. Milewski, C. A. Galeano-Rios, A. Nachbin, and J. W. M.
Bush, J. Fluid Mech. 778,361 (2015).
[32] J. W. M. Bush, A. U. Oza, and J. Molacek, J. Fluid Mech. 755,
R7 (2014).
[33] J. Molacek and J. W. M. Bush, J. Fluid Mech. 727,612 (2013).
[34] W. Batson, F. Zoueshtiagh, and R. Narayanan, J. Fluid Mech.
729,496 (2013).
[35] F. Blanchette, Phys. Fluids 28,032104 (2016).
[36] A. Eddi, D. Terwagne, E. Fort, and Y. Couder, Europhys. Lett.
82,44001 (2008).
[37] D. M. Harris and J. W. M. Bush, J. Sound Vib. 334,255 (2015).
[38] L. de Broglie, Annales de la Fondation Louis de Broglie 12,1
(1987).
[39] L. de Broglie, On the Theory of Quanta, Annales de Physique,
10e serie, t.III (1925).
[40] C. Borghesi, J. Moukhtar, M. Labousse, A. Eddi, E. Fort, and
Y. Couder, Phys. Rev. E 90,063017 (2014).
[41] S. Kocsis, B. Braverman, S. Ravets, M. J. Stevens, R. P. Mirin,
L. K. Shalm, and A. M. Steinberg, Science 332,1170 (2011).
[42] E. Nelson, Phys. Rev. 150,1079 (1966).
042202-15
... Indeed, this work is inspired by fluid-mechanical systems that were discovered about a decade ago by Couder, Fort and collaborators [13][14][15], and since then investigated in great detail also by other 4 teams, notably by Bush and collaborators [16][17][18] and Gilet and collaborators [46][47]. These systems, oil droplets walking over a vibrating fluid film, exhibit a remarkable series of quantumlike features, all induced by a pilot-wave. ...
... Under precise experimental conditions the droplets rapidly bounce on the oil surface, while being propelled by the surface wave they create. With such 'walking' droplets an impressive series of experiments can be performed exhibiting properties that mimic quantum phenomena, such as single particle diffraction and interference, quantization of angular momentum (the droplets can only rotate on a discrete series of radiuses), a form of tunneling, etc. [13][14][15][16][17][18][46][47]. When the oil bath is rotated, two droplets attract each other via the surface wave they generate. ...
... Moreover these experimental analogies are backed-up by certain intriguing formal analogies 1 (e.g. [47]). For our purpose, we do not need to take into account any of the detailed mechanisms involved (for in-depth Newtonian modeling of the system, see e.g. ...
Preprint
Recently it was shown that certain fluid-mechanical 'pilot-wave' systems can strikingly mimic a range of quantum properties, including single particle diffraction and interference, quantization of angular momentum etc. How far does this analogy go? The ultimate test of (apparent) quantumness of such systems is a Bell-test. Here the premises of the Bell inequality are re-investigated for particles accompanied by a pilot-wave, or more generally by a resonant 'background' field. We find that two of these premises, namely outcome independence and measurement independence, may not be generally valid when such a background is present. Under this assumption the Bell inequality is possibly (but not necessarily) violated. A class of hydrodynamic Bell experiments is proposed that could test this claim. Such a Bell test on fluid systems could provide a wealth of new insights on the different loopholes for Bell's theorem. Finally, it is shown that certain properties of background-based theories can be illustrated in Ising spin-lattices.
... A two-dimensional billiards -type model was developed by Shirokoff [5], then Gilet [6] presented perhaps the simplest dynamical model, which considers straight-line walking in a semi-confined geometry. Later, Gilet [7] derived a more complex, in both the philosophical and mathematical sense, discrete model for walkers on a circular corral. ...
... Like the standard map [2,3] it is assumed that the droplet receives a "kick" at each impact n. Other assumptions include, a constant damping, C ∈ [0, 1], of the velocity due to the inertia of the droplet [5,6,7], and a constant undamped kick strength of K ∈ R + . Furthermore, we use nondimensional variables and parameters from the onset. ...
Preprint
Recent experiments on walking droplets in an annular cavity showed the existence of complex dynamics including chaotically changing velocity. This article presents models, influenced by the kicked rotator/standard map, for both single and multiple droplets. The models are shown to achieve both qualitative and quantitative agreement with the experiments, and makes predictions about heretofore unobserved behavior. Using dynamical systems techniques and bifurcation theory, the single droplet model is analyzed to prove dynamics suggested by the numerical simulations.
