Você está na página 1de 118

Diploma Thesis

Experimentelle Untersuchung der Basset


Gedchtniskraft auf eine starre Kugel in
instationrer Bewegung
Experimental measurement of the history
forces on a rigid sphere in unsteady motion
prepared by

Sebastian Lambertz
from Duderstadt
at the Institute for Nonlinear Dynamics

Thesis period:

18th July 2011 until 18th July 2012

First referee:

Prof. Dr. E. Bodenschatz

Second referee:

Prof. Dr. A. Tilgner

Abstract
The aim of the present study was to examine the impact of the different forces acting
on a rigid sphere moving in a Newtonian fluid. The equation of motion proposed
and improved by Basset [Bas88], Boussinesq [Bou85], Oseen [Ose10], Tchen [Tch47],
Maxey and Riley [MR83] and Gatignol [Gat83] was solved for the quasi-steady and
transient case.
A linear motion system with a travel distance of 4.05 m and a maximal force of
7 600 Newton was build up. With a net weight of 115 kg, we were able to run the
moving platform with a velocity up to 7.5 m/sec and accelerations up to 28 m/sec2 .
In the constant driving mode the velocity deviation was 0.1% at a velocity of 5 m/sec.
With this system we investigated the response of a sphere to sinusoidal forcing.
Therefore we filmed the motion of the millimetre sized spheres in a container with
silicon oil. The data fitted the solution quite well and correspond with the measurements found in the literature. We checked the dependence on the Reynolds number
and found that the solution is still correct when both Reynolds number and Strouhal
number are small but of the same order. For the first time, the break-down of the
solution due to non-linear effects was measured.
The results are especially important for the moving of tracer and inertia particles
in unsteady flows.
Keywords:
Basset history force, memory term, fractional calculus, relaxation, low Re flow, nonlinear effects

iii

Contents
1 Introduction
1.1
1.2

1.3

1.4

1.5

The governing equations . . . . . . . . . . . . . . . . . . . . . . . . .

1.1.1

The Navier Stokes equations . . . . . . . . . . . . . . . . . . .

Flow around a sphere . . . . . . . . . . . . . . . . . . . . . . . . . . .

1.2.1

Low Reynolds number flow . . . . . . . . . . . . . . . . . . . .

1.2.2

High Reynolds number flow . . . . . . . . . . . . . . . . . . .

Force measurement . . . . . . . . . . . . . . . . . . . . . . . . . . . .

1.3.1

Low Re solution for the Navier-Stokes equations . . . . . . . .

1.3.2

Forces acting on a particle in fluid . . . . . . . . . . . . . . . .

1.3.3

High Reynolds number flow . . . . . . . . . . . . . . . . . . . 14

1.3.4

Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18

Velocity measurement

2.2

. . . . . . . . . . . . . . . . . . . . . . . . . . 18

1.4.1

Special solutions for the Navier-Stokes equations . . . . . . . . 19

1.4.2

Experimental studies . . . . . . . . . . . . . . . . . . . . . . . 22

1.4.3

High Reynolds number flow . . . . . . . . . . . . . . . . . . . 27

1.4.4

Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32

Motivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32

2 Theory
2.1

35

The equation of motion . . . . . . . . . . . . . . . . . . . . . . . . . . 35


2.1.1

Inertial frame of reference . . . . . . . . . . . . . . . . . . . . 35

2.1.2

Non-inertial frame of reference . . . . . . . . . . . . . . . . . . 36

Quasi - steady state solution for oscillating sphere . . . . . . . . . . . 36


2.2.1

Assumptions and Conditions . . . . . . . . . . . . . . . . . . . 37

2.2.2

Solution with complex wave ansatz . . . . . . . . . . . . . . . 38

2.3

Characteristics of the resulting equation . . . . . . . . . . . . . . . . 38

2.4

Introduction to fractional calculus . . . . . . . . . . . . . . . . . . . . 41


2.4.1

Different definitions of the fractional derivative . . . . . . . . . 41

Contents

2.5

2.4.2 Laplace Transform of fractional derivatives . . . . . . . . .


General solution for the Basset equation . . . . . . . . . . . . . .
2.5.1 Laplace Transformation of the Basset equation . . . . . . .
2.5.2 Excursion: Dielectric relaxation and the response function
2.5.3 Solution for oscillatory motion . . . . . . . . . . . . . . . .
2.5.4 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . .

.
.
.
.
.
.

.
.
.
.
.
.

42
44
45
46
47
51

3 setup - censored

53

4 Results
4.1 Quasi-steady state . . . . . . . . . . . . . . . . . .
4.1.1 Full solution versus Quasi-steady state . . .
4.1.2 Assumptions . . . . . . . . . . . . . . . . . .
4.1.3 Measured trajectories . . . . . . . . . . . . .
4.1.4 Results . . . . . . . . . . . . . . . . . . . . .
4.1.5 Comparison with the data in the literature .
4.2 High Reynolds number regime . . . . . . . . . . . .
4.2.1 Constant frequency . . . . . . . . . . . . . .
4.2.2 Constant amplitude . . . . . . . . . . . . . .
4.2.3 The role of the forcing . . . . . . . . . . . .
4.2.4 High Reynolds number and density variation
4.2.5 Summary and application of the results . . .
4.3 Discussion . . . . . . . . . . . . . . . . . . . . . . .

55
55
55
56
58
60
63
65
67
69
70
75
76
77

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

5 Outlook

81

6 Appendix
6.1 Solution in the inertial frame . . . . . . . . . . . . . .
6.1.1 Analysis of the result in the inertial frame . .
6.2 Solution for the co-moving frame . . . . . . . . . . .
6.2.1 Analysis of the result in the co-moving frame .
6.3 Inverse Laplace Transform of the Basset equation . .

83
83
85
88
90
93

vi

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

Nomenclature
Latin Letters
Variable

Meaning

Unit

a
As = A
Ap
(Ap /As )0
AcJ

Radius of the particle


Amplitude of sled motion
Amplitude of particle motion
Amplitude ratio, predicted value from linear theory
Acceleration number according to Jourdan et al.
dup
2a
[JHI+ 07] AcJ = (vu
2 dt
p)

m
m
m

AcOH

Acceleration number according to Odar and Hamildu


ton [OH64] AcOH = 2a
u dt
Acceleration number according to Temkin and Kim



r
[TK80] AcT = fp 1 U 2 2a dU
dt

AcT

r,T emkin

cD
cH
cA
cT emkin
d
D
E
f
F
Fdrag
FHistory
Fadded
Fdisturbance,T chen

Correction factor for drag force


Correction factor for Basset History force
Correction factor for added mass term
Correction factor according to Temkin and Kim (see
[TK80] ) cT emkin = cD + f (Ac)
Diameter of sphere
Diameter of container
Electrical field
Frequency
Normalized force
Stokes drag force
Basset History force
Added mass term
Flow disturbance term according to Tchen [Tch47]

m
m
kg m s3 A1
sec1
m sec2
kg m sec2
kg m sec2
kg m sec2
kg m sec2

vii

Nomenclature
Variable

Meaning

Unit

Fdisturbance,CL

kg m sec2

Fdrag
Fn

Flow disturbance term according to Corrsin and


Lumley [CL56]
Flow disturbance term according to Buevich
[Bue66]
Flow disturbance term according to Maxey and Riley [MR83]
Stokes force
a2 As
As
Force number
= 9S A
Ap
p

g
g(t)

Gravity acceleration
Time dependent part of force

m sec2

Fdisturbance,Buevich
Fdisturbance,M R

r

Ga
KBasset
KM ei
Ka
Kr
L
mp
mf
mnonlinear
n
p
P
Ref
ReJourdan

kg m sec2
kg m sec2

1 F a3

m
m sec2.5

Galileo number Ga =

Kernel of the History force integral according to


Basset
Kernel of the History force integral according to Mei
Rubber constant in the axial direction
Rubber constant in the radial direction
Length scale
Mass of the fluid
Mass of the particle
Mass of fluid and particle in the non-linear regime
Integer number
Pressure
Polarization
Reynolds number based on the force on the particle
p
Ref = aA

Reynolds number according to Jourdan et al.


[JHI+ 07]

m sec2.5
N m1
N m1
m
kg m3
kg m3
kg m3

kg m1 sec2
A m2 sec

ReJourdan =

r

2a vup


dup
dt

Res

p
f

kg m sec2


dup 2
dt

dup,y
dt

2

Reynolds number based on the sled force Res =


aAs

Rev
Ro

viii

Reynolds number based on the velocity Ref =


Roshko number

aAp

Nomenclature
Variable

Meaning

Unit

s
S
Sl
T
u

Integration variable for Laplace transform. s = i


Dimensionless frequency
Strouhal number
Time scale
Relative velocity of particle and fluid in the frame
moving with the fluid u = up v
Velocity of the particle in the inertial frame


2

Stokes terminal velocity uts = 29 a fp 1 g

sec
m sec1

Decay of the initial conditions


Response function
Mean velocity
Relative velocity of fluid and particle Ur,T emkin =
v up
Velocity of the fluid in the inertial frame
Velocity of the fluid in the frame moving with the
fluid
Volume of the particle
Integration variable

m sec1
m sec1

up
uts
u0
u
umean
Ur,T emkin
v
v
V
x0

m sec1
m sec1

m sec1
m sec1
m3

Greek Letters
Variable

Meaning

Unit

Density ratio =

Order of fractional derivative

f
p

q q

Inverse Stokes number =

2
f

Ratio of container diameter D to sphere diameter d


= Dd
Ratio for importance of wall effects 2 = Dd Rev

external force time


Force time ratio f = unsteady force time


0

Expansion parameter
Electric permittivity of free space
(dynamic) Viscosity
Integration variable

2
9

1
S

A2 sec4 kg1 m3
kg m1 sec1

ix

Nomenclature
Variable

Meaning

Unit

Fitting parameter for van der Pol model


Dimensionless parameter from Stokes solution
2
= i a
Fitting parameter for dielectric model by Hilfer
[Hil02]
Fitting parameter for van der Pol model
(kinematic) Viscosity
Factor for non dimensionalization of velocity

= pp+ 1 f F

kg sec2

f =
p

Mathematical constant = 3.141...


