Download as pdf or txt
Download as pdf or txt
You are on page 1of 25

European Journal of Physics

PAPER • OPEN ACCESS You may also like


- Worm-like instability of a vibrated sessile
A modular experimental system for teaching fluid drop
A. Hemmerle, G. Froehlicher, V. Bergeron
dynamics with Faraday waves et al.

- Effect of depth ratio on Faraday instability


in a binary liquid system
To cite this article: Henrik B Pedersen et al 2023 Eur. J. Phys. 44 065002 K P Choudhary, S P Das and Shaligram
Tiwari

- Droplet time crystals


Tapio Simula
View the article online for updates and enhancements.

This content was downloaded from IP address 177.37.250.13 on 21/03/2024 at 16:33


European Journal of Physics
Eur. J. Phys. 44 (2023) 065002 (24pp) https://doi.org/10.1088/1361-6404/acf1df

A modular experimental system for


teaching fluid dynamics with Faraday waves
Henrik B Pedersen , Albert Freud Abildgaard,
Morten Søtang Jacobsen and Henrik Juul
Department of Physics and Astronomy, Aarhus University, DK-8000 Aarhus C,
Denmark

E-mail: hbp@phys.au.dk

Received 2 April 2023, revised 12 July 2023


Accepted for publication 18 August 2023
Published 12 September 2023

Abstract
We describe a modular setup for the observation of Faraday waves on a
vibrating bath. The setup will be used as a project exercise on fluid dynamics
in a first-year course on experimental physics at Aarhus University as well as
for future research on fluids. As a demonstration of the setup, the acceleration
threshold for the onset of Faraday waves on a silicone oil bath as a function of
the driving frequency is measured and compared to thresholds calculated using
different existing models. The possibility to characterize surface waves with
the system is demonstrated by recording and analyzing images of Faraday
waves, e.g. showing explicitly that the Faraday waves in the present case are
subharmonic and establishing the dispersion relation for the waves.

Supplementary material for this article is available online

Keywords: Faraday waves, dispersion relation, image reconstruction, fluid


dynamics, vibrating bath

(Some figures may appear in colour only in the online journal)

1. Introduction

An interesting model system in fluid mechanics consists of a bath containing a fluid that is
vibrated vertically at a driving frequency fd and with an amplitude Ad, i.e. with a time-varying

Original content from this work may be used under the terms of the Creative Commons
Attribution 4.0 licence. Any further distribution of this work must maintain attribution to the
author(s) and the title of the work, journal citation and DOI.

© 2023 European Physical Society


0143-0807/23/065002+24$33.00 Printed in the UK 1
Eur. J. Phys. 44 (2023) 065002 H B Pedersen et al

Figure 1. Illustration of the model system in fluid dynamics for which a modular
experiment for university teaching is described. The relation of the frequency fF of the
Faraday wave and the driving frequency are indicated for a subharmonic wave. The
indicated coordinates for the description (x, y, z) are moving with the bath (accelerated
coordinate system).

acceleration amplitude of a = Ad ´ (2pfd )2 in addition to the constant gravitational accel-


eration g. The fluid is characterized by a density ρ, a kinematic viscosity ν (dynamic viscosity
η = ν × ρ), and a surface tension σ, and another fluid can be added to the bath or be the
atmospheric air (ρ2 ∼ 1.2 kg m−3, η2 ∼ 1.8 × 10−5 Pa s) defining an upper boundary of the
fluid. Figure 1 shows a graphical illustration of this physical model system. For a given
forcing frequency, the liquid surface becomes unstable once the applied acceleration exceeds
a threshold value, i.e. a ac( fd), which also depends on the properties (ρ, ν, and σ) of the
liquid as well as on the liquid height (h) in the bath (plus the fluid properties of a possible
additional fluid layer). Just above this threshold, the liquid surface displays Faraday waves
with wavelength λF and frequency fF = 1/2fd (subharmonic) or fF = fd (harmonic) and their
interdependence (dispersion relation) reflects the nature of the waves (gravity and capillary
waves) and potentially the nature of the interaction between the liquid and the walls of
the bath.
This model system allows for controlled studies of surface waves [1–5] and liquid surface
interactions [6–11] and as such it is a prototype system to examine fluid dynamics [12]. The
model system has a special role in the development of fluid mechanics as it in particular
allows isolated investigations of the instabilities of the fluid interfaces as pioneered by
Michael Faraday [13], and later was the exact model system investigated in the seminal
papers by Benjamin and Ursell [14], who provided a linear stability analysis without visc-
osity, and by Kumar and Tuckermann [12], who gave a full linear stability analysis including
viscosity. Several papers, e.g. [15–17] have reviewed both early and modern experimental
and theoretical studies on the Faraday instability. A full exploration of Faraday waves, i.e.
solving the Navier–Stokes equations with appropriate boundary conditions for the model
system depicted in figure 1, especially above threshold has not been realized analytically and
most likely new aspects of this system are still to be found experimentally [1] or through
computational investigations [17–19].

2
Eur. J. Phys. 44 (2023) 065002 H B Pedersen et al

Recently, vibrating baths have gained additional attention as it was discovered [20–23] that
a small droplet of the same liquid can bounce and walk on the oscillating surface in the region
below the acceleration threshold (a < ac) for generation of Faraday waves where the surface
is still stable. An oscillating liquid with a bouncing droplet is a fascinating dynamical system,
both as a research object and for teaching, which connects a particle, the droplet, and its wave
field, i.e. the surface wave it introduces on the liquid surface by bouncing. While the fluid
system is of course purely classical in nature, the droplet-liquid system in some sense
resembles the de Broglie-Bohmian pilot-wave view of quantum mechanics where a particle is
associated explicitly with a wave field, and several analogs to quantum theories and
experiments have been pursued with the droplet-liquid system [24–30]. Evidently, despite its
conceptual simplicity, the coupled motion of a droplet and its wave field is very challenging
to describe and its scientific exploration is still ongoing both experimentally and theoretically.
The experimental system with an oscillating bath (figure 1) thus forms the core of more
presently active research fields within fluid dynamics. The relative simplicity of the exper-
imental system and its fundamental role as an interesting model system in fluid mechanics
makes it attractive for teaching physics at undergraduate level, since direct illustrations of
essential fluid dynamical concepts (e.g. the Faraday instability and dispersion relations) can
be done quantitatively and, moreover, direct connections to fascinating modern research (e.g.
on bouncing and walking droplets) can be made without complicated instrumentation.
In this paper, we describe an experimental realization of a system for studying both surface
waves on the oscillating liquid as well as bouncing and walking droplets. The setup is made
modular with the possibility to add inserts to the bath, for example, to study the occurrence of
surface waves under different boundary conditions, e.g. straight walls or brimful conditions,
and droplet motion in designed structures. Further, the setup is realized with laboratory
equipment available or realizable at low cost in standard teaching laboratories at the uni-
versity. The setup is embedded as a project exercise in a course on experimental physics for
first-year students at Aarhus University and will also find future use on more advanced levels
of study and in student-driven research on fluid properties and droplet dynamics.
The focus in the present paper is on describing and characterizing the oscillating bath by
showing results on Faraday waves obtained using it, while results demonstrating its ability to
study bouncing and walking droplets will be described in a separate article.