... Over the past two decades, a number of experiments have demonstrated that a millimetric oil droplet bouncing on the surface of a vertically vibrating fluid bath exhibits certain phenomena previously thought to be exclusive to the microscopic quantum realm [6,7]. These include quantized orbital radius [3,27,39] and angular momentum [18,68], wavelike statistics in cavities [31,40,71], tunneling [22,43,58,76], Friedel oscillations [72] and Anderson localization [1], among others [23,28,29,30,57,65,81]. The droplet bounces on the surface of the bath and thus generates a guiding or 'pilot' wave field. ...
... While the walker's dynamics may be chaotic, its statistics often displays wave-like patterns reminiscent of those exhibited by quantum particles. Specifically, experiments and numerical simulations have suggested that the walker's long-time statistical behavior exhibits a wave-like signature in a rotating frame [39,64], harmonic potential [18,49,68] and corral geometry [15,31,40,70,71]. There has thus been an interest in understanding the long-time behavior of this hydrodynamic pilot-wave system. ...
Preprint
We conduct an analysis of a stochastic hydrodynamic pilot-wave theory, which is a Langevin equation with a memory kernel that describes the dynamics of a walking droplet (or "walker") subjected to a repulsive singular potential and random perturbations through additive Gaussian noise. Under suitable assumptions on the singularities, we show that the walker dynamics is exponentially attracted toward the unique invariant probability measure. The proof relies on a combination of the Lyapunov technique and an asymptotic coupling specifically tailored to our setting. We also present examples of invariant measures, as obtained from numerical simulations of the walker in two-dimensional Coulomb potentials. Our results extend previous work on the ergodicity of stochastic pilot-wave dynamics established for smooth confining potentials.
... Over the last two decades, a classical macroscopic system that couples a particle and a wave [12] has shown a number of analogies with quantum particles [13], including quantized orbital radius [14-16] and angular momentum [17,18], wavelike statistics in cavities [19][20][21], tunneling [22][23][24][25] and Friedel oscillations [26], among others [27][28][29][30][31][32][33]. In this system, the particle is a droplet coupled to the wave field that it generates by bouncing on the surface of a vibrating liquid bath [12,34,35]. ...
... Over the last two decades, a classical macroscopic system that couples a particle and a wave [12] has shown a number of analogies with quantum particles [13], including quantized orbital radius [14][15][16] and angular momentum [17,18], wavelike statistics in cavities [19][20][21], tunneling [22][23][24][25] and Friedel oscillations [26], among others [27][28][29][30][31][32][33]. In this system, the particle is a droplet coupled to the wave field that it generates by bouncing on the surface of a vibrating liquid bath [12,34,35]. ...
Preprint
Full-text available
A macroscopic hydrodynamic system that couples a particle and a wave has recently renewed interest in the question as to what extent a classical system may reproduce quantum phenomena. Here we investigate single-particle diffraction with a pilot-wave model originally developed to describe the hydrodynamic system. We study single-particle interactions with a barrier and slits of increasing width by focusing on the near field. We find single-particle diffraction arising as wavelike patterns in the particles' position statistics, which we compare to the predictions of quantum mechanics. We provide a mechanism that rationalizes the diffractive behavior in our system.
... (iii) There is memory in the system. Since the waves generated by the droplet decay very slowly in time, the motion of the droplet is not only influenced by the wave it generated on the most recent bounce, when the droplet is confined spatially by either applying an external potential or the presence of boundaries, one gets chaotic motion of a WPE that leads to coherent wave-like statistics [13][14][15][16]. These are some of the many hydrodynamic quantum analogs exhibited by the walking-droplet system (for a comprehensive review see [17,18]). ...
... It is clear from the previous analyses that there exists an intimate connection between chaotic dynamics and the statistics of tunneling behavior. Emergence of wave-like statistics from underlying chaotic dynamics has been observed in experiments and simulations with walking droplets confined in a small domain whose extent is a few times the wavelength of droplet-generated waves [10,[13][14][15][16]26,54]. Wavelike statistics emerge in these confined systems when (i) the decay time (i.e. ...