Density of the fluid
Density of the particle

kg m3
kg m3

Non dimensional parameter =

Factor for non dimensionalization of time

1 , 2

kg m2 sec2
m2 sec1
m sec1

2 f

p
9
2 (p + 12 f )

sec1

1
a2 p + 2 f

2
9

Relaxation time
2
r = 92 a fp

sec1

e
0
00

Fitting parameter for van der Pol model


Dielectric susceptibility
Real part of the dielectric susceptibility
Imaginary part of the dielectric susceptibility
Fitting parameter for van der Pol model
(angular) Frequency = 2f

kg sec1

kg m sec2
rad sec 1

Nomenclature

miscellaneous
Meaning

Symbol

Partial derivative
x f (t) =
Total derivative following the particle

f (t)
f (t)
df (t)
=
+ up
dt
t
x

Total derivative following the


fluid

f (t)
f (t)
Df (t)
=
+v
Dt
t
x

Partial derivative with respect to


the time
Riemann - Liouville fractional integration (left hand)

Riemann - Liouville fractional integration (right hand)

Riemann - Liouville fractional


derivative

f (t)
x

f (t)
f = t f (t) =
t

a It =

t Ib

1 Zt
(t )(1) f ()d
() a

1 Zb
=
( t)(1) f ()d
() t

m
a Dt f (t) =

dn
n
f (t)
aI
dxn t

Laplace Transformation
F (s) = L {f } (s) =

Z
0

est f (t) dt,

sC

xi

1 Introduction
In this thesis work the motion of particles in unsteady flows is examined theoretically
and experimentally. It is organized as follows. We start with a general introduction
and identify proper parameters to characterize the problem. The previous works
are summarized and discussed. Based on the parameter range and the results of the
previous experiments we chose the configuration of our experiment. The following
theory part contains solutions for the governing equations. Then we go to the setup
and explain the constituent parts of our experiment. The fourth section contains
the results of the experiment and we compare our results with the previous results.
The last part contains a outlook based on the new results.
We start this section with a short introduction and explain why the problem is
important for a wide range of applications. Then we go on to the experimental
findings. We compare the flow around fixed and freely moving spheres and discuss
the implications for our experiment. The various experimental and theoretically
work on this problem is divided into two parts depending on the ansatz. We start
with the force measurements at fixed velocity and go on to the velocity measurements
at controlled forcing. The last part of this section summarizes the previous work
and gives an overview, which part of the parameter space we explore experimentally
and why we do this.

1.1 The governing equations


The motion of particles in gases or fluids is a very interesting problem with numerous
applications in nature and technology. Examples are water droplets forming clouds,
pollutant dispersion through air or the motion of gas bubbles inside the ocean. Also
combustion, the mixing of food ingredients and many others can only be understood,
if the motion of the particle (or bubble) in the fluids can be calculated. The examples
gave already an impression of the wide parameter range. The different sizes of
particles clearly affects the motion. The motion of millimetre-sized rain drops is

1 Introduction
different from the motion of small pollen. Another important quantity is the density
ratio of the particle and the fluid. The motion of a bacteria with almost the same
density as the surrounding water will be different from the motion of an air bubble
with the same size. Also the properties of the flow are important. A strong wind
distributes pollution particles faster than a light breeze. The problem gets even more
complicated, if we allow the flow to be unsteady. A flow with a time dependent
velocity is the most complicated, but also most interesting problem we want to
discuss in this thesis.

1.1.1 The Navier Stokes equations

The motion of pure fluids and suspensions has been studied for very long time. For
the first time (in 1823) Navier derived the equations, which were named after him
and Stokes. Navier [Nav23] - like Cauchy in 1823 [Cau23], Poisson in 1829 [Poi28]
and Saint-Venant in 1843 [dSV43] - began with the elasticity theory and applied his
equations to fluids, whereas Stokes (in 1845) [Sto46] derived the equations directly
and pointed out their connection to the elasticity theory. Let the fluid velocity be
denoted by v, the dynamic viscosity by and the fluid density by . The external
forces acting on the fluid are denoted by f . Then the Navier-Stokes equations for
an incompressible fluid are given by:

~v
+ ~v ~v = p + 2~v + f~
t

(1.1)

The discussion of the different terms is easier, if we make the equation non-dimensional.
Let T define a typical time scale of the change of the boundary conditions. The velocities are rescaled with the mean velocity U , and the length is rescaled with a
typical length scale L to get the Navier-Stokes Equation in non-dimensional form.
The prime indicates a dimensionless parameter:
v0 =

v
,
U

p0 =

p
,
U2

0 = L,

t0 =

t
.
T

1.1 The governing equations


Therewith we get the non-dimensional Navier-Stokes equations:
L/U v 0

02 v 0 + forces
+ v 0 0 v 0 = 0 p0 +
0
T t
LU
v 0
1 02 0
Sl
v + forces
+ v 0 0 v 0 = 0 p0 +
0
t
Re
v 0
Ro
+ Re v 0 0 v 0 = Re0 p0 + 02 v 0 + forces
t0
|

{z

unsteady term

{z

convective term

{z

inertial terms

| {z }

pressure term

| {z }

viscous term

In the last step we merged the Strouhal number Sl = L/U


with the Reynolds number
T
LU
L2
Re = to get the Roshko number Ro = Sl Re = T . The Strouhal number
is the ratio of the convective flow time scale L/U to the flow time scale T
Sl = L/U
T
and therefore relates the convective and unsteady terms. Only for very small values
of the Strouhal number (Sl  1) the flow is steady and the unsteady term can
be neglected. In our experiments the flow is generally unsteady and the Strouhal
number is bigger than unity.
If the Reynolds number is small (Re  1), the inertial terms also become small
and the flow is dominated by the viscous terms. Nevertheless the convective terms
can become important, if a body is immersed in the fluid. Near the body, the viscous
terms will dominate, but far away from the body the convective terms have to be
considered, since these terms are of order O(Re L2 ) and the viscous terms are
of order O(L3 )(see [VIG07]). If the Reynolds number is high, we can distinguish
between two cases: if the Reynolds number is high, but the Strouhal number is small,
only the convective terms have to be included. These terms are non-linear and cause
very interesting phenomena, but the mathematical handling of the problem gets
difficult. If both the Strouhal and the Reynolds number are high, the unsteady and
convective terms have to be included. According to Lovalenti and Brady [LB93b]
the convective and unsteady effects are not simply additive. If both unsteady and
convective terms contribute to the dynamics, the form of the unsteady term will
change.
2

The Unsteadiness of a flow is also measured by the Roshko number Ro = TL ,


which is the ratio of viscous time scale L2 / and boundary time scale T . The
Roshko number indicates the importance of the viscous terms in relation to the
unsteady terms.

1 Introduction

1.2 Flow around a sphere


Now we move from the general equations to a more specific problem. A particle
is immersed in the fluid and we want to describe the motion of the particle. We
can differentiate between two cases: either the force on the particle is given or the
velocity is known. Our task is either to compute the velocity resulting from a given
force or to compute of the forces on the particle when the velocity is given.
Before we go into the details of the computations, we briefly discuss the flow field
for four different cases. The flow is either at low or high Reynolds numbers and the
sphere is either fixed or freely moving in the flow.

1.2.1 Low Reynolds number flow


In the low Reynolds flow, the flow fields around a fixed and a freely moving sphere in
the comoving frame are the same. Picture 1.1 shows the flow around a sphere fixed
. Here we replace the length scale
in a windtunnel at low Reynolds number Re = av

L by the radius of the sphere a and the velocity scale U by the flow velocity v of
the windtunnel. The freely moving sphere generates the same flow field. The fluid
flows around the sphere and one can think of different flow layers that are slightly
deformed by the sphere. The flow is also called laminar. An important observation
is that the flow field is axisymmetric.

Figure 1.1: The flow around a sphere at low Reynolds numbers. Picture by Taneda
[Tan56] at Re = 4.575.

1.2 Flow around a sphere

1.2.2 High Reynolds number flow


At higher Reynolds numbers the convective terms become important and the flow
field will change due to these non-linear terms. In the laminar flow around the
sphere, the flow had different symmetries, namely axisymmetry and planar symmetry. For higher Reynolds numbers, the different symmetries will be broken. First,
the flow loses its axisymmetry and vortices behind the sphere occur. A further
increase of the Reynolds numbers leads to an unsteady flow field and the vortices
behind the sphere detach. For even higher Reynolds numbers the planar symmetry
is also lost and the flow field is called turbulent.
Figure 1.2 taken from Bodenschatz and Eckert [BE11] shows the turbulent flow
field around a sphere. It was taken by L. Wieselsberger and L. Prandtl in a wind
tunnel. The sphere was fixed inside the measurement section and the flow field was
visualized with smoke.

Figure 1.2: Photo by L. Prandtl. Flow around a sphere in a windtunnel, visualized


with smoke. Taken from Bodenschatz and Eckert [BE11].

Freely moving sphere When the sphere is not fixed, the problem has additionally
six degrees of freedom, namely the three position and the three velocity coordinates
of the sphere. In the low Reynolds number regime, the asymptotic state of the freely
moving sphere is steady and the wakes of the free and fixed sphere are identical. As
shown by Natarajan and Acrivos [NA93] the wake is axisymmetric until a Reynolds

1 Introduction
number Re = ua
up to 105. We will call this case A. At this critical number,

the wake loses its axisymmetry due to a forward Hopf bifurcation. Jenny, Bochet
and Dusek [JBD03] showed that the additional degrees of freedom influence the
symmetry breaking and the wake of the free and fixed sphere cannot be considered
identical anymore. After the first bifurcation the wake loses its axisymmetry, but
the particle trajectory is still planar. For higher Reynolds numbers there are the
steady regime with planar symmetry (105 < Re < 137, 5, which we call case
B) and the unsteady periodic regime (Re > 150, which is called case C). The loss
of the planar symmetry characterizes the chaotic regime, which appears for higher
Reynolds numbers (Re > 250) and will be called case D. For spheres lighter than
the fluid another regime, the zig-zag regime, occurs as shown by Jenny, Dusek and
Bouchet [JDB+ 04]. The pictures 1.3 and 1.4 show the different flow regimes.
The result found in the paper mentioned before were tested in a series of experiments, which we will not discuss completely. An overview was given by Ern et al.
[ERFM12]. The first experiment with freely moving spheres focusing was undertaken by Veldhuis et al. [VBVWL05] in 2005. They filmed the motion of spheres
settling or rising in water. By imposing a small temperature gradient, they were
able to visualize the wake with the Schlieren technique. The plastic spheres had a

radius in the range of 0.75 to 10 mm and the density ratio = fp was between 0.35
a
and 2. The resulting Reynolds number Re = umean
was between 102.5 and 2311.5.