2. Faraday waves

2.1. Qualitative outline of the description of the Faraday instability

For an inviscid liquid (η = 0), the Faraday instability was first described with a linear stability
analysis by Benjamin and Ursell [14] exploiting the fact that Mathieu equations for each
Fourier amplitude of normal modes of the surface waves can be formulated in this limit. For
viscous fluids, a complete linear stability analysis was given by Kumar and Tuckerman [12]
for two superimposed fluid layers and by Kumar [31] for one fluid layer. They demonstrated
how the acceleration threshold for the onset of Faraday waves at a given driving frequency
and at a given wavenumber can be evaluated by solving an eigenvalue problem, and they
demonstrated how, for a given driving frequency, the corresponding observable values of
acceleration threshold and wavenumber of the Faraday waves can be obtained. They pre-
sented a full hydrodynamical model (FHM) [12, 31] and a phenomenological model (PhM)
[12] in which an estimated representation of the viscous damping rate was used. In the limit of
weak dissipation, Müller et al [32] derived an explicit analytical model (AnM) for both the
acceleration threshold and the corresponding wavenumber at a selected driving frequency.

3
Eur. J. Phys. 44 (2023) 065002 H B Pedersen et al

The mathematical details of these evaluations [12, 31, 32] are somewhat complicated and
actual computations can potentially be difficult to perform for undergraduate students. As
supplementary information, we therefore provide example codes written for MATLAB [33]
from which such calculations can be done. Below, we give a brief outline of the main physical
arguments leading to the analysis of the Faraday instability and illustrate the properties of
Faraday waves by explicit calculations for selected cases with a fluid similar to the Silicone
oil used in the present work. We have decided to follow a notation similar to the one given in
the paper by Kumar [31].
The description of Faraday waves on the surface of an incompressible (i.e. ρ is constant)
and Newtonian (i.e. η is independent of stress and flow rate) fluid proceeds from the equation
of continuity (mass conservation)
 · u = 0, (1)
where u = (v, u, w) is the velocity field of the fluid at position x = (x, y, z), and the
corresponding Navier–Stokes equations for the fluid motion
¶u
r + (u · ) u =  · p = -p + h  2u - [g - a ´ cos (wd t )] e z (2)
¶t
where p is the local pressure in the fluid, and the components of the stress tensor π can be
written:

¶uj ¶uk ⎞
pjk = pd jk + h ⎛⎜ + ⎟ + r [g - a ´ cos (wd t )] zd jz dkz. ( 3)

⎝ kx ¶xj ⎠

with j and k representing x, y, and z, and ωd = 2π × fd.


The fluid motion can then be further defined by (1) imposing a no-slip condition at the
bottom of the bath, i.e. requiring the vertical fluid velocity to follow the velocity of the bath’s
bottom at all contact points, and (2) kinematic boundary conditions at the side walls, i.e.
requiring the fluid velocities perpendicular to the boundaries to vanish at the boundaries.
When considering a laterally infinite container [12, 31], the second condition becomes irre-
levant. At the surface, a kinematic surface condition, stating that fluid elements must remain
within the fluid or similarly that the surface is adverted by the fluid motion, must be imposed.
The lateral stability of the surface can be described by a dynamical boundary condition at the
surface, namely that the tangential components of the stress tensor (πxz and πyz, equation (3))
are continuous [12], or vanish [31] if we neglect the influence of the fluid above the surface.
Finally, the vertical destabilization of the surface, represented by the occurrence of Faraday
waves, can be represented by a discontinuity in the normal component (πzz) of the stress
tensor corresponding to the stress imposed by the curvature of the surface with surface
tension σ.
A small (infinitesimal) vertical oscillation (instability) of the position of the surface relative
to a flat interface can be described by normal modes of the horizontal plane with wave-
numbers kx and ky, e.g. for a surface of infinite lateral extension [12, 31] as
z (x, t ) = z (t ) ´ sin (k x x + k y y). (4)
and with a corresponding vertical velocity
w (x, t ) = w (t ) ´ sin (k x x + k y y). (5)
For standing waves in a rectangular geometry of dimension Lx × Ly the corresponding
description is

4
Eur. J. Phys. 44 (2023) 065002 H B Pedersen et al

z (x, t ) = z (t ) ´ sin (k x.lx ) sin (k y.my) (6)


where the wavenumbers kx.l and ky.m obey
p l
Lx = ´ n = x. n ´ n (7)
k x. l 2
and
p l y. m
Ly = ´m= ´m ( 8)
k y. m 2
where l and m are integers. For other geometries of the bath, e.g. a circular bath, the spatial
part of equation (4) as well as equations (7)–(8) will evidently be different.
The actual motion of the surface can then be represented by a Fourier expansion on the
spatial normal modes as
w (t ) = exp (s + iawd ) ´ å wn exp(inwd t ) ( 9)
n

and
z (t ) = exp (s + iawd ) ´ å zn exp(inwd t ) (10)
n

where s represents the viscous decay rate of the surface waves (∼ −2νk2) and α effectively
acts to modify the resulting frequency of the generated instability. For α = 1/2, subharmonic
waves with fF = 1/2fd, 3/2fd, 5/2fd... are described, while for α = 0 harmonic waves with
fF = fd, 2fd, 3fd, ... are described. Which type of waves, subharmonic or harmonic, and with
which wavenumber they will occur in the physical (experimental) situation, depend on the
actual acceleration threshold for the different types of waves. For inviscid fluids, the Fourier
amplitudes ζn each obey a differential equation of the Matheiu form [14].

2.2. Dispersion relation for free surface waves


Without forcing (a = 0 or ωd = 0), Kumar [31] established an analytical dispersion relation
for surface waves on a viscous fluid

s 3 4q k 2 (q02 + k 2) - C0 cosh (q0 h) cosh (kh) + D 0 sinh (q0 h) sinh (kh) ⎞


0 = gk + k - n 2 ⎛⎜ 0 ⎟
r q0 cosh (q0 h) sinh (kh) - k sinh (q0 h) cosh (kh)
⎝ ⎠
(11)
where the complex wavenumber q0 is defined by
s + iw 0
q02 = k 2 + (12)
n
and q0 is taken with the positive real part. The coefficients C0 and D0 are given by
C0 = q0 (q04 + 2q02 k 2 + 5k 4) (13)
and
D 0 = k (q04 + 6q02 k 2 + k 4). (14)
Requiring both real and imaginary parts of equation (11) to be zero simultaneously, the
angular frequency ω0 of free surface waves and their decay rate s can be obtained. In the limit
where ν ≈ 0, equation (11) reduces to an equation for the angular frequency only

5
Eur. J. Phys. 44 (2023) 065002 H B Pedersen et al

s
w 20 = ⎛gk + k 3⎞ tanh (kh)
⎜ ⎟ (15)
⎝ r ⎠

with two terms corresponding to free gravity waves (gk) and free capillary waves (σ/ρk3).