Article
Full-text available
A macroscopic, self-propelled wave-particle entity (WPE) that emerges as a walking droplet on the surface of a vibrating liquid bath exhibits several hydrodynamic quantum analogs. We explore the rich dynamical and quantum-like features emerging in a model of an idealized one-dimensional WPE in a double-well potential. The integro-differential equation of motion for the WPE transforms to a Lorenz-like system, which we explore in detail. We observe the analog of quantized eigenstates as discrete limit cycles that arise by varying the width of the double-well potential, and also in the form of multistability with coexisting limit cycles. These states show narrow as well as wide energy level splitting. Tunneling-like behavior is also observed where the WPE erratically transitions between the two wells of the double-well potential. We rationalize this phenomena in terms of crisis-induced intermittency. Further, we discover a fractal structure in the escape time distribution of the particle from a well based on initial conditions, indicating unpredictability of this tunneling-like intermittent behavior at all scales. The chaotic intermittent dynamics lead to wave-like emergent features in the probability distribution of particle's position that show qualitative similarity with its quantum counterpart. Lastly, rich dynamical features are also observed such as a period doubling route to chaos as well as self-similar periodic islands in the chaotic parameter set.
... Experimental [61] and numerical [22,43] studies of a walker in a harmonic potential similarly revealed that, in the long-time limit, the trajectories exhibit a quantization in their radius and angular momentum. Studies of a walker in circular [21,32,37,63] and elliptical [65] 'corral' geometries have shown that the longtime statistical behavior of the walker's position is related to the eigenmodes of the domain. ...
... We leave the detailed characterization of the invariant measure of (1.1) for future work. Moreover, we expect that results analogous to theorems 2.10 and 2.12 can be obtained for other hydrodynamic pilot-wave models that have been proposed in the literature, such as discrete-time (iterated map) models for droplets in the presence [24,31,32] or absence [22,30] of boundaries, and models for multiple interacting droplets [2,16,17,[69][70][71]. ...
Article
Full-text available
We study the long time statistics of a walker in a hydrodynamic pilot-wave system, which is a stochastic Langevin dynamics with an external potential and memory kernel. While prior experiments and numerical simulations have indicated that the system may reach a statistically steady state, its long-time behavior has not been studied rigorously. For a broad class of external potentials and pilot-wave forces, we construct the solutions as a dynamics evolving on suitable path spaces. Then, under the assumption that the pilot-wave force is dominated by the potential, we demonstrate that the walker possesses a unique statistical steady state. We conclude by presenting an example of such an invariant measure, as obtained from a numerical simulation of a walker in a harmonic potential.
... This active system is dynamically rich, with two key features -(i) the droplet and its self-generated wave coexist as a wave-particle entity (WPE); the droplet creates localized waves which in turn guide the droplet motion, and (ii) the system is non-Markovian and has memory since the waves generated by the droplet can decay very slowly in time and hence the droplet motion is affected by the history of waves along the droplet's trajectory. Such WPEs exhibit hydrodynamic quantum analogs [19][20][21][22] and some examples include quantization of orbits [23][24][25][26][27][28], tunneling across submerged barriers [29][30][31], wave-like statistics in confined geometries [32][33][34][35][36], Friedel oscillations [37] and hydrodynamic superradiance [38,39]. Most of the hydrodynamic quantum analogs have focused on a single WPE system, and investigations of many interacting WPEs have been limited [40][41][42][43][44]. ...
Preprint
Full-text available
Active particles are non-equilibrium entities that uptake energy and convert it into self-propulsion. A dynamically rich class of active particles having features of wave-particle coupling and memory are walking/superwalking droplets. Such classical, active wave-particle entities (WPEs) have been shown to exhibit hydrodynamic analogs of many single-particle quantum systems. We numerically investigate the dynamics of several WPEs and find that they self-organize into a bound cluster, akin to a nucleus. This active cluster exhibits various modes of collective excitations as the memory of the system is increased. Dynamically distinct excitation modes create a common time-averaged collective potential indicating an analogy with the bag model of a nucleus. At high memory, the active cluster can destabilize and eject WPEs whose decay statistics follow exponential laws analogous to radioactive nuclear decay. Hydrodynamic nuclear analogs open up new directions to pursue, both experimentally and numerically, within the nascent field of hydrodynamic quantum analogs.