Veldhuis et al. confirmed the existence of the different flow regimes (case A-D) that
Jenny et al. predicted. They showed that the wakes of a freely moving sphere in the
different cases displays remarkable differences when compared to the wakes behind
spheres held fixed. Also, the wake for a sphere lighter than the fluid differs from
the wake of a heavier sphere, even for the same Reynolds number.
Summary We discussed the different flow fields for fixed and freely moving spheres.
At low Reynolds numbers the flow field will be the same for both cases, but if the
Reynolds number is higher the symmetry braking mechanism is different. For the
freely moving sphere, we identified four different cases, which can be characterized
by the number of symmetries. In the first case (A) the wake of the sphere is both
axisymmetric and steady. Case B is characterized by a loss of axisymmetry, but the
wake has planar symmetry and is steady. Case C is unsteady, but the wake is still
planar. Case D has neither temporal nor spatial symmetries.
This result affects the interpretation of our experiment in the following way: The

1.2 Flow around a sphere

(A)

(B)

Figure 1.3: Picture by Veldhuis et al. Flow field visualized with Schlieren rechnique.
The left part of each picture shows the flow in the xz-plane and the right part shows
the flow in the yz-plane. Left: Flow around a sphere in water at Re = 102.5. The
flow field is steady (case A): Right: Flow around a sphere in water at Re = 162.5.
The axisymmetry is broken (case B).

(C)

(D)

Figure 1.4: Picture by Veldhuis et al. Flow field visualized with Schlieren rechnique.
The left part of each picture shows the flow in the xz-plane and the right part shows
the flow in the yz-plane. Left: Flow around a sphere in water at Re = 225. The
axisymmetry is broken and the flow field is oscillating (case C). Right: Flow around
a sphere in water at Re = 985. The flow field is chaotic (case D).

1 Introduction
force measurement is done typically with a sphere fixed in the flow, whereas the
velocity measurement is done on freely moving spheres. Hence, the result might
differ at high Reynolds numbers.
We start with the force measurements at low and high Reynolds numbers. After
this section the velocity measurements will be discussed.

1.3 Force measurement


Lets assume the velocity up (t) of the particle is given. The force f (t) on the particle
will then depend on the fluid and particle properties (fluid viscosity , fluid density
f and particle density p , particle radius a). Also the values of the fluid velocity
v(t) affects the result.
We start with a rather simple problem: the sphere is fixed and we measure the
force F on the sphere. The motion of the sphere should be sinusoidal with the
frequency f and amplitude A. We can then describe the problem with the additional
parameters particle radius a and dynamic viscosity completely. These parameters
have three different dimensions (kg, m, sec) and we can define three independent
non-dimensional parameters. An appropriate choice is:
normalized force
Reynolds number
dimensionless frequency

F
aAf
aup
Rev =

2f a2
a2
S=
=
9
9
F =

(1.2)
(1.3)
(1.4)

We reduced the five dimensional parameter space (F, f, A, a, ) to a three dimensional parameter space (F , Rev , S). Now we have to answer the following question:
how does the normalized force F depend on Rev and S?
The first part of this section focuses on the low Reynolds number regime and we
discuss the different contributions to the force on the sphere in the frequency and
time domain. Then we turn to the high Reynolds number case and examine the
influence of this control parameter.

1.3.1 Low Re solution for the Navier-Stokes equations


Low Reynolds number flow is characterized by the relative unimportance of the
inertial terms. Under the assumption of vanishing convective terms, Stokes was

1.3 Force measurement


able to solve the Navier-Stokes equations for the flow around a sphere. He used the
method of stream functions and derived the hydrodynamic forces on a pendulum,
which he assumed to be correctly described by a sphere. The peak velocity of the
flow should be denoted by U = A. Then the result is given by (see [Sto51]):

F = < 6U a(

1
|{z}

Stokes drag

|{z}

Basset History

1 2
)eit

|{z}

added mass

(1.5)

with the dimensionless parameter


a2
2
= i
= iS.
9
9
The dimensionless parameter denotes a dimensionless (complex) frequency. The
plot 1.5 shows the result for the (angular) frequency = 2 rad/sec (all other
parameters are set to unity). We can identify the different terms in the solution
as the Stokes drag, the Basset memory force and the added mass term. Since they
scale different with and therewith with the frequency , they can be separated in
the frequency domain. This was already discussed in the paper by Lawrence and
Weinbaum [LW86]. The Stokes drag (first term) is dominant for small frequencies,
whereas the added mass term (third term) is dominant for high frequencies. The
second term is the Basset force, which will only give the biggest contribution at
intermediate frequencies. The graph 1.6 shows the scaling of the different forces
individually.

1.3.2 Forces acting on a particle in fluid


In the last section the solution for sphere moving at low Reynolds numbers in a
steady flow was discussed. The total drag on the sphere was given by the sum of
different forces, namely the Stokes drag, the added mass term and the Basset History
force. In this section, we discuss the correct form of these force terms and explain
the appearance of additional forces at higher Reynolds numbers and non-uniform
flow conditions.
So far, five different force acting on a particle moving in a fluid have been discovered (see Brennen [Bre05]). If body forces act on the fluid, they will also act on
the particle inside the fluid. From the Stokes solution we identified three different

1 Introduction

10

Stokes regime

Basset regime

added mass regime

Force (arb. units)

10

10

10

Stokes
History force
Added mass
2

10 1
10

10
S = a2 /9 (nondim)

10

Figure 1.5: The force on an oscillating sphere in dependence of the time. All parameters set to unity.

1
0.8

amplitude (arb. units)

0.6
0.4
0.2
0
0.2
0.4
0.6

Stokes
History force
Added mass

0.8
1
0

5
6
time (arb. units)

10

Figure 1.6: Scaling of the forces on an oscillating sphere as function of the dimen2
sionless frequency S = a9 . All parameters set to unity.

10

1.3 Force measurement


forces that will also act on a sphere moving inside a fluid. The fifth force acting
on the particle appears for non-uniform flows. A non-uniform flow is a flow, which
has not everywhere the same velocity (both magnitude and direction). If the flow
is non-uniform, the flow disturbance term has to be considered.
A typical equation of motion will include the following forces:
body forces Typically the motion is driven by body forces. This can be the action
of gravity on a settling sphere or a rising bubble. In our experiment the motion will
be driven by the shaking of the fluid, which results in an inertial force acting on the
particle.
drag When a particle moves inside the fluid, the particle will be decelerated
due to the action of drag. For Rev = 0 the drag is given by Stokes formula
Fdrag = 6f a(up v) (with viscosity , fluid density f , particle radius a and particle (up ) and flow (v) velocity). For higher Reynolds numbers the drag force has to be
derived from experimental data (for example Allen [All00], Arnold [Arn11], Williams
[Wil15], Wieselsberger [Wie22], Liebster and Schiller [LS24], Liebster [Lie27], Schmiedel
[Sch28]).
added mass When the particle accelerates, one would expect that the surrounding
fluid will accelerate with the sphere. In the equation the added mass term can be
merged with the mass terms to give an effective inertial mass for the sphere and the
fluid. The added mass term will be constant for the parameter range studied.
Nevertheless the constancy for higher Reynolds numbers is still under discussion.
Most authors regard the added mass term as constant, which is also supported by
numerical and theoretical papers (see Auton Hunt and PrudHomme [AHP88], Mei
and Klausner [MK92], Kim, Elghobashi and Sirignano [KES98] and Wakaba and
Balachandar [WB07]). In contrast, Brennen states that the added mass (and drag
term) will be a function of the frequency of the unsteady term and the Reynolds
number (see Brennen [Bre05]). He refers to experiments by Sarpkaya [Sar86] on
cylinders in oscillatory flow, which showed that both added mass and drag terms
vary with Reynolds number and frequency.
History force The history force occurs due to the lagging development of the
boundary layer. At the surface of the sphere, vorticity will be generated and diffuses
away from the sphere. According to Brennen [Bre05], the History force results

11

1 Introduction
from the fact that additional vorticity created at the solid particle surface due to
relative acceleration diffuses into the flow and creates a temporary perturbation in
the flow field. Like all diffusive effects it produces an 1/2 term in the equation for
oscillatory motion. In the Maxey Riley model, the convective terms are neglected
and the history force is modelled by
6a ()
2

1/2

Z t
0

d(u)
d

(t )1/2

d.