2.3. Dispersion relation and threshold acceleration for Faraday waves

Kumar and Tuckerman [12] and Kumar [31] devised a general method for evaluating the
threshold acceleration for a given driving frequency and wavenumber as an eigenvalue
problem. As mentioned, we will not repeat the derivation nor the explanation here, but we
give as supplementary information some example code where these explicit methods are
implemented.
Figure 2 illustrates the predicted acceleration thresholds for the occurrence of Faraday
waves from the full hydrodynamical model and the phenomenological model formulated by
Kumar et al [12, 31] for a liquid similar to the Silicone oil used in the experimental
investigation. For a given driving frequency, the wavenumber with lowest acceleration
threshold will evidently represent the observable Faraday wave. For a relatively deep liquid
(h = 5 mm), as displayed in figures 2(a)–(c), the subharmonic waves with fF = 1/2fd gen-
erally display the lowest acceleration thresholds over the entire frequency range and hence the
observable Faraday waves at threshold are subharmonic. For more shallow liquids of say
h = 1 mm, figure 2(d) exemplifies how harmonic waves with fF = fd or subharmonic waves
with fF = 3/2fd may display the lowest acceleration thresholds.
Thus, experimentally, Faraday waves will occur on the liquid surface at the threshold with
the wavenumber that displays the lowest acceleration threshold for forming the instability. In
this representation, the actual dispersion relation for Faraday waves is hence directly linked to
the analysis of the onset of the Faraday instability. Müller et al [32] gave analytical
expressions for the wavenumber and the acceleration threshold as a function of the driving
frequency for Faraday waves, valid in the limit of weak dissipation. Finally, assuming a
dispersion relation dominated by capillary waves, a surface wave decay rate of s = −2νk2,
and an infinitely deep liquid bath (h = ∞), a simple analytical form of the acceleration
threshold can be derived [9, 22, 34]

r 1 3 h
ac = 24 3 ´⎛ ⎞ ´ ⎛ ⎞ ´ (2pfd )5 3.
⎜ ⎟ (16)
⎝s ⎠ ⎝r ⎠
In figure 3, we show examples of calculated dispersion relations for free waves and for
Faraday waves as well as acceleration thresholds for Faraday waves for the Silicone oil used
in the experimental demonstration using different model descriptions and for two different
heights of liquids in the bath. As seen in figure 3(a), for the present liquid with a height of
h = 5 mm, the dispersion relation for the Faraday waves are almost identical to the dispersion
relation for free surface waves. For more shallow fillings of the bath, a small deviation
between the dispersion relation for free surface waves and for Faraday waves can be observed
around a wavelength of 15 nm.
Figures 3(c)–(d) illustrate the predictions from different models of the acceleration
threshold for a liquid height of h = 5 mm and h = 1 mm. The full hydrodynamical model
(FHM) [12, 31] represents the most reliable model as it relies only on the assumption of
infinitesimal waves (threshold). For a liquid height of h = 5 mm, the analytical model of
Müller et al [32] is in excellent agreement with the full hydrodynamical model (FHM)
[12, 31] for subharmonic waves while the phenomenological model (PhM) [12] deviates
significantly. Also for harmonic waves ( fF = fd), the deviation of the phenomenological

6
Eur. J. Phys. 44 (2023) 065002 H B Pedersen et al

Figure 2. Faraday acceleration thresholds as a function of wavenumber and for different


driving frequencies and different liquid heights in a laterally infinite bath as calculated
with the full hydrodynamical model formulated by Kumar and Tuckerman [12] and
Kumar [31]. The liquid parameters are ν = 2.0 m2 s−1, ρ = 938 m−3,
η = ν × ρ = 19 × 10−3 Pa s, and σ = 2.09 × 10−2 N m−1. The panels (a)–(f) show
acceleration thresholds for the onset of subharmonic (blue solid points) and harmonic
(red open circles) Faraday waves with liquid heights h and driving frequencies fd as
indicated. In panel (b), the regions of subharmonic ( fF = 1/2fd, 3/2fd, ...) and harmonic
( fF = fd, fd, ...) Faraday waves are indicated. The dashed lines in each panel mark the
corresponding values of acceleration threshold ac/g and wavenumber kF for
subharmonic and harmonic waves; symbols are displayed for the subharmonic waves
in panel (f). In panels (a)–(c) and (e)–(f) the subharmonic wave of fF = 1/2fd has the
lowest acceleration threshold. In panel (d), the harmonic wave with fF = fd and
subharmonic waves with fF = 3/2fd have almost the same acceleration threshold, while
the acceleration threshold for subharmonic waves with fF = 1/2fd is higher.

model and the analytical model relative to the full hydrodynamical model are significant. As
seen in figure 3(d), the discrepancies between the models are evident for all types of waves in
the case of a shallow liquid filling.

3. Experimental setup

3.1. Mechanical setup


Figure 4 shows a graphical illustration of the constructed experimental system with the
oscillating bath and the instrumentation used for the observations made in this study. As
supplementary information, we display in figure S1 more details on the actual construction of
the vibrating system, and in figure 5 we illustrate some details of the bath and give examples
of constructed inserts.
As illustrated in figure 4, the complete system is setup on a table with two vertically
displaced levels. The mechanical oscillator is mounted on the lower level (Level 1) with a

7
Eur. J. Phys. 44 (2023) 065002 H B Pedersen et al

Figure 3. Calculated properties of Faraday waves for a laterally infinite container with
liquid heights of (a)–(b) h = 5 mm and (c)–(d) h = 1 mm using different models
[12, 31, 32]. The liquid parameters are chosen to correspond approximately to the
Silicone oil used in the present experimental demonstration, i.e. ν = 2.0 m2 s−1,
ρ = 938 m−3, η = ν × ρ = 19 × 10−3 Pa s, and σ = 2.09 × 10−2 N m−1. In panels (a)
and (c) the dispersion relations for subharmonic Faraday waves as calculated from the
full hydrodynamical model of Kumar et al [12, 31] (see also figure 2) are compared to
free surface waves for an inviscid liquid (Equation (15)) and a viscous fluid
(equation (11)). In panels (b) and (d) different model calculations for the harmonic
( fF = fd) and subharmonic ( fF = 1/2fd) Faraday waves are compared. The models are
the full hydrodynamical model (FHM) of Kumar et al [12, 31], the phenomenological
model (PhM) of Kumar and Tuckerman [12], and the Analytical model (AnM) of
Müller et al [32].

vertically oscillating rod guided through a hole to the upper level (Level 2) where the rod is
attached to an aluminum base on which the bath is mounted.
The mechanical oscillator is made from a commercially available audio vibrator (Fischer
Amps Buttkicker Mini Concert). The audio vibrator was modified as illustrated in figure S1,
i.e. a hole was drilled in its top cover and a 10 mm diameter steel rod was mounted (glued)
directly on the oscillating magnet. Additionally, plastic inserts (3D printed) were positioned
on each side of the oscillating magnet to limit the stroke of the oscillator and plane springs
where inserted below and above the magnet to obtain a resulting harmonic motion of the
magnet and rod. Thus, hard collisions between the moving magnet and the top and bottom
sides of the house of the vibrator are avoided. The oscillator is excited with a standard
laboratory function generator (GF467F, Centrad) combined with a commercially available
(non-expensive) audio amplifier (NS-03G SUB, Nobsound).

8
Eur. J. Phys. 44 (2023) 065002 H B Pedersen et al

Figure 4. Schematic drawing of the constructed experimental system with a vertically


vibrating bath, showing the principal mechanical construction with a mechanical
oscillator fixed on Level 1 of a table that supports and vibrates a bath above Level 2 of
the table. The distance sensor monitors directly the vibration of the bath. Laser 1 and
the photodiode are used to probe the state (stable/unstable) of the liquid surface (see
figure 6 for an example of a measurement). Laser 2 is focused on the top face of one of
the baths walls from a shallow angle (∼10 degrees), and its laser spot thus moves along
the face as the bath vibrates vertically (see figure 12). Camera 1 (CAM1) observes the
bath from above (example in figures 7(a)–(b)) and CAM2 from and an angle of 45
degrees. The dotted pattern in the bottom of the bath is used for the reconstruction of
the surface from the picture taken with CAM1 (see section 4 for a detailed description).