... Walkers interact with submerged boundaries through non-specular reflection (Pucci et al. 2016) and tunnelling (Eddi et al. 2009;Carmigniani et al. 2014;Nachbin et al. 2017). More remarkably, when they are confined in a cavity, they experience chaotic trajectories for which the corresponding statistics is close to the quantum statistics predicted by the Schrödinger equation (Harris et al. 2013;Gilet 2014Gilet , 2016. They also seem to be diffracted when they pass one-by-one above a submerged slit (Couder & Fort 2006Dubertrand et al. 2016), although it is not yet clear that the underlying statistics is really similar to that of a diffracted quantum particle (Andersen et al. 2015). ...
Preprint
A walker is a fluid entity comprising a bouncing droplet coupled to the waves that it generates at the surface of a vibrated bath. Thanks to this coupling, walkers exhibit a series of wave-particle features formerly thought to be exclusive to the quantum realm. In this paper, we derive a model of the Faraday surface waves generated by an impact upon a vertically vibrated liquid surface. We then particularise this theoretical framework to the case of forcing slightly below the Faraday instability threshold. Among others, this theory yields a rationale for the dependence of the wave amplitude to the phase of impact, as well as the characteristic timescale and length scale of viscous damping. The theory is validated with experiments of bead impact on a vibrated bath. We finally discuss implications of these results for the analogy between walkers and quantum particles.
Preprint
Full-text available
We present a new hydrodynamic analogy of nonrelativistic quantum particles in potential wells. Similarities between a real variant of the Schr\"odinger equation and gravity-capillary shallow water waves are reported and analyzed. We show that when locally oscillating particles are guided by real wave gradients, particles may exhibit trajectories of alternating periodic or chaotic dynamics while increasing the wave potential. The particle probability distribution function of this analogy reveals the quantum statistics of the standard solutions of the Schr\"odinger equation and thus manifests as a classical deterministic interpretation of Born's rule. Finally, a classical mechanism for the transition between quasi-stationary states is proposed.
Article
Full-text available
Small drops bouncing across a vibrating liquid bath display many features reminiscent of quantum systems.
Article
Full-text available
A millimetric droplet bouncing on the surface of a vibrating fluid bath can self-propel by virtue of a resonant interaction with its own wave field. This system represents the first known example of a pilot-wave system of the form envisaged by Louis de Broglie in his double-solution pilot-wave theory. We here develop a fluid model of pilot-wave hydrodynamics by coupling recent models of the droplet's bouncing dynamics with a more realistic model of weakly viscous quasi-potential wave generation and evolution. The resulting model is the first to capture a number of features reported in experiment, including the rapid transient wave generated during impact, the Doppler effect and walker–walker interactions.
Article
Full-text available
It has recently been demonstrated that droplets walking on a vibrating fluid bath exhibit several features previously thought to be peculiar to the microscopic realm. The walker, consisting of a droplet plus its guiding wavefield, is a spatially extended object. We here examine the dependence of the walker mass and momentum on its velocity. Doing so indicates that, when the walker's time scale of acceleration is long relative to the wave decay time, its dynamics may be described in terms of the mechanics of a particle with a speed-dependent mass and a nonlinear drag force that drives it towards a fixed speed. Drawing an analogy with relativistic mechanics, we define a hydrodynamic boost factor for the walkers. This perspective provides a new rationale for the anomalous orbital radii reported in recent studies.
Article
Full-text available
Bouncing walking droplets possess fascinating properties due to their peculiar wave/particule interaction. In order to study such walkers in a 1d system, we considered the case of one or more droplets in an annular cavity. We show that, in this geometry, walking droplets form a string of synchronized bouncing droplets that share a common coherent wave propelling the group at a speed faster than single walkers. The formation of this coherent wave and the collective behavior of droplets is captured by a model.
Article
Full-text available
Yves Couder, Emmanuel Fort, and coworkers recently discovered that a millimetric droplet sustained on the surface of a vibrating fluid bath may self-propel through a resonant interaction with its own wave field. This article reviews experimental evidence indicating that the walking droplets exhibit certain features previously thought to be exclusive to the microscopic, quantum realm. It then reviews theoretical descriptions of this hydrodynamic pilot-wave system that yield insight into the origins of its quantumlike behavior. Quantization arises from the dynamic constraint imposed on the droplet by its pilot-wave field, and multimodal statistics appear to be a feature of chaotic pilot-wave dynamics. I attempt to assess the potential and limitations of this hydrodynamic system as a quantum analog. This fluid system is compared to quantum pilot-wave theories, shown to be markedly different from Bohmian mechanics and more closely related to de Broglie’s original conception of quantum dynamics, his double-solution theory, and its relatively recent extensions through researchers in stochastic electrodynamics.