(1.6)

In this notation we used the velocity u in the comoving frame (u = up v). The
correct form of the History force kernel is discussed in several publications and many
different forms of the equation of motion have been proposed. The discussion on
the correct form of the History force started with the empirical formula by Odar
and Hamilton [OH64], who proposed different pre-factors for the history force. Note
that the Kernel of (t ) leads to a scaling of in the frequency range. A further
discussion will be given later in the theory chapter. Mei and Klausner [MLA91]
observed in a numerical study that the history force decays much faster than t1/2 .
In subsequent studies Mei and Adrian [MA92] solved the problem with the matched
asymptotic method and a t2 decay was proposed.
Lovalenti and Brady [LB93b] and [LB93a] used the reciprocal theorem to derive
an expression for the hydrodynamic force on a particle in time-depended motion.
They also found a t2 decay for suddenly accelerated particles from rest, but a
exponential decay for the change from velocity u1 to u2 with u2 > u1 > 0. According
to Mei [Mei94], this behaviour was caused by the linearisation of the Navier-Stokes
equations in the analysis by Lovalenti and Brady.
Kim, Elghobashi and Sirignano [KES98] corrected the Kernel for very heavy par
ticles (particle-to-fluid density ratio = fp < 0.2), but the temporal decay is also
t2 . The equation by Kim et al. is much more complex than the model by Mei.
Recently Loth and Dorgan [LD09] published another proposal for the equation of
motion.
In summary, the different history Kernels are given by:
Fhistory = 6f a

12

Z t

K(t )

du
d
d

(1.7)

1.3 Force measurement


with
16(t )
=
a2
"

KBasset

#1/2

"

KM ei

16(t )
=
a2

(1.8)
#1/2c1

4(t )2
+
f H a2
"

Rev
2

3 #1/c1 c1

and

(1.9)
(1.10)

1
fh = (0.75 + c2 Rev )3
2

(1.11)

Loth and Dorgan used the same formula as Mei, but with different constants:
c1,M ei = 2

c2,M ei = 0.0105

c1,Loth = 2.5

c2,Loth = 0.2

Both models result in an faster decay of the history force for long times, but the
decay remains algebraically.
Note that the term long time behaviour of the Kernel is misleading. Of course, the
correction terms will dominate for long times, but they also scale with the Reynolds
number. If the Reynolds number is very small, the long time behaviour remains
t1/2 and therefore the correction terms should be better denoted Reynolds number corrections.
Disturbance of the flow In the literature, the flow disturbance term denotes the
change of a non-uniform flow due to the presence of a particle inside the fluid.
Starting from the sphere settling under gravity in unsteady, but uniform flow, Tchen
[Tch47] derived an equation for non-uniform flow. He replaced the particle velocity
up by (up v) and observed that the presence of a particle will result in a fluid
acceleration. This fluid acceleration requires a pressure gradient throughout the
fluid. In his equation the term for the disturbance of the flow field
Fdisturbance,T chen = mf luid

v
v
Dv
= mf luid
+ mf luid v
Dt
t
x

was added to the equation of motion. The term v denotes the undisturbed flow field.
Corrsin and Lumley [CL56] and Buevich [Bue66] pointed out inconsistencies in the
work by Tchen. They criticized that Tchen used only the static pressure gradient

13

1 Introduction
and not the full Navier Stokes equations. The corrected term proposed by them is
Fdisturbance,CL = mf luid

Dv
2 v
Dt

and
Fdisturbance,Buevich = mf luid

d
v.
dt

d
denotes a time derivative following the particle. In 1983, Maxey
The derivative dt
and Riley [MR83] published their form of the equation of motion, which is now
generally accepted as the best approximation. Gatignol derived the same formula
independently [Gat83]. They concluded that the flow field disturbance is given by

Dv
= mf luid .
Dt X(t)

Fdisturbance,M R

Thereby the total derivative has to be taken at the particle position X(t). This is
the same result, which Tchen derived by an ad hoc extension of the equation of
motion.

1.3.3 High Reynolds number flow


So far we have focused on the dependence of the force F on the dimensionless
frequency S. This section focuses on the dependence on the third parameter, the
Reynolds number. There are two kinds of experimental techniques to measure the
force on a sphere in an unsteady flow at different Reynolds numbers. The first
type of experiments was performed by Odar and Hamilton and later by Karanfilian
and Kotas. The sphere is fixed to a bar and set into motion. The amplitude A
or the frequency f of the bar can be raised to increase the Reynolds number. The
other experimental approach is also straight-forward. The sphere is fixed inside the
measurement section of a windtunnel. The fluid velocity is known and the force on
the sphere is measured. The Reynolds number can be increased by a higher wind
speed v.
This section is organized as follows. First we discuss the results for the sphere
moved inside the fluid. Then we go on to the windtunnel experiments and look at
the results for steady and unsteady flow configurations. The last part gives a short
summary of the conclusions from the force measurements.

14

1.3 Force measurement


Oscillating sphere The hydrodynamic force on a sphere in a viscous fluid was first
measured by Odar and Hamilton [OH64] and the results are still discussed today
(see Michaelides [MR10]). Odar and Hamilton fixed a sphere to a rod, which was
driven by a motor. The connection between the rod and driving wheel of the motor
was adjustable, so that they could vary both the frequency and amplitude of the
motion. Due to the setup with the moving rod, the horizontal motion of the sphere
was not exactly sinusoidal. The speed of the motor was recorded and the force on
the sphere was measured with force transducers placed inside the sphere. Since the
original data are not available, we use the data from the reconstruction in [TKT91]
and [MR10] to estimate the parameters of this experiment. The experiments were
done in an tank filled with oil (viscosity = 79.002 kg/(sec m) and = 889.07
kg m3 ) and a sphere (different constituent parts, mass m = 148 g). The Reynolds
(u is the instantaneous velocity of the sphere) ranged from 0
numbers Rev = au

to 31 and a wide range of frequencies was studied. The dimensionless frequency


2
S = a9 of the measurements ranged from 0 to 6.37.
The aim of their experiment was the determination of correction factors for the
different forces. They assumed that the Reynolds number correction for the drag
force cD is correct. Also they assumed that the history force is zero at t = 34 .
Then they were able to fit the results to get correction factors for the added mass
term cA and the Basset History term cH . Odar and Hamilton proposed the empirical
formula 1.12 for the hydrodynamic forces.
Z t d
u()
4
1
dt

d
F = cD a2 |u|u + cA a3 a + cH a2 ()1/2
2
3
0
t

(1.12)

The correction factors for added mass cA , Basset History cH and Stokes drag cD all
depend on the Reynolds number. The proposed equation was questioned later by
Mei and Adrian [MA92]. The errors in the measurement of the added mass term
have most likely compensated the errors in the history force measurement (and vice
versa). With the assumption that the added mass term is constant, Michaelides
proposed a revised formula [MR10].
Karanfilian and Kotas [KK78] performed an experiment similar to that by Odar
and Hamilton. They also fixed a sphere to a rod and measured the force on the
sphere. In their experiment the oscillation direction was vertically and the sphere
velocity was higher. The amplitude of the oscillation was between 10 and 60 mm and
the angular frequency was between 0.566 and 2.262 rad/sec. The experiments were

15

1 Introduction
done in either water ( = 0.918106 m2 /s) or diesel oil ( = 3.68106 m2 /s) and the
was in
sphere radius was 19.02 mm. The resulting Reynolds number Rev = aup,max

a2
the range 50 to 5000. The dimensionless frequency S = 9 was not explicitly given
in the paper, but with the parameters above the available range was 20 < S < 400.
They measured the (total) drag force on the sphere for different velocities and
accelerations. Instead of the dimensionless frequency S, they used the acceleration
du
number AcOH = 2a
and the Reynolds number to characterize the flow. They
u dt
concluded that the drag coefficient is not a function of the Reynolds number alone.
It also depended on the acceleration number and they gave an empirical formula for
the dependence. The problem with this approach is that the acceleration number
is not constant over one cycle, as pointed out by Tsuji, Kato and Tanaka [TKT91].
The definition of the acceleration number leads to a problem for the oscillating
sphere, since the maximum value goes to infinity during one cycle. Unfortunately,
the parameters at the different data points were not reported in their data, so that
we were not able to check the dependence on the frequency and amplitude variation.
Wind tunnel experiments Now we move on to the wind tunnel experiments with
steady flow conditions. The sphere is fixed in the measurement section of a windtunnel and the forces are measured in dependence on the Reynolds number. As
shown before, the Stokes drag will dominate in the low frequency regime. At higher
flow velocities, the flow field itself will become unsteady and it is not possible to
isolate the different contributions. The result is usually given in terms of the total
drag force
Fdrag cD v 2
with the drag coefficient cD , which is defined as follows
cD =

2Fdrag
(total) drag force
= 2 2.
dynamic pressure projected area
v a

The results are summarized in the following plot 1.7 from H. Schlichtings book
Boundary Layer Theory [SG00]. In contrast to the rest of this thesis, the Reynolds
number in this plot is based on the diameter of the sphere. The plot shows that for
24
small Re the value of cD is proportional to Re
. For intermediate Reynolds number,
the cD coefficient is constant and at high Reynolds numbers a sudden decrease of
cD is apparent. A further discussion would lead to far and we conclude that relation
between force and velocity is very complex. The force on a sphere in unsteady flow

16

1.3 Force measurement

Figure 1.7: Dependence of the drag coefficient of spheres on the Reynolds number.
Curve 1 and 2 represent the prediction by Stokes and Oseen and curve 3 is the
numerical result by B. Fornberg (1988).

conditions was also measured by Tsuji, Kato and Tanaka [TKT91]. They fixed the
sphere in the windtunnel and blocked the flow periodically behind the sphere to
achieve an unsteady flow field. They measured in the range of 4000 < Rev = au
<

8000 with pulsation frequencies from 2.5 to 8 Hertz. The interpretation of their
results is rather complicated. They used the ansatz by Odar and Hamilton (see
[OH64]) and tried to fit the data to get correction factors for the Stokes drag cS ,
Basset history cH and added mass force cA . Under the assumption that the history
force can be neglected, the total drag coefficient cD is given by
4
cD = cS + (1 + cA )AcOH .
3

(1.13)

AcOH is the acceleration factor Odar and Hamilton [OH64] and given by AcOH =
2a du
( u being the velocity of the sphere relative to the fluid). The equation 1.13
u2 dt
indicates a linear relation between the difference cD cS and the acceleration factor,
if the added mass correction factor cA is constant. For low frequencies this statement
is true. Nevertheless for high frequencies, the data deviated from the linear relation,
which the authors explain by the effect of the history force. They conclude that
their results agree with the experimental results by Odar and Hamilton.

17

1 Introduction

1.3.4 Summary
We discussed force measurements for different Reynolds numbers both in the time
and frequency space. At low Reynolds numbers, the force on the sphere is the sum
of the different forces (Stokes drag, Basset history and added mass force), which
scale different with the frequency. Therefore they can be separated in the frequency
domain for the case of sinusoidal motion. The measurements at higher Reynolds
numbers indicate that the low Reynolds number solution needs correction terms.
The form of the corrections and their justification is unclear.