Figure 5. The applied bath and illustrations of possible inserts. (a) The plexiglas bath
that holds the liquid. The four holes in the outer frame are for attachment to the Al-
base. The wells in the bottom of the bath are placed above the corresponding wells that
holds permanent magnets in the Al-base and serves for the fixation of inserts. (b) An
example of an insertable frame for modifying the interactions between the liquid and
the wall. With the shown frame (height 5 mm), brimful conditions can for example be
realized. (c) Two examples of reflective structures that can be placed inside the bath.
The two iron screws that are attached to each structure fit into a pair of wells in the bath
and fix the structure via the coupling to the permanent magnets below.

9
Eur. J. Phys. 44 (2023) 065002 H B Pedersen et al

The oscillating rod is fixed to and supports the Al-base on the upper level (Level 2) of the
table. To ensure a stable vertical motion of the Al-base, the rod is guided into the upper level
via ceramic plain bearings and additionally, one side of the Al-base is similarly guided by a
stationary vertical rod fixed on Level 2.
As seen in figure S1, the Al-base has 16 cylindrical wells for permanent magnets (diameter
10 mm and height 6 mm) below the liquid bath. The magnets inserted in these wells serve to
fix structures that can be inserted into the bath. The bath itself is made of plexiglass with an
outer footprint of 110 × 120 mm2, which fits into a recess in the Al-base. The inner dimension
of the bath is 100 × 100 mm2 and the bath has a total depth of 10 mm. Figure 5 shows more
details of the bath construction and also exemplifies inserts that can be placed inside the bath.
One type of insert is a frame, figure 5(b), that follows the edge of the bath to allow mod-
ification of the liquid-wall boundary condition, e.g. straight walls or brimful conditions,
relevant when observing Faraday waves in the bath. Another type of insert includes dedicated
structures, e.g. chicanes or reflective elements, exemplified in figure 5(c), that can affect the
wave (and droplet) motion at specific regions of the bath.
The upper level (Level 2 in figure 4) is made as a breadboard so all diagnostic instruments
(lasers, photodiode, distance sensors, and cameras (see figure 4) can be positioned and fixed
around the vibrating bath using standard optic components.
For the experiments reported here, intended to demonstrate the functionality of the setup,
we use a Silicone oil (Polydimethylsiloxane, PDMS, CAS no. 63 148-62-9, Merck) with
density ρ = 938 kg m−3, dynamic viscosity η = 1.94 × 10−2 Pa s, kinematic viscosity ν = η/
ρ = 2 × 10−2 m2 s−1 = 20 cSt (at 25 degree), surface tension σ = 2.09 × 10−2 N m−1, and
index of refraction nr = 1.403. The gravitational acceleration in our location is g =
9.816 m s−2 [35].

3.2. Observation of bath vibration and onset of Faraday waves


The vibration of the bath is directly observed via a distance sensor (IR08.D03S, Baumer,
0–3 mm), fixed on the upper level (Level 2) of the table, that hence measures the vertical
position of the bath. The output from the distance sensor is coupled to one channel of a digital
oscilloscope (PicoScope 2207A, Pico Technology). An example of the signal measured with
the distance sensor is shown in figure 6(a) for a driving frequency of fd = 38 Hz.
The occurrence of Faraday waves on the liquid surface is observed from the reflection of a
laser (Laser 1 in figure 4) that illuminates an area of the liquid surface in the central part of
bath with an inclination angle of approximately 45 degrees. The reflected light is detected
with a photodiode (FD100S, Thorlabs) positioned oppositely to the illuminating Laser 1 at a
position where the laser’s transverse size approximately equals the size of the photodiode.
The signal from the photodiode is coupled to a second channel of the digital oscilloscope.
Figure 6(b) shows examples of signals observed with the photodiode for a stable and an
unstable surface. For the stable surface, the photodiode signal displays a near sinusoidal
pattern (blue line in figure 6(b)) corresponding to the vibration of the bath, while a distinct
change occurs when the surface becomes unstable (red line in figure 6(b)). Hence, the onset of
Faraday waves on the liquid surface can for instance be identified from the mean of the
observed signal. This particular way of identifying the acceleration threshold for the onset of
Faraday waves is illustrated in figure 6(c).

3.3. Observation of wave properties


The oscillating bath is observed with two high-speed cameras (Chronos 2.1-HD, Kron
Technologies inc.). The first camera, CAM1 in figure 4, is positioned 345 mm above the bath

10
Eur. J. Phys. 44 (2023) 065002 H B Pedersen et al

Figure 6. Illustration of the measurement of the bath’s oscillation and the onset of
Faraday waves on the liquid surface for a driving frequency of fd = 38 Hz. (a) Signal
from the distance sensor (see figure 4) converted to an actual distance (0.15 mm V−1).
The amplitude of 0.20 mm corresponds to an acceleration of a/g = 1.18. (b) Signal
from the photodiode (see figure 4) for conditions of acceleration below and above the
threshold for the onset of Faraday waves. (c) Determination of the acceleration
threshold from the variation of the mean of the photodiode signal as a function of
acceleration.

and observes the bath at normal incidence. The other camera (CAM2) is located to the side of
the bath and observes the bath at 45 degrees from a distance of approximately 150 mm. The
recording with the cameras reported here are all taken at full resolution (1920 × 1080 pixel)
with a frame rate of 1000 Hz and with an exposure time of 500 μs. The area of the bath is
supplied with light from two commercial fluorescent lamps (not shown in figure 4) allowing
the camera gains to be set low. The air from the ventilation system of the cameras disturbs the

11
Eur. J. Phys. 44 (2023) 065002 H B Pedersen et al

air in front of the cameras and could therefore potentially disturb the observed liquid surface.
Hence, this air flow is guided away from the liquid surface by sets of inserted plastic plates.
Further, another laser (Laser 2 in figure 4) is setup with an approximately 10 degree
inclination angle towards horizontal and with a focus on an up-facing side wall of the bath
where a piece of black colored paper is attached. As the bath vibrates vertically, this laser spot
moves horizontally allowing for the state of the oscillation to be observed from above
with CAM1.