Article
Full-text available
A bouncing droplet, self-propelled by its interaction with the waves it generates, forms a classical wave-particle association called a "walker." Previous works have demonstrated that the dynamics of a single walker is driven by its global surface wave field that retains information on its past trajectory. Here, we investigate the energy stored in this wave field for two coupled walkers and how it conveys an interaction between them. For this purpose, we characterize experimentally the "promenade modes" where two walkers are bound, and propagate together. Their possible binding distances take discrete values, and the velocity of the pair depends on their mutual binding. The mean parallel motion can be either rectilinear or oscillating. The experimental results are recovered analytically with a simple theoretical framework. A relation between the kinetic energy of the droplets and the total energy of the standing waves is established.
Article
Full-text available
A walker is a droplet bouncing on a liquid surface and propelled by the waves that it generates. This macroscopic wave-particle association exhibits behaviors reminiscent of quantum particles. This article presents a toy model of the coupling between a particle and a confined standing wave. The resulting two-dimensional iterated map captures many features of the walker dynamics observed in different configurations of confinement. These features include the time decomposition of the chaotic trajectory in quantized eigenstates and the particle statistics being shaped by the wave. It shows that deterministic wave-particle coupling expressed in its simplest form can account for some quantumlike behaviors.
Article
Full-text available
A bouncing drop and its associated accompanying wave forms a walker. Based on previous works, we show in this article that it is possible to formulate a simple theoretical framework for the walker dynamics. It relies on a time scale decomposition corresponding to the effects successively generated when the memory effects increase. While the short time scale effect is simply responsible for the walkerʼs propulsion, the intermediate scale generates spontaneously pivotal structures endowed with angular momentum. At an even larger memory scale, if the walker is spatially confined, the pivots become the building blocks of a self-organization into a global structure. This new theoretical framework is applied in the presence of an external harmonic potential, and reveals the underlying mechanisms leading to the emergence of the macroscopic spatial organization reported by Perrard et al (2014 Nature Commun. 5 3219).
Article
We present a first-principles model of drops bouncing on a liquid reservoir. We consider a nearly inviscid liquid reservoir and track the waves that develop in a bounded domain. Bouncing drops are modeled as vertical linear springs. We obtain an expression for the contact force between drop and liquid surface and a model where the only adjustable parameter is an effective viscosity used to describe the waves on the reservoir’s surface. With no adjustable parameters associated to the drop, we recover experimental bouncing times and restitution coefficients. We use our model to describe the effect of the Bond, Ohnesorge, and Weber numbers on drops bouncing on a stationary reservoir. We also use our model to describe drops bouncing on an oscillated reservoir, describing various bouncing modes and a walking threshold.
Article
In a thought-provoking paper, Couder and Fort [Phys. Rev. Lett. 97, 154101 (2006)10.1103/PhysRevLett.97.154101] describe a version of the famous double-slit experiment performed with droplets bouncing on a vertically vibrated fluid surface. In the experiment, an interference pattern in the single-particle statistics is found even though it is possible to determine unambiguously which slit the walking droplet passes. Here we argue, however, that the single-particle statistics in such an experiment will be fundamentally different from the single-particle statistics of quantum mechanics. Quantum mechanical interference takes place between different classical paths with precise amplitude and phase relations. In the double-slit experiment with walking droplets, these relations are lost since one of the paths is singled out by the droplet. To support our conclusions, we have carried out our own double-slit experiment, and our results, in particular the long and variable slit passage times of the droplets, cast strong doubt on the feasibility of the interference claimed by Couder and Fort. To understand theoretically the limitations of wave-driven particle systems as analogs to quantum mechanics, we introduce a Schrödinger equation with a source term originating from a localized particle that generates a wave while being simultaneously guided by it. We show that the ensuing particle-wave dynamics can capture some characteristics of quantum mechanics such as orbital quantization. However, the particle-wave dynamics can not reproduce quantum mechanics in general, and we show that the single-particle statistics for our model in a double-slit experiment with an additional splitter plate differs qualitatively from that of quantum mechanics.