1.4 Velocity measurement


So far we discussed experiments, in which the sphere was fixed inside a flow and
the resulting force was measured. In this section, we discuss the inversion of the
problem. The force f on the particle is specified and the velocity of the sphere up is
unknown. We use the relative velocity u = up v of the particle velocity up and the
fluid velocity v to describe the problem. The particle velocity depends on the radius
a of the sphere, the densities of the particle p and fluid p and the fluid viscosity
. Also the typical time scale of the flow T is important. Since the parameters have
three different dimensions (kg, m, sec), we can choose three independent parameters.
An obvious choice might be:
density ratio
Reynolds number
dimensionless frequency

f
p
au
Rev =

a2
S=
T
=

(1.14)
(1.15)
(1.16)

This choice is not very good, because the velocity is the variable of the problem and
should not be connected to the control parameter. Instead we define a Reynolds
number based on the force F on the particle:
viscous time scale
a2 /
=
force time scale
a/ F a

a Fa
=

Ref =

18

(1.17)
(1.18)

1.4 Velocity measurement


The connection to the Reynolds number is obvious: the velocity was replaced by the

velocity scale F a as noted by [VBVWL05]. The three dimensionless parameters


, S and Ref define the parameter space of the problem. We have to find the
functional dependence of the velocity on these parameters. The three parameters
cover different features of the physics: the density ratio is connected to the material
properties, the dimensionless frequency S defines the dominant response and the
Reynolds number Ref (based on the force) measures the strength of the external
forces relative to the viscous forces. In the literature different definitions of the
dimensionless parameters can be found, but we use the definitions given above, if
not said otherwise.
This section is organized as follows. We start with the low Reynolds number
flows. The Reynolds number is very small and its effect can be neglected. First, we
discuss the solution for the different density ratios and frequencies. Then we go on
to the different experiments. The third part focuses on the high Reynolds number
flow. Here the Reynolds number is an additional control parameter for the problem.
We discuss the previous experiments and define the needs for our experiment.

1.4.1 Special solutions for the Navier-Stokes equations


In the last section, we saw that the different forces acting on the sphere can be
separated in the frequency domain. Lawrence and Weinbaum [LW86] discussed this
behaviour in detail and Sobral, Oliveira and Cunha [SOC07] tested in a numerical
study, if this conclusion is also true for the velocity measurements at given forcing.
Therefore the relative strength of the different hydrodynamic forces was examined
for the sedimentation of rigid particles. The simulation showed that the Basset force
affects the unsteady drag for the early stages significantly. The time was normalized
2
mp
= 92 a fp . For particle density ratios close
with the relaxation time scale r = 6a
to one, the time to reach terminal velocity changes from O(1) to O(100), when the
Basset force is considered. The result for a fluid-particle density ratio of 1/2 is shown
in graph 1.8. The added mass terms dominates for small times (and therefore high
frequencies) and decays after a few r . The Basset term reaches its maximum after
approximately one r , but decays slowly. After 100 r the Basset force is as small
as at the beginning. The Stokes drag increases very slowly and dominates the other
forces at values greater than r 10 (corresponding to small frequencies). Also
they state that the Basset force will change with presence of inertial effects. As the
Reynolds number based on radius and Stokes terminal velocity increases, the sphere

19

1 Introduction

Figure 1.8: The relative strength of the hydrodynamic forces for a settling sphere.
Taken from [SOC07], page 136. The particle Reynolds number was Rep = 0.5.
Sobral, Oliveira and Cunha use to define to density ratio, but the definition is
the same: = f /p . The virtual mass corresponds to the added mass term in
our notation. The time was made non-dimensional with the relaxation time scale
2
r = 92 a fp .

will reach the steady state faster. Due to the high computation costs of the Basset
force, only a qualitative analysis was possible. They conclude that the classical
Basset force is correct for higher Reynolds numbers, if the dimensionless frequency
S is high. The conclusion can be questioned, because Sobral et al. simulated only
the effects of the known forces and included the non-linear corrections for the drag
force. The inclusion of this correction factor is for sure not enough to capture the
non-linear effects at different Reynolds and dimensionless frequencies.
In the case of sinusoidal forcing and therewith oscillatory motion of the sphere,
the first analytical result was given by Coimbra et al. in 2004 [CLL+ 04]. The
authors derived a general solution for the equation of motion (see Coimbra and
Rangel [CR98] or Coimbra and Kobayashi [CK02]) and solved it with a complex wave
ansatz in the laboratory frame. They gave the solution in terms of the dimensionless
2

frequency S = a9 and the density ratio = fp . The velocity of the fluid is given
by Aei S t and the particle responds with a velocity Bei( S t+) . Then the relation

20

1.4 Velocity measurement


between the fluid and particle velocity is given by:
B i
2iS( 1)

e =1+
A
iS( + 2) + + 3 Sei/4

(1.19)

The equation 1.19 contains both the amplitude ratio and the phase difference. The
result is shown in the plots 1.9 and 1.10, which give the dependence on the frequency
S and the density ratio .
2.5

Ampl. ratio (nondim)

=5
=2
= 0.5
= 0.2

1.5

0.5

0 6
10

10

10

10
S (nondim)

10

10

10

Figure 1.9: The amplitude ratio of fluid and particle velocities as a function of the

dimensionless frequency S and density ratio = fp as given in 1.19. Solid lines


with Basset History force, dashed lines without.
The difference between the amplitudes of particle and fluid velocities increases
monotonically with the frequency. Particles lighter than the fluid ( > 1) have a
bigger amplitude and heavy particles have a smaller amplitude than the fluid. For
very small frequencies both types of particles have the same amplitude as that of
the fluid. Particle and fluid are in phase for low and high frequencies. Depending on
the density of the sphere, the phase is different for intermediate frequencies. Heavy
spheres lag behind the fluid, whereas light spheres have a positive phase difference.
The plots show also the influence of the Basset History term: the amplitude ratio
increases faster and the phase anomaly between forcing and sphere motion occurs

21

1 Introduction

=5
=2
= 0.5
= 0.2

[rad]

/8

/8

10

10

10

10
S (nondim)

10

10

10

Figure 1.10: The phase difference between fluid and particle velocity as a function

of the dimensionless frequency S and density ratio = fp as given in 1.19. Solid


lines with Basset History force, dashed lines without.
at smaller frequencies. A detailed discussion can be found in the theory chapter.
In summary, the velocity of the fluid defines the force on the sphere and with
formula 1.19, we can calculate the velocity of the sphere. This is the formula that
gives us the velocity of the sphere in dependence on S and . In the next sections,
we compare this solution with the experimental data to see, if the formula is correct.

1.4.2 Experimental studies


1.4.2.1 Experiments with oscillating spheres
We start with the experiments at low Reynolds numbers. The time scale of the flow
9
T is defined by the oscillating frequency f of the sphere. We choose T = 2f
= 9 .

The parameter space is then defined by the two parameters density ratio = fp and
2
dimensionless frequency S = a9 . Since the forcing on the particle was not reported
in every experiment, we chose the Reynolds number based on the velocity to compare
the experiments. All values have been converted into our notation Re = au
. Figure

1.11 summarizes the parameter space of the different experiments in the literature.

22

1.4 Velocity measurement


4

10

10

Abbad b
Weinstein

LEsperance
0

10

Abbad a
Coimbra

10

Gonzales
4

10 2
10

10

10

10

Figure 1.11: Data sets in the different experiments.


Coimbra et al. [CLL+ 04] and L0 Esperance et al. [LCTR05] measured the response
of a particle to an external shaking in two studies in 2004 and 2005. The first study
focuses on the influences of different fluids, whereas the second study only varies the
particle properties. In both studies they used the same technique. The particle was
fixed with a small copper wire (diameter d = 19 m) inside the fluid cell, which was
mounted on a shaker. The frequency range was 20 Hz to 70 Hz and the amplitudes
last from 250 m to 400 m. The motion of the fluid container and the particle was
recorded with a CCD camera using digital holographic techniques. In the study from
2004 they used three different fluid-particle configurations (Krytox/polypropylene
(PP), Krytox+dryer/Mg alloy, oil/brass) and different particle radii a ranging from
0.5 mm to 2 mm. Therewith the Reynolds numbers based on the radius a and the
relative velocity u = up v was less than 0.5 and the dimensionless frequency S
ranged from 0.015 to 5. The density ratio was between 0.1(brass) and 2.1(PP). In
the second study all experiments were performed with Krytox. The particle size
ranged from 1.98 to 2.23 mm with density ratios = f /p from 0.214 (brass) to
13 (styrofoam). The Reynolds number was less than 5 and S ranged from 1 to 5.
Both studies had the same result: in the explored parameter range the history force
effect is significant and its neglecting leads to errors greater than 10 percent. The
proposed form of the equation is the correct form and at least in the explored range,
there is no need for corrections.

23

1 Introduction
While Coimbra et al. performed horizontal shaking of the fluid cell, Addad and
Souhar designed an experiment with vertical shaking direction. The whole experiment was assembled on a plate and sinusoidally moved up and down by a motor.
Again the motion of the particles was recorded with a high speed camera, but this
time via shadow imaging. The maximal vertical vibration frequency was 10 Hz and
the amplitude was fixed to A = 5 mm. They published two papers on the same
apparatus [AS04a] and [AS04b]. The first study covered the motion of rigid spheres
with different densities and radii ranging from 1 to 2.5 mm. By using silicon oil and
glycerine as fluid and teflon and polyamide spheres, they could measure with density
ratios = f /p from 0.43 to 1.11. The viscosities were oil = 2 104 m2 /s and
glycerine = 6.9 104 m2 /s. Using the definition of the Reynolds number given above,
the Reynolds number in this experiment was less than 1.25. The second experiment
was done with both light bubbles and heavy droplets and the Reynolds number in
this experiment was smaller (Re < 0.25). They report four different experiments.
Two experiments were done with air bubbles rising in silicon oil ( = 5 104 m2 /s)
and glycerin ( = 7 104 m2 /s). The other two experiment included settling glycerine droplets in different silicon oils ( = 5 and 2 104 m2 /s). The density ratio
ranged from 0.77 to 974. Unfortunately, Abbad and Souhar presented their data in
a different way, so the comparison with the data by Coimbra is not possible directly.
They plotted the data in dependence of the penetration depth of vorticity , which
is related to the dimensionless frequency by S = 92 12 . The velocity was normalized