4. Reconstructing surface waves from camera images

To observe and analyze the structure of the Faraday waves on the liquid surface using CAM1,
we use the idea of a free surface synthetic Schlieren (FSSS) technique introduced by Moisy
et al [36] and further developed for example by Wildeman [37] and Li et al [38]. We
explicitly apply a simple FSSS surface reconstruction technique based on direct simulations
of light rays through the used imaging system to reveal the parameters needed for surface
reconstruction. In the present implementation of a FSSS technique, we position a regular
pattern of points (dots) below the liquid, i.e. in practice we place a paper sheet with a dotted
pattern between the Al-base and the bath as indicated in figure 4.
Figures 7(a)–(b) shows examples of pictures taken with CAM1 for a flat surface and for
a surface with Faraday waves excited at fd = 40 Hz and with a vertical acceleration of
a/g = 1.45. The spacing between the dots on the paper below the bath is 1 mm and at the
widest part, there are N = 66 dots in the x-direction and M = 71 dots in the y-direction. For
the analysis shown in this paper, the rectangular band of 66 × 39 dots is used as marked with
the red dashed squares. Around the central dotted area, small perturbations due to the wells in
the plexiglas for mounting inserts (see figure 5(a)) above the permanent magnets can be seen.
When the liquid surface is flat (figure 7(a)), the pattern of dots is seen unperturbed by
CAM1, however, when the liquid surface is curved (figure 7(b)), as induced by waves on the
surface, the dots are displaced in the camera image in a manner that is related to the gradient
of the surface at the point on the surface from which the light originates. As mentioned, the
relation between the apparent displacements of the dots in the camera image and the surface
gradients can be revealed by direct simulations of light rays through the complete imaging
system.
To illustrate in detail the principle behind the presently applied simulations, an example of
simulated light rays used to determine the parameters for the surface reconstruction for the
applied imaging system (liquid, lens and camera) is shown in figure 8. The green lines in
figure 8 thus illustrates light rays from an example point yP in the object plane (i.e.
corresponding to a dot on the paper below the bath) to the image plane (the camera sensor).
For a curved surface, the transmitted light (green lines in figure 8) from the example point yP
below the liquid passes the liquid surface at the mean position yL , and is imaged to the point
yS at the sensor. For a flat surface, the similar light rays (not shown in figure 8) from the same
point yP passes the liquid surface at yL and are imaged at the position yS. With a complete
description (simulation) of the imaging system, the experimental observation of the points yS
(flat reference surface) and yS (curved surface) with the camera sensor, allows the gradient of
the surface ay ( yL ) (i.e. ∂z/∂y) at the point yL as well as the point yL itself to be determined.
Using a full pattern of well-known dimension below the liquid, a map of the gradients of the
surface can be obtained, and the actual surface structure can be reconstructed by integration.
To actually reveal the parameters for the surface reconstruction, simulations as exemplified
in figure 8 are performed. Thus, the light rays (green lines in figure 8 originating from points

12
Eur. J. Phys. 44 (2023) 065002 H B Pedersen et al

Figure 7. Examples of pictures taken with CAM1 (see figure 4). (a) Image of a stable
flat surface showing the pattern of dots on the paper below the liquid bath. (b) Image of
an unstable surface with Faraday waves on the liquid obtained with fd = 40 Hz and
a/g = 1.45, showing the characteristic displacement of the imaged dots that reflect the
variation of the spatial gradient of the liquid surface. To the left in the images, the laser
spot from Laser 2 (see figure 4) is seen on the side wall of the bath. (c) Reconstructed
surface as described with equations (17)–(19).

13
Eur. J. Phys. 44 (2023) 065002 H B Pedersen et al

Figure 8. Illustration of simulations of light rays from the bottom of the liquid bath to
the camera sensor as used to derive the parameters needed for the surface
reconstruction with the present imaging system. The green lines show simulated light
rays propagating from a point (yP) on the paper below the bath through the liquid and
its curved surface, through air, through the lens, and finally to the image sensor of the
camera. The inserts show details of the wave propagation around the liquid surface and
at the imaging sensor, illustrating how the position (yP) of the dot on the paper is
displaced from yS for a flat surface to yS for a surface curve disturbed by the wave. Note
for the lower insert that the asymmetrical distance scales of the vertical and horizontal
axes chosen to emphasize the positions of the points (yP, yS, and yS ) distort the display
of the light rays (green lines). For the simulation, the refractive index of the liquid and
air was set to nr = 1.403 and nair = 1.000, the height of the liquid in the bath was
h = 5.00 mm and the surface wave was simulated with a wavelength λ = 5.73 mm with
an amplitude of 0.40 mm. The lens has a focal length fL = 35 mm and a diameter of
30 mm. The two indicated distances are d1 = 39.55 mm and d2 = 305.69 mm. The size
of the sensor is 10.80 mm in the displayed cut through the setup.

yP on the paper surface below the liquid are propagated numerically as simple rays (straight
line trajectories) (1) through the liquid, (2) across the flat or curved liquid surface where the
refractive index changes from nr = 1.403 to nair = 1 (using Snell’s law), (3) through air from

14
Eur. J. Phys. 44 (2023) 065002 H B Pedersen et al

the liquid surface to the lens, (4) across the lens with focal length fL = 35 mm, and finally (5)
to the image sensor. For the example shown in figure 8, an actual surface wave is modeled as
a harmonic wave so that the height of the liquid surface is model by
h ( y) = h 0 + Dh ´ cos (2py l ), with h0 = 5.00 mm, Δh = 0.40 mm, and λ = 5.73 mm. The
wavelength applied in the simulation is explicitly chosen not to be a multiple of the point
spacing on the paper sheet.
From ray simulations as exemplified in figure 8, the parameters needed for the evaluation
of the surface gradients from the experimental observations of positions yS and yS can now be
determined explicitly for the applied imaging system. For a flat surface, figure 9(a) shows the
direct linear imaging of points yP from the object plane to points yS on the image plane, and
figure 9(b) shows the displacements (yL − yP) of the probed points on the liquid surface
relative to points on the paper. As expected, both these relations are perfectly linear, and the
value of the coefficient aSP (= −0.130) reflects the de-magnification of the object (dotted
paper) to the image plane (camera sensor), while the coefficient aLP (= −0.012) mainly
reflects the light acceptance of the imaging system and the refractive index of the liquid. The
simulated value of aSP can be compared directly to the observed images since the distance
between points on the paper sheet is well known (1 mm), and hence the correspondence
between the simulations and actual physical system can be directly verified through the value
of aSP. For a curved surface, figure 9(c) shows the almost linear relationship between the
slope of the liquid surface and the measurable displacements yS - yS of the imaged points
positions from the curved surface relative to their imaged positions for a flat surface. The
division of the displacement yS - yS with the de-magnification aSP of the imaging system is
done for convenience and simply represents the scaling of the camera sensor image to the
actual size of the object plane. Finally, figure 9(d) shows the relation between the displace-
ments yL - yL of the probed point on the curved surface relative to the flat surface and the
measurable displacement yS - yS . As seen, the relations in figures 9(c)–(d) are linear to a very
good approximation with coefficients determined to be aαS = −0.719 mm−1 and aLS = 0.977
for the present imaging system.
Summarizing, for each point, yP on the paper surface, the corresponding points yS (flat
surface, example in figure 7(a)) and yS (curved surface, example in figure 7(b)) are directly
observable with the camera CAM1 (see figure 4) and the transformations from these obser-
vable positions to actual surface gradients are obtained by a linear transformation revealed by
direct light ray simulations. In practice, to obtain the map of gradients of the curved surface
from the recorded images of the dot-pattern for a flat and a curved liquid surface as displayed
in figure 7, first the probed positions on the flat surface are calculated
yL = yP ´ (1 + aLP ) = yS aSP ´ (1 + aLP ). (17)

Second, the gradients of the surface at the point y˜L are determined as
a y ( yL ) = aaS ´ ( yS - yS ) aSP , (18)

and the corresponding coordinates on the liquid surface are determined as

yL = yL + aLS ´ ( yS - yS ) aSP (19)

With the complete map of gradients ( xL , yL , ax (xL ), ay ( yL )), the actual surface can be
reconstructed by integration. In this particular case, we used the grad2Surf software [39]
developed for MATLAB [33]. Figure 7(c) shows the final reconstructed surface based on the
images in figures 7(a)–(b) using the described technique.