1
by the Stokes terminal velocity uts = 2ga2 ( fp 1) 9
. The graphical extraction
of their data leads to graph 1.12. They found the same result as Coimbra et al,
namely that the History forces cannot be neglected for small Reynolds numbers and
in this range, the form of the equation needs no corrections. The deviation in the
graph 1.12 result almost completely from the graphical extraction, though Abbad
and Souhar estimated their error less than 5 percent.
Another experiment of the same kind was done by Weinstein, Kassoy, and Bell
[WKB08]. They focused on the effects of the history forces on the accuracy of Coriolis flow meters. Therefore they measured the moving of bubble or solid spheres in
a fluid column, which was moved by a shaker table horizontally. They performed
all experiments with an amplitude of 0.35 mm and a frequency of 20 Hz. Additionally all experiments were done with the same kind of oil, leading to a very small
density variation. To achieve a wide range of different Reynolds and dimensionless
frequencies S, different aluminium, steel and POM spheres (0.8 to 4.8 mm) and

24

1.4 Velocity measurement

0.35
0.3

0.2

A /A

0.25

Abbad 1
Abbad 2
with History force
without History force

0.15
0.1
0.05
0
0

0.2

0.4

0.6

0.8

Figure 1.12: Data by Abbad and Souhar


air bubbles with diameter ranging from 0.5 to 4.0 mm were used. Therewith the
highest Reynolds number was less than 1.25 and the dimensionless frequency was
between 0.214 and 0.98. The density ratio = f /p ranged from 0.112 for the
steel spheres to 711 for the air bubbles. Weinstein, Kassoy, and Bell were able to
collapse their data, but instead of the dimensionless frequency S = a
they used the
9u
q q
2
1
inverse Stokes number = 9 S . Nevertheless their results agree very well with
the theory and the two experiments mentioned before. For the solid particles the
highest aberration was 6% , while the values for the bubbles were bigger because of
impurities in the oil.
Most of the experiments suffer from the very low dimensionless frequencies. The
application of a mechanical oscillation with high frequency is not easy. In contrast,
the acoustic forcing of the particle is much easier. Several experiments have been undertaken and we discuss only the experiments with direct connection to our problem.
The study from Gonzales, Hoffmann and Gallego [GHG00] is based on the technique
introduced by Hoffmann and Koopmann [HK94]. They used two loudspeakers to
oscillate glass particles falling in a tube with air. The glass particles had a density
of p = 2.4 103 kg m3 leading to a density ratio of = air /p = 0.5 103 . The
Reynolds number was less than 2.6 and S ranged from 0.01 to 5. They concluded
that for these conditions the data for the extremely low density ratio case are welldescribed by the model using only Stokes drag. However, a careful check of their

25

1 Introduction
data shows a discrepancy from the theoretical prediction at intermediate frequencies. This error can be explained by the effect of the Basset history force. Also the
data for high frequencies are below the prediction, which is most likely caused by
the added mass term. Nevertheless at low frequencies the Stokes drag dominates
the other forces and the agreement between their data and the linear model is very
well.
Garbin et al. [GDO+ 09] were able to measure the history force on bubbles at
frequencies in the MHz range. They propelled a pair of air bubbles (diameter
2m) with ultrasound at f = 2.25 MHz. Then they recorded the motion of the
microbubbles with the ultra-high speed camera Brandaris 128 at frame rates near
15 106 frames/second. The Reynolds number was below 5. Their model included a
system of two coupled equations of motion. The oscillation of one bubble generates
a fluid field and the other bubble will be accelerated in this field. From the motion
of one bubble the forces on the other bubble could be derived. They compared their
results with the prediction for the history force for bubbles proposed by Magnaudet
and Legendre [ML98] and found that the History force has a crucial influence on
the dynamics. They concluded that the disregard of the Basset History force will
lead to an underestimation of the total viscous dissipation in the case of a bubble
moving in uniform unsteady flow. The bubbles have to be separated far enough, so
that the assumption of a uniform flow holds.
The experiments is the frequency domain all verified the equation of motion including the Basset History force. In the next section, we look at the experiments in
the time domain.
1.4.2.2 Experiments with settling spheres
As mentioned, the solution by Coimbra was only obtained in the domain of the
dimensionless frequency S and not in the time domain. Coimbra and Rangel [CR98]
also gave a general solution, but only in term of complex functions. When the
forcing is given by a step function, the solution can be formulated in terms of real
functions, as shown by Vodopyanov, Petrov and Shunderyuk [VPS10], who obtained
the equation for the settling of a sphere in the time domain. They checked their
solution in an experiment, in which they recorded the settling of steel spheres in
silicone oil. The diameter of the spheres was between 0.237 and 0.396 mm, whereas
the viscosity of the silicon was between 6.7 and 8.9 104 m2 /s. Therewith the
Reynolds number based on the terminal velocity and radius ranged from 0.0753

26

1.4 Velocity measurement


to 0.257. They concluded that the Basset equation describes the settling of the
particle correctly. Due to the limitations of their container, they had to include the
corrections for the finite size of the container to obtain the correct result.
Park, Klausner and Mei [PKM95] undertook an experiment with both falling
ruby spheres and rising air bubbles in different fluids. By using distilled water,
water/glycerine solution and ethyl alcohol as working fluids, Reynolds numbers between 13 and 212 were achieved. The results showed an excellent agreement with
the history force proposed by Mei (see equation 1.8 or [Mei94]). This experiment is
not contradicting to the previous results, since we can understand the Mei History
force term as finite Re correction to the Basset History force. We should note that
this experiment does not give an contribution to the discussion of the long time
behaviour rather than a contribution to the Re dependence of the Basset term.
We can conclude that the Basset History force is correct for very low Reynolds
numbers, whereas for higher Re correction terms are needed. The next section gives
an overview of the experiments done at higher Reynolds numbers.

1.4.3 High Reynolds number flow


The parameter space is now defined by the three dimensionless parameters =

f
and the Reynolds number Re = a F a . The parameter space (, Re)
, S = a
p
9
was explored by different experiments. We already introduced the experiment by
Veldhuis et al. [VBVWL05]. In the literature often a combination of the density
ratio and the Reynolds number is used to describe the problem. This new parameter
is called the Galileo number and is given by:
Ga =

viscous time scale


force time scale

a2 /
= q p
=
a/ |( f 1)|F a

(1.20)
|( f 1)|F a3

q
p

(1.21)

For the case of the settling sphere, the force F is given by gravitational acceleration
g and the parameter space is characterized by different flow regimes (see first part of
the Introduction 1.2.2). The experimental studies cover the whole parameter space,
as shown in graph 1.13. Unfortunately, Veldhuis et al. did not measure the velocity
of the sphere, so we can not use their experiment for our discussion.

27

1 Introduction

|( p 1)|F a3

(denoted by G in the plot) and


Figure 1.13: The Galileo number Ga =

density ratio define the parameter space for the motion of spheres under gravity.
The different flow regimes are marked in the phase diagram. Picture taken from
[VBVWL05].
f

We need experimental data, which were taken for defined forces and known flow
fields. If the flow field is unknown, we can not specify the fictitious forces due to the
acceleration of the fluid and the total force acting on the particle is unclear. There
are two types of experiments that fits our restrictions. The first type of experiment
was already introduced. The measurement of a settling sphere in still fluid provides
data for the velocity at a known force (gravity) and velocity field (still fluid). The
other type of experiment is drag measurement in shock tubes. Due to the shock
wave, the velocity of the fluid is increased and this acceleration of the fluid leads
to an fictitious force on the particle. This experiment can also be used, hence the
flow field is known from the measurement of the shock wave and the force can be
computed from the acceleration.
This section focuses on the experiments with high Reynolds number flow, in which
the velocity was measured. We start with the experiments with steady forces and
go on to the unsteady case.
Settling spheres An experiment at higher Reynolds numbers was performed by
Mordant and Pinton [MP00]. The settling motion of different spheres was measured
by the Doppler shift of acoustic waves. Using glass, steel and tungsten spheres with
radii ranging from 0.25 to 3 mm and water as fluid, they achieved Reynolds numbers

28

1.4 Velocity measurement


based on the mean velocity of the sphere umean of 20.5 < Re = aumean
< 3850.

Thereafter, they solved the Maxey Riley equation with the Basset force for the
History force numerically and compared the result with their measurements. For
short times, the settling motion was described correctly, while the memory term
decreased faster than t1/2 for long times. They proposed an experimental fit for
the decay. Their conclusion should be taken with caution, because the decay of
the Basset History force is not measured at constant Reynolds number. A possible
decay of the Basset term couldqresult from the increase of the Reynolds number.

|( p 1)|F a3

for the experiments was in the interval


The Galileo number Ga =

from 15.46 to 1911.86. The advantage the Galileo number is twofold. This number is constant for the whole experiment and it can be computed before. The
Reynolds number based on the velocity increases during the experiment and has to
be computed with an implicit equation (see [MP00]). With the Galileo number, we
can easily identify the different flow regime in comparison with the experiment by
[VBVWL05]. For example, Mordant and Pinton compared the motion of different
(see the graphs in 1.14, which were
particle at the same Reynolds number Rev = au

taken from their paper). The bigger sphere shows transitory oscillations, whereas
the trajectory of the smaller sphere does not have these oscillations. This observation
becomes clear, when the definition Ga is considered. Although both experiments
have the same Reynolds number based on the velocity, the Galileo number, which
is a Reynolds number based on the force, is different. This is why, the trajectories
of the two experiments differ.
f

Shock tube experiments As explained before, we can also use shock tube experiments for our discussion, since the force on the sphere and the flow field are
known. Temkin and Kim [TK80] and Temkin and Metha [TM82] used weak shock
tube experiments to measure the velocity of moving particles at a given force. The
force acting on the particle is given by the fictitious forces due to the acceleration of
the fluid. Temkin and Kim [TK80] measured the drag coefficients for high Reynolds
numbers in a shock tube. The particles were micrometer sized droplets of different
densities and the Reynolds number based on the relative velocity of fluid and particle
was between 1.6 and 38.5. The total drag coefficient for steady flows is known from
other experiments, but the interesting aspect of this experiment was the velocity
change. Due to the shock wave, the velocity of the fluid is raised very fast and the
acceleration of the fluid is big. Since the experiment is done in a short time after

29

1 Introduction

(A)

(B)

Figure 1.14: Picture (A): Velocity measurements for Rev = au


400. The dashed

curve shows the velocity of a 1 mm tungsten sphere and the solid curve represents
a 2 mm glass sphere. Picture (B): Velocity measurements for Rev = au
630.

The dashed curve shows the velocity of a 1 mm steel sphere and the solid curve
represents a 1.5 mm glass sphere.

the shock wave, the unsteady effects can not be neglected. Temkin and Kim used
the Acceleration number
!
2a dUr
p
1
(1.22)
Ac =
f
Ur2 dt
to characterize the unsteadiness of the flow. The variable Ur is the fluid velocity in
the frame moving with the particle. They found that the drag coefficient cT emkin
depends both on the Reynolds and Acceleration number, but the influence can be
simplified:
cT emkin = cD + f (Ac),
where cD is the steady drag coefficient that captures the dependence on the Reynolds
number and f (Ac) represents the functional dependence on the acceleration number. For their experimental data they found a linear function.
The results by Temkin and Kim were tested in a subsequent study by Temkin
and Metha [TM82]. Again, micrometer sized particle were injected into a shock
tube and the particles were accelerated by weak shock waves. The maximal flow
velocity was 10.1 m/sec and the Reynolds number based on the relative velocity of
fluid and particle was between 4.5 and 57.5. The data were taken for both accelerating and decelerating flows. The acceleration number was defined as in equation

30

1.4 Velocity measurement


1.22 and both values smaller and bigger than zero could be tested. Temkin and
Kim found that the drag coefficient is smaller than the steady value for accelerating
flows (Ac > 0) and bigger for decelerating flows (Ac < 0). According to them, this
can be explained by the size of the recirculating region. They also plotted the drag
value for the same Reynolds number, but different Acceleration numbers Ac. An increase of the unsteadiness of the flow (higher |Ac|) lead to an higher drag coefficient.
Schlffel et al. [SBB+ 09] investigated the motion of spheres behind a shock wave.
They used Laser Doppler Velocimetry (LDV) to measure the velocity of aluminium
particles with a radius of 2.5 micrometer in different gases. The particle were accelerated due to the fictitious forces at velocities of 1000 m/s and the Reynolds
p)
decreased from 45 to 0 in the experiments. They tried to
number Re = 2a(vu

fit the experimental data with different models, which did not include the Basset
history and added mass force. Also the unsteady effects due to the change in the
fluid velocity were ignored. The discrepancy between these models and the experiment was quite big and they were not able to explain this result. But especially
the omission of the unsteady effects leads to the big discrepancy, which was also the
case for the experiment by Temkin and Kim [TK80].
An overview of other results from wind tunnel measurement can be found in the
paper by Jourdan et al. [JHI+ 07]. They also report their shock tube experiment
with different plastic spheres (density from 25 to 1550 kg m3 ) and different gases.
The Reynolds number was defined as follows:
ReJourdan =


v
u

2a v up u
t


dup
dt

dup
dt

!2

dup,y
+
g
dt

!2

The velocity up,y is the velocity of the particle in the direction of gravity, which was
perpendicular to the direction of the shock wave. With this definition, the Reynolds
number was in the interval 471 < ReJourdan < 68500. The acceleration number
defined as
2a
dup
Ac =
(v up )2 dt
was between 2.5105 and 3.5102 . They found that the measured drag coefficient
deviates from the steady drag coefficient. They argue that the influence of the
unsteady terms can be neglected due to the small value of the acceleration number

31

1 Introduction
Ac. According to them, this can be seen from the comparison with the experiment
by Karanfilian and Kotas [KK78] (discussion in section 1.3.3). This argument should
be questioned because of two facts: the acceleration number is not constant during
the experiment and for this reason it is not a good choice for the measurement of
the unsteadiness of the flow. Also the data from the experiment by Karanfilian and
Kotas are not very precise due to flow disturbance caused by the rod and the walls.
Unfortunately, the data given in the references are not enough to define the force
variation consistently. In contrast to the settling sphere experiments, the force in
the shock tube experiments will decrease linearly with time.

1.4.4 Summary
In this section we discussed the velocity measurement at a given force on the particles. We started with the solution of the equation for low Reynolds numbers and
went on to the experiments. We saw that the Reynolds number should be defined
based on the forces instead of the velocity in the case of force measurements. The
equation of motion with the Basset History force was verified in a big number of
different experiments. For the high Reynolds number case the experiments showed
that the equation of motion needs corrections. In the case of the particle settling
under gravity, the symmetry braking mechanism can explain the break down of the
low Reynolds number solution, but the implication for the correction of the equation of motion are unclear. The shock tube experiments also could not be used to
improve the equation of motion, since the parameters to describe the unsteadiness
of the flow have not been reported.
The next section gives a comparison between the previous experiments and the
parameter space of our experiment.

1.5 Motivation
The motion of particles in an unsteady flow is only partially understood so far. We
have identified the three control parameters: the density ratio , the dimensionless
frequency S and the Reynolds number Re based on the velocity (for force measurements) respectively based on the force (for velocity measurements). For the velocity
measurements the Galileo number is also a suitable choice. The problems can be divided into two main questions: what is the correct form of the forces and how do the

32

1.5 Motivation
forces change due to a variation of the control parameters, especially the Reynolds
number? At a very small Reynolds number, the appearance of the Stokes drag and
added mass force are quite well understood and the correct terms are experimentally
verified. The correct form of Basset history force is also experimentally verified, but
the interpretation of the physical background is under discussion.
The following plots summarize the previous experiments and the parameter range
of this experiment. When the Reynolds number is very small, only the density ratio
and the frequency characterize the problem. The plot 1.15 shows the parameters
and S, at which previous experiments have been done and which parameterspace
our experiment covers. There are experiments for a wide range of density ratios,
so we decided to expand the frequency range towards smaller values. This range is
important, because the temporal decay of the Basset History force for long times
(small frequencies) is unclear.
The data base for the Reynolds number and Galileo number dependence is very
small. We discussed some experiments, but the parameter space , S, Re has not
been examined in detail. The plot 1.16 shows our available parameter space. The
experiment of Veldhuis et al. only checked the dependence on and Ga, while S
was zero.
Our experimental approach is comparable with the experiments by Coimbra et
al. and Weinstein et al. We oscillate a fluid column with a certain frequency
and amplitude and measure the response of a sphere to this external forcing. This
oscillation is conducted by a linear motor system, which enables us to explore a wide
parameter range in the most interesting parameter subspace (Re/Ga, S). There are
two goals of the experiment: The range of experimental data at low Galileo numbers
will be extended to lower frequencies. The second goal of the experiment is the
analysis of the Reynolds number dependence. By increasing the amplitude of the
linear motor the dependence of the velocity response on the Reynolds number will
be checked.

33

1 Introduction

10

10

Abbad b

Weinstein
LEsperance

own data

10

Abbad a
Coimbra

10

Gonzales
4

10 3
10

10

10

10

10

S
Figure 1.15: Data sets in the different experiments.

0.25
0.2

0.15
0.1

own data

0.05
0 1
10

Veldhuis
0

10

10
10
Ga = (f2 A d3)1/2/

10

10

Figure 1.16: Parameter space for the Galileo number dependence

34

2 Theory
In this section we solve analytically the equation of motion for the sphere moving
in a fluid. The force on the sphere is specified and an expression for the velocity in
dependence on the force is derived. In the last chapter one expression was already
given for the case of an inertial frame and we solve the equation again in the frame
moving with the fluid. We start with the quasi-steady state motion and both the
fluid and the particle motion are modelled with a complex wave ansatz. Afterwards,
we solve the full equation. A short introduction to the fractional calculus is given,
since this is very useful for the solution. The last part summarizes the results of
this chapter.

2.1 The equation of motion


2.1.1 Inertial frame of reference
We start with the equation of motion in the inertial frame of reference. The motion
of a rigid sphere (radius a, density p ) in a fluid (density f , (kinematic) viscosity
) is calculated. The fluid velocity v(t) is specified and the particle velocity up (t) is
unknown. Maxey and Riley [MR83] derived the following equation of motion under
the two assumptions. The first assumption was that the particle Reynolds number
p)
 1). With the angular frequency as representing scale for
is small (Rev = a(vu

2
the velocity gradient we can rewrite the second condition as a  1. The following
equation was obtained for the motion:
mp

dup
Dv
1
d(up v)
= (mp mf )g + mf
6af (up v) mf
dt
Dt
2
dt
6a ()
2

1/2

d(up v)
d
d
0 (t )1/2

Z t

35

2 Theory
We used the short notation for the mass of the particle mp =
4 a3
mf = 3f .

4p a3
3

and the fluid

2.1.2 Non-inertial frame of reference


Since a moving object is studied (the particle inside the fluid), it is natural to study
the motion in the comoving frame of reference, which is an non-inertial frame. We
choose the reference frame, which is moving with the fluid. Let the fluid velocity in
the frame moving with the fluid be denoted by v(t), then this velocity v 0 from
the definition of the coordinate system. The velocity of the particle in the comoving
frame u(t) is related to the velocity in the inertial frame by u(t) = up (t) v(t). The
external forces will now appear as inertial forces and the terms change as follows:
3

The inertial force term is now given as (p f ) 4a


3

dv
dt

The term due to the disturbance of the flow disappears, because

D
v
Dt

= 0.

The drag, added mass and history force do not change, though up (t)v(t)
has to be replaced by u(t).
As a result, the equation of motion changes to:
Z t
du
du
4a3
1
du
2
1/2
d
mp
= (p f )
v 6af u mf
6a () f
d.
dt
3
2
dt
0 (t )1/2

It can be rearranged to:


Z t
du
du
4a3
1
2
1/2
d
+ 6af u + 6a () f
d
=
(

)
v
(mp + mf )
p
f
2
dt
3
0 (t )1/2
r
du
1
du
9 Zt
9
d
(p + f )
+
f
d + 2 f u = (p f )v
(2.1)
1/2
2
dt
2a
2a
0 (t )

Equation 2.1 was first derived by Basset in 1888 [Bas88] for the case of a sphere
settling under gravity. Therefore, the equation of motion 2.1 will be referred to as
the comoving Basset equation or just Basset equation.