15
Eur. J. Phys. 44 (2023) 065002 H B Pedersen et al

Figure 9. Properties of the imaging of dots at positions yP the on paper below the liquid
to the the camera sensor through a liquid surface that is either flat or curved due to
surface waves (see figure 8 for an illustration of the simulated light rays). (a)
Correspondence of dot-positions on the sensor (yS) and on the paper (yP) for a flat
surface. (b) Correspondence of probed position on the liquid surface (yL) and the dot-
position (yP) for a flat surface. (c) Correlation of the surface slope (ay ( y˜L )) and the
measured displacements of dot-positions on the sensor image. For convenience, the
displacements have been scaled with the de-magnification factor aSP of the imaging
system. (d) Correlation of the displacements of the probed positions of the liquid
surface (yL - yL ) and the measured displacements of the dot-positions on the sensor
image.

5. Results

5.1. Acceleration threshold for the onset of Faraday waves


Figure 10 displays the measured acceleration thresholds for the onset of Faraday waves as
determined in the present experiment (red dots) as a function of the driving frequency for a
situation where brimful boundary conditions are used, i.e. with the insert displayed in
figure 5(b) inserted.
Using tabulated values for the kinematic viscosity (ν) and surface tension (σ) for the used
Silicone oil, the experimental data are compared to the full hydrodynamical model (dashed
blue curve) [12, 31], the phenomological model (full green curve) [12], the analytical model

16
Eur. J. Phys. 44 (2023) 065002 H B Pedersen et al

Figure 10. Experimentally determined acceleration thresholds for the onset of Faraday
waves in the bath filled with a Silicone oil (ρ = 938 kg m−3, ν = 2.0 × 10−5 m2 s−1,
and σ = 20.9 × 10−3 N m−1) to a height of h = 5 mm and using an insert into the bath
to obtain brimful conditions at the bath boundaries (see figure 5). The data are
compared to the full hydrodynamical model (FHM) of Kumar et al [12, 31], the
phenomenological model (PhM) of Kumar and Tuckerman [12], the analytical model
(AnM) of Müller et al [32] and the approximate model for capillary waves
(equation (16)) [9, 22, 34]. The full blue line shows an adaption of the FHM to the
data with ν = 2.2 × 10−5 m2 s−1 and σ = 17 × 10−3 N m−1.

(gray curve) [32], as well as the model for capillary waves (dashed black line, equation (16))
in a bath of infinite depth. As expected, the model for capillary waves fails to make a
quantitative prediction of the acceleration threshold. With the tabulated values for the used
oil, the other models also fail to account for the observed acceleration thresholds. The
phenomenological model in particular fails to account for the frequency dependency at low
frequencies. The full hydrodynamical model and the analytical model essentially give iden-
tical predictions and qualitatively predicts the correct shape of the frequency dependence,
however, generally underestimating the acceleration threshold at higher frequencies.
Allowing the viscosity and the surface tension to vary as free parameters, a good fit to the
experimental data can be obtained with both the full hydrodynamical model and the analytical
model.

5.2. Properties of Faraday waves


Figure 11 shows examples of measured and reconstructed surfaces for three different forcing
frequencies at accelerations slightly above the Faraday thresholds at these frequencies. It is
clearly seen how the standing wave pattern depends strongly on the excitation frequency
reflecting the dispersion relation for the waves and the conditions for standing waves
(equation (7)–(8)). The contribution from two components of standing waves can be seen in
the formation of square structures, most visible for 30 Hz and 40 Hz.

17
Eur. J. Phys. 44 (2023) 065002 H B Pedersen et al

Figure 11. Examples of reconstructed liquid surfaces and their associated Fourier
transforms obtained with the experimental system (see figure 4). The experimental
settings are (a)–(b): forcing frequency fd = 15 Hz and acceleration a/g = 0.40, (c)–(d):
fd = 30 Hz and a/g = 0.88, and (e)–(f): fd = 40 Hz and a/g = 1.14. The red circles in
panels (a), (c), and (e) mark the positions where the surface amplitudes are traced as
displayed in figure 12.

In figure 12, the time evolution of selected points on the waves corresponding to
figure 11(a), (c) and (e) are illustrated. It is evident that the waves observed are all sub-
harmonic, fF = 1/2fd, to a very high degree, i.e. a deviation from this exact correspondence
cannot be seen. The bath position as monitored with Laser 2 (figure 4) is seen to be well
described by a harmonic function, while small disturbances are seen for the Faraday waves,
i.e. most clearly seen for the lower frequency (highest amplitude). The phase shift is close to
zero for 15 Hz and 30 Hz, but has a finite value for the data displayed for 40 Hz, showing that
the applied acceleration is higher than the threshold value. The actual accuracy of the
selectable acceleration is presently determined by the voltage resolution of our function
generator and in the present case this limited a more accurate match of the Faraday threshold
for 40 Hz.
To analyze the wavelength composition of the Faraday waves in more detail, a Fourier
transform analysis is applied, i.e. Fourier amplitudes are explicitly calculated as
1
AFT (k x , k y) =
M´N
åå A (x, y) exp (-ikx x) exp (-ik y y) (20)
n m

where AFT(kx, ky) is the (complex) amplitude of the surface as a function of position and kx =
2π/λx and ky = 2π/λy are wavenumbers in the two directions corresponding to wavelengths
λx and λy. Panels (b), (d), and (f) of figure 11 display the actual real amplitude given by
1
A (k x , k y ) = AFT ´ conj (AFT ) . (21)
M´N
for the corresponding surfaces in panels (a), (c), and (e). The direct square symmetry is still
evident (especially for 15 Hz and 30 Hz), but also shows preference for linear combinations as
most clearly seen for 40 Hz.
The time domain variations (figure 12), showing fF = 1/2fd, and the spatial variations
(wavelength) can be combined to obtain a direct measurement of the dispersion relation for

18
Eur. J. Phys. 44 (2023) 065002 H B Pedersen et al

Figure 12. Surface oscillations as a function of time compared to the oscillation of the
bath. The black dots in panels (a), (b), and (c) show the vertical amplitudes of the points
marked with a red circle in figure 11(a), (c), and (e). The blue points in the three panels
show the oscillation of the bath as recorded from the motion of the laser spot from
Laser 2 (see figure 4).

the Faraday waves. The result of this analysis is shown in figure 13. The upper panels (a), (b),
and (c) show the Fourier amplitude as a function of the combined wavelength

1 l= 1 l2x + 1 l 2y (22)

i.e. integrating over the azimuthal angle for the spectra in figures 11(b), (d), and (f).
Figure 13(d) shows the finally obtained dispersion relation in comparison to the dispersion
relation for free waves on an inviscid liquid (equation (15)) as well as in comparison to the
dispersion relation obtained with the full hydrodynamical model of Kumar et al [12, 31]. For
illustration, the gravity and capillary components of the dispersion relation for free surface
waves are plotted individually showing for example that the total dispersion relation has about
equal contributions from gravity and capillary waves at 12–13 Hz.

6. Discussion

6.1. Experimental improvements

Our setup is constructed with an awareness of keeping the cost affordable for teaching
laboratories at university or high school, and clearly improved results could be obtained with
other choices of instrumentation and design for certain parts of the setup.
With a function generator with better resolution in the setting of the amplitude level (at
present the resolution is 0.01 V below 1 V and 0.1 V above), a finer variation of the bath’s
acceleration would be possible and hence, the Faraday threshold as a function of the driving
frequency could be mapped out with better accuracy. Also, the development of the waves just
above the Faraday threshold could be investigated in higher detail. This concerns for example
the phase shift of the oscillating waves relative to the bath’s oscillation. In a future upgrade of
the experiment, we would indeed foresee the addition of a better function generator. Alter-
natively, a high-quality sound card from a personal computer could also be applied instead of
a function generator.

19
Eur. J. Phys. 44 (2023) 065002 H B Pedersen et al

Figure 13. Dispersion relation for Faraday waves on the oscillating bath. (a)–(c) show
Fourier transforms corresponding to figure 11 as a function of the 1/λ = 1 l2x + 1 l2y .
The black lines show Gaussian fits to the main peaks in the spectra. (d) Experimentally
determined dispersion relation (black dots) for the Faraday waves on the used Silicone
oil compared to model calculations using the full hydrodynamical model of Kumar
et al [12, 31] and to the dispersion relation (equation (15)) for free surface waves on an
inviscid fluid (dashed red line). For illustration, the individual contributions from free
gravity (dashed green line) and capillary (dashed blue line) in equation (15) are
displayed.

With a larger bath, we could obtain a higher resolution in the Fourier transforms
(figures 13(a)–(c)) allowing for better accuracy in the measurement of the dispersion relation,
however, this would also require cameras of higher resolution, and would in general be
associated with a relatively large cost.
The applied oscillator is made with a commercially available audio vibrator (Buttkicker),
modified with two plane springs to make the resulting oscillation harmonic, which reduces
significantly the price tag for the oscillating system. With a stronger vibrator and possibly
with an improved design of the vibrating system [40], we could examine other regions of

20
Eur. J. Phys. 44 (2023) 065002 H B Pedersen et al

Faraday waves, for instance at higher frequencies, however, for the investigation of Faraday
waves in a teaching perspective, the accessible (15–60 Hz) region of excitation is sufficient
where the transition between gravity waves and capillary waves is highlighted.
For the observation of the dispersion relation, good results could also have been obtained
with cameras from modern smartphones. As our system will be used as an integrated sta-
tionary exercise in a laboratory course as well as for research, we decided to use (still
affordable) permanently installed high-speed cameras.

6.2. Surface reconstruction

In the present paper, we implemented a FSSS surface reconstruction method based on direct
simulations of light rays from points in a regular dotted pattern below the liquid to the
corresponding points registered by the imaging sensor. The simulations provide explicit
information on all parameters of the specific imaging system (aSP, aLP, and aαS) needed to
reconstruct the gradient of the liquid surface (see equations (17)–(19) and figure 9) from
measurements of the displacement of the dots in the pattern as compared to the dot-positions
measured with a flat surface. An advantage of this approach using direct simulation of light
rays, is that it allows a direct clarification of the properties (e.g. possible distortions or non-
linearity) of the applied imaging system. Moreover, the direct propagation of light rays is
relatively easily implemented numerically and promote a deep understanding of the sig-
nificance of the applied imaging system for students. The analytical implementation of the
FSSS technique by Moisy et al [36] relied on the paraxial approximation, the weak slope
approximation, and the weak amplitude approximation. Compared to the present case, these
approximations are essentially equivalent to making the approximation yL = yS aSP , i.e. the
relation in figure 9(d) would be flat, which in our case would lead to significant derivations in
the reconstruction of the image.
In other implementations of the FSSS method, randomly distributed points or regular
patterns have been applied (see e.g. Wildeman [37] and Li et al [38] for overviews of
techniques applied). For the purpose of surface reconstruction applied in this paper, a regular
pattern (in the present case dots) makes the point identification and image reconstruction
simple and easy to implement numerically. More advanced FSSS techniques [37, 38] could
evidently also be implemented for the present experimental system, however, we consider the
relative simplicity of the concept of direct ray propagation and the simple linear transfor-
mations (equations (17)–(19)) as appealing in a teaching situation, and the obtained results
(figures 11 and 13) are quite satisfactory.

6.3. Examples of future directions for student-driven research

The setup will be integrated as a project exercise into a course on experimental physics at the
bachelor level at Aarhus University (see also [41] for some description of the course).
We believe that the aspects of the setup demonstrated in this paper, could for example
enable students to develop an intuitive understanding of wave phenomena including in
particular standing waves and dispersion relations. In this respect, it should be noted that the
setup operates in a regime where both gravity and capillary waves are of equal importance
(figure 13).
The models established by Kumar et al [12, 31] and Müller et al [32] are well suited for
training numerical calculations at the university level. Additionally, several aspects, beyond
the demonstrations of the Faraday instability and the mapping of the dispersion relation
shown here, can be imagined with the present setup either for teaching purposes or as
dedicated student-driven research projects. One possibility would be to explore further the

21
Eur. J. Phys. 44 (2023) 065002 H B Pedersen et al

idea of Kumar and Tuckermann [12] that the measurement of the acceleration threshold for
Faraday waves as a function of the driving frequency can be used to actively determine the
fluid viscosity (ν or η) and the surface tension σ simultaneously by a fit with the (exact) full
hydrodynamical model to measured data. The potential of this method, which to the best of
our knowledge has not been examined experimentally before, was indicated for the data set
shown in figure 10, where a very good match of the model to the experimental data could be
obtained by adjustment of the tabulated viscosity and surface tension for the used Silicone oil.
It would thus be a very interesting line of advanced student research to validate this method in
more detail. Moreover, upon validation, the method could constitute a universal way to
examine fluid properties for example as a function of temperature. In the student laboratory in
the first year of study, fluids like water, household oil, and syrup could be investigated.
Another interesting direction for future student-driven research could be the observation
and investigation of harmonic waves ( fF = fd) and higher subharmonic waves ( fF = 3/2fd)
and in particular the transitions between subharmonic and harmonic waves that could occur
for low liquid depths and at low frequencies as demonstrated with the calculations displayed
in figure 2(d) and figure 3(d).
The explicit importance of liquid-wall interactions [3] can also be examined with the
present setup. In the demonstration reported here, we used the brimful conditions, but clearly
the effect of straight walls or other types of boundaries can be realized with appropriate inserts
into the bath. For example, the onset and significance of meniscus waves (emitted with
frequency fd) in competition with subharmonic waves at 1/2fd could be investigated. Also, the
dependence of the standing waves on the bath geometry, i.e. square, circular, oval, etc can
relatively easily be studied with simple inserts.
Finally, the setup is also meant to examine the dynamics of bouncing and walking dro-
plets. In this respect, the description in this paper serves to demonstrate the general func-
tioning of the setup, while actual results on droplet dynamics will be given in a separate
manuscript.

7. Conclusion

We have described a setup for investigations of Faraday waves for laboratory teaching at
university. In its present form, the setup allows for example for the examination of the
acceleration threshold for Faraday waves as a function of driving frequency as well as for
investigations of the nature for the Faraday waves. At a given driving frequency, the onset of
the Faraday instability of the liquid surface can be directly deduced from the reflection of a
laser from the liquid surface in a simple and objective way. Images acquired at high speed
(∼1 kHz) of the bath taken from above, allows reconstruction of the surface using a free
surface synthetic Schlieren technique. These surfaces allows the deduction of both the wave
frequency and wavelength and can thus be used to determine the dispersion relation for
Faraday waves of the investigated liquid. Similar setups can be used for teaching fluid
dynamics at different levels at university, and, moreover, with the present setup, we foresee
several interesting projects within fluid dynamics to be realized. The setup will also be used
for student-driven investigations of bouncing and walking droplets.

Acknowledgments

We thank the mechanical and electrical workshops at the Department of Physics and
Astronomy, Aarhus University for excellent support during the development of the setup.

22
Eur. J. Phys. 44 (2023) 065002 H B Pedersen et al

Data availability statement

The data cannot be made publicly available upon publication because they are not available in
a format that is sufficiently accessible or reusable by other researchers. The data that support
the findings of this study are available upon reasonable request from the authors.

ORCID iDs

Henrik B Pedersen https://orcid.org/0000-0002-7617-8919

References

[1] Kityk A V, Knorr K, Müller H-W and Wagner C 2004 Spatio-temporal Fourier analysis of
Faraday surface wave patterns on a two-liquid interface Europhys. Lett. 65 857–63
[2] Kityk A V, Embs J, Mekhonoshin V V and Wagner C 2005 Spatiotemporal characterization of
interfacial faraday waves by means of a light absorption technique Phys. Rev. E 72 036209
[3] Rajchenbach J, Clamond D and Leroux A 2013 Observation of star-shaped surface gravity waves
Phys. Rev. Lett. 110 094502
[4] Rajchenbach J and Clamond D 2015 Faraday waves: their dispersion relation, nature of bifurcation
and wavenumber selection revisited J. Fluid Mech. 777 R2
[5] Salgado Sánchez P, Yasnou V, Gaponenko Y, Mialdun A, Porter J and Shevtsova V 2019
Interfacial phenomena in immiscible liquids subjected to vibrations in microgravity J. Fluid
Mech. 865 850–83
[6] Douady S 1990 Experimental study of the Faraday instability J. Fluid Mech. 221 383–409
[7] Milner S T 1991 Square patterns and secondary instabilities in driven capillary waves J. Fluid
Mech. 225 81–100
[8] Christiansen B, Alstrøm P and Levinsen M T 1995 Dissipation and ordering in capillary waves at
high aspect ratios J. Fluid Mech. 291 323–41
[9] Bechhoefer J, Ego V, Manneville S and Johnson B 1995 An experimental study of the onset of
parametrically pumped surface waves in viscous fluids J. Fluid Mech. 288 325–50
[10] Batson W, Zoueshtiagh F and Narayanan R 2013 The Faraday threshold in small cylinders and the
sidewall non-ideality J. Fluid Mech. 729 496–523
[11] Jiang L, Ting C-L, Perlin M and Schultz W W 1996 Moderate and steep faraday waves:
instabilities, modulation and temporal asymmetries J. Fluid Mech. 329 275–307
[12] Kumar K and Tuckerman L S 1994 Parametric instability of the interface between two fluids
J. Fluid Mech. 279 49–68
[13] Faraday M 1831 XVII on a peculiar class of acoustical figures; and on certain forms assumed by
groups of particles upon vibrating elastic surfaces Philos. Trans. R. Soc. Lond. 121 299–340
[14] Benjamin T B and Ursell F 1954 The stability of the plane free surface of a liquid in vertical
periodic motion Proc. R. Soc. LondonA 225 505–15
[15] Miles J and Henderson D 1990 Parametrically forced surface waves Ann. Rev. Fluid Mech. 22
143–65
[16] Westra M-T, Binks D J and Van De Water W 2003 Patterns of Faraday waves J. Fluid Mech. 496
1–32
[17] Perinet N, Juric D and Tuckerman L S 2009 Numerical simulation of Faraday waves J. Fluid
Mech. 635 1–26
[18] Takagi K and Matsumoto T 2015 Numerical simulation of faraday waves oscillated by two-
frequency forcing Phys. Fluids 27 032108
[19] Kahouadji L, Périnet N, Tuckerman L S, Shin S, Chergui J and Juric D 2015 Numerical simulation
of supersquare patterns in Faraday waves J. Fluid Mech. 772 R2
[20] Couder Y, Fort E, Gautier C-H and Boudaoud A 2005 From bouncing to floating: noncoalescence
of drops on a fluid bath Phys. Rev. Lett. 94 177801
[21] Couder Y, Protière S, Fort E and Boudaoud A 2005 Walking and orbiting droplets Nature 437
208–208

23
Eur. J. Phys. 44 (2023) 065002 H B Pedersen et al

[22] Protière S, Boudaoud A and Couder Y 2006 Particle-wave association on a fluid interface J. Fluid
Mech. 554 85–108
[23] Eddi A, Boudaoud A and Couder Y 2011 Oscillating instability in bouncing droplet crystals
Europhys. Lett. 94 20004
[24] Couder Y, Boudaoud A, Protière S and Fort E 2010 Walking droplets, a form of wave-particle
duality at macroscopic scale? Europhys. News 41 14–8
[25] Harris D M, Moukhtar J, Fort E, Couder Y and Bush J W M 2013 Wavelike statistics from pilot-
wave dynamics in a circular corral Phys. Rev.E 88 011001
[26] Bush J W M 2015 Pilot-wave hydrodynamics Ann. Rev. Fluid Mech. 47 269–92
[27] Andersen A, Madsen J, Reichelt C, Ahl S R, Lautrup B, Ellegaard C, Levinsen M T and Bohr T
2015 Double-slit experiment with single wave-driven particles and its relation to quantum
mechanics Phys. Rev. E 92 013006
[28] Bohr T 2018 Quantum physics dropwise Nature Phys. 14 209–10
[29] Turton S E, Couchman M M P and Bush J W M 2018 A review of the theoretical modeling of
walking droplets: Toward a generalized pilot-wave framework Chaos 28 096111
[30] Bush J W M and Oza A U 2020 Hydrodynamic quantum analogs Rep. Prog. Phys. 84 017001
[31] Kumar K 1996 Linear theory of Faraday instability in viscous liquids Proc. Roy. Soc.A 452
1113–26
[32] Müller H W, Wittmer H, Wagner C, Albers J and Knorr K 1997 Analytic stability theory for
Faraday waves and the observation of the harmonic surface response Phys. Rev. Lett. 78
2357–60
[33] Matlab version: 9.3.0.713579 (r2017b), 2017.
[34] Edwards W S and Fauve S 1994 Patterns and quasi-patterns in the Faraday experiment J. Fluid
Mech. 278 123–48
[35] Nielsen J E, Strykowski G and Forsberg R 2017 Absolute gravity measurement in Denmark and
Greenland Tech. Rep. -Ver. 1.2 (https://doi.org/https://ftp.space.dtu.dk/pub/JEMNI/DTU_
AG_2017ver12_w_data.pdf)
[36] Moisy F, Rabaud M and Salsac K 2009 A synthetic Schlieren method for the measurement of the
topography of a liquid interface Exp. Fluids 46 1021
[37] Wildeman S 2018 Real-time quantitative schlieren imaging by fast fourier demodulation of a
checkered backdrop Exp. Fluids 59 97
[38] Li H, Avila M and Xu D 2021 A single-camera synthetic schlieren method for the measurement of
free liquid surfaces Exp. Fluids 62 227
[39] Harker M and O’Leary P (2022) Surface reconstruction from gradient fields: grad2surf version 1.0.
MATLAB Central File Exchange. (https://doi.org/https://www.mathworks.com/matlabcentral/
fileexchange/43149-surface-reconstruction-from-gradient-fields-grad2surf-version-1-)
[40] Harris D M and Bush J W M 2015 Generating uniaxial vibration with an electrodynamic shaker
and external air bearing J. Sound Vib. 334 269
[41] Pedersen H B, Andersen J E V, Nielsen T G, Iversen J J, Lyckegaard F and Mikkelsen F K 2019
An experimental system for studying the plane pendulum in physics laboratory teaching Eur. J.
Phys. 41 015701

24

You might also like