2.2 Quasi - steady state solution for oscillating sphere


This chapter presents a solution in the comoving reference frame, whereas the detailed solution in both reference frames is given in the appendix.

36

2.2 Quasi - steady state solution for oscillating sphere

2.2.1 Assumptions and Conditions


We start with the assumption that have to be fulfilled so that the equation is valid:
low Reynolds number Rev < 1: The solution is only valid for small particle
. This restriction was also formulated by Maxey
Reynolds numbers Rev = ua

and Riley [MR83].


big Strouhal number Sl = a
> 1: Let the typical time scale of the motion
9u
be given by a (angular) frequency . We replace the time scale T by the scale
9
and get the following dimensionless Navier Stokes equations:

Sl
St Rev

1 2
u
+ uu = p +
u + forcing
t
Rev

u
+ Rev uu = p0 + 2 u + forcing
t

The unsteady term is kept, whereas the non-linear convective terms are neglected. Hence, the following restriction has to be satisfied
Sl Rev > Rev
Sl > 1
a
a
=
>1
y
9u
9Ap
a

>1
9Ap
Exemplary, this condition can be found in [CK02].
Now we move to the conditions, under which the solution of the equation is valid:
quasi-steady state:We have to assume that the motion of both the particle
and the fluid can be described by a complex wave ansatz. This is only true
after the transient state. In 2.5.3 a solution including the transient state is
given.
known initial conditions: We assume that the velocity at t = 0 is zero.
Also the velocity for t < 0 has to be zero.
no wall effects: In order to solve this equation, it is assumed that the walls
have no hydrodynamic effect on the sphere.

37

2 Theory

2.2.2 Solution with complex wave ansatz


Now we solve the Basset equation 2.1 for the complex wave ansatz for both sled and
particle velocity. This ansatz is justified for the conditions given above. We start
with the evaluation of the Basset History force integral and obtain with the ansatz
u = Ap e(it+) :
Z t

du(t0 )/dt0 0 Z t d(Ap e(it+) )/dt0 0


dt =
dt
(t t0 )1/2
(t t0 )1/2
0

= Ap ei ieit

A detailed derivation can be found in the appendix. The other terms are straight
forward and with v = As e(it) and = fp we get:

9
9
1
(it+)
(it+)
+
iAp e
+
Ap e(it+) = (1 )As ie(it)
(1 + )Ap ie
2
2a
2a
The exponential term can be dropped and the terms are rearranged to get the input
on the left side.
Ap e

1
9
i(1 + ) +
2
2


9
i + 2
2
a
2a

= (1 )As i

which gives
(1 21 )i
Ap i
q
e =

As
i(1 + 12 ) + 92 a2 i +
=

(2.2)

2a2

2iS(1 )

iS(2 + ) + 3 iS +

In the last step we introduced the non-dimensional frequency S

(2.3)
a2
.
9

2.3 Characteristics of the resulting equation


We are interested in the physical conclusions of this result and therefore we look at

the solution for different density ratios = fp and different values for the dimensionless frequency S, which also captures the dependence on the particle diameter
a and the fluid viscosity .

38

2.3 Characteristics of the resulting equation


Influence of the frequency We start with the dependence on the frequency and
mark, which parts result from which input.
2iS(1 )
Ap i
e =
2
As
iS( + 2) + 3 S (1 + i) +
2
|
{z
} |
{z
}
added mass

history force

(2.4)

|{z}

Stokes drag

For a very low frequency S  1 all terms containing the dimensionless frequency
S have a very small influence, thus the motion is largely determined by the Stokes
drag. For intermediate frequencies all three terms contribute. For very high frequencies S  1 the added mass term will dominate.
This is well in agreement with the results found by Lawrence and Weinbaum for
the motion of spheroids, see [LW88]. Also LEsprance et al. [LCTR05] found a
corresponding result. Another interesting observation is the different behaviour of
the phase difference. For very high frequencies the result will be proportional to i
and therefore the phase difference will be /2. For very low frequencies, the real
part will dominate and the phase difference goes to 0.
Influence of the density ratio From the formula 2.4 the influence of the density

ratio = fp cannot be easily seen, thus it has to be plotted. Figure 2.1 shows the
p
effect on the amplitude ratio A
. The dashed lines show the result without History
As
force, whereas the solid lines show the results including the History force. With
the density ratio deviating from unity, the amplitude increases. Both small and
large result in increased amplitude, whereas the effect for light particle ( > 1)
is greater than for heavy particles ( < 1). For high frequencies the amplitudes are
the same with or without the Basset History force. The effect of the History force is
a shift towards higher frequencies. This means if the History force is included, the
amplitude increases slower with increasing frequencies.
The result for the phase is shown in figure 2.2. Again the result for small and
large frequencies coincide. For small frequencies the particle and fluid are in phase,
whereas the phase difference goes to /2 for high frequencies. We start the discussion
with the curves without Basset History force. The effect of is only a shift of the
frequency, at which the phase difference jumps to /2. The form of the curve
remains unchanged. If the Basset History force is considered, the slope of the curves
is smaller. The phase difference starts to deviate from 0 at smaller frequencies and

39

2 Theory

1.4

Ampl. ratio (nondim)

1.2
1

=5
=2
= 0.5
= 0.2

0.8
0.6
0.4
0.2
0 6
10

10

10

10
S (nondim)

10

10

10

Figure 2.1: Amplitude ratios for different density ratios. The solid (dashed) lines
show the result including (excluding) the History force.
reaches /2 later. Also the form of the curves for different is changed. The phase
difference increases faster for heavy particles ( < 1) than for light particles ( >).

[rad]

/2

=5
=2
= 0.5
= 0.2

/4

0 6
10

10

10

10
S (nondim)

10

10

10

Figure 2.2: Phase for different density ratios. The solid (dashed) lines show the
result including (excluding) the History force.

40

2.4 Introduction to fractional calculus

2.4 Introduction to fractional calculus


In the previous section the Basset equation was solved by the application of the
complex wave ansatz. To get a more general solution for the transient case, the
fractional calculus is needed. In this section the fractional analogue for some basic
mathematical concepts like derivative, integral and Laplace and Fourier transformations are defined.

2.4.1 Different definitions of the fractional derivative


The history of fractional calculus started with a letter from LHopital to Leibniz in
1695. In this letter he discussed the result for the n-th derivative of a function f (x)
d2
and asked what happened for an non-integer n. The symbol dt
2 is a short notation
for the calculation that has to be done twice:
d d
d2
f (t) =
f (t)
2
dt
dt dt
This might raise the question, which calculation has to be done twice to obtain
d
d1/2 d1/2
f (t) = f (t)
1/2
1/2
dt dt
dt
1/2

d
Several mathematician have given definitions for the operator dt
1/2 , but the discussion will be restricted to those definitions that are used in this text. A more detailed
discussion can be found e.g. in the textbooks by S. Das [Das11] or R. Hilfer [He00].

Riemann-Liouville First, the Cauchy formula for the n-th integration (a function
f should be locally integrable on (a, ), n-times) is defined
Z
|

...

Z t

{z a}

f ()d =

Z t
1
(t )(n1) f ()d = I n
(n 1)! a

n times

by introducing the gamma function (n) = (n1)! we can extend the formula above
to > 0:

a It

1 Zt
=
(t )(1) f ()d
() a

(2.5)

41

2 Theory
The notation of the three indexes for the operator is essential. The sub-index before
the operator gives the lower bound, whereas the sub-index after the operator gives
the upper bound. The function has to depend on one of these two sub-indexes. Two
cases can be defined:

a It

1 Zt
(t )(1) f ()d
=
() a

t Ib

right hand

and
1 Zb
( t)(1) f ()d
() t

left hand

The operator I is generally considered as Riemann-Liouville fractional integral. Let us consider a number m = n with 0 < < 1 and n being an integer
number. The fractional derivative for the right hand definition is then given by
m
a Dt f (t)

dn
n
f (t)
a It
n
dx

(2.6)

The left hand definition is straight-forward, however we will not use it. In fact, we
will use only the right hand one-half derivative from zero to t:
1/2

d1
1/2
f (t)
0I
dx1 t
Z t
1
f (1) ()(t )1/2 d
=
(1 1/2) 0

0 Dt f (t) =

df ()
1Zt
d1/2 f (t)
dt
d
=
0 (t )1/2
dt1/2

(2.7)

In the third step we used the permutability of integration and derivation. Also we
d1/2
defined the meaning of the operator dt
1/2 in the last step.

2.4.2 Laplace Transform of fractional derivatives


In order to solve differential equations of fractional order, the derivation of the
Laplace Transform of the fractional integral and derivative is useful. The Laplace
transformation of a function is defined by:
L{f (t)} =

42

Z
0

est f (t)dt = fe(s)

2.4 Introduction to fractional calculus


The Laplace transform of the first order derivative of the function f(t) is given by
L{f 1 (t)} = s1 fe(s) f (0)
Thereby we used the initial condition f (0). Another important property of the
Laplace transform is the convolution theorem:
Z

L{f (t) g(t)} = L{

f (t )g()d}

= fe(s)ge(s)
The convolution in the time domain is replaced by a multiplication in the s-domain.
As stated, the Riemann-Liouville definition is used only. We follow the procedure given by Gorenflo and Mainardi [GM08] to define the Laplace transformation.
1
t 2
and use the convoluTherefore we start with the definition of a function = (
1
)
2
tion theorem to observe
1 (t) f (t) =
2

Z t
0

(t ) 2
f ()d
( 12 )

The integral kernel has to vanish for t 6 0 and we get for the half derivative
1 Zt
1 Zt
1/2
21
(t

)
f
()d
=
(t )1/21 f ()d = 0 It f (t)
( 12 ) 0
( 12 ) 0
In the last step we used the Riemann-Liouville definition given in 2.4.1 and we found
a convolution of two functions, namely and f (t), which give the half-derivative of
f (t). With the Laplace transform1 of t1/2 :
1/2

L{t

}=

1 1/2
s
2

 

we observe a very useful formula for the Laplace transformation of the fractional
integral:
1/2

L{0 It f (t)} = s1/2 fe(s)


With the definition of the Laplace transformation of a differentiated function we
1

for example Doetsch [DN74].

43

2 Theory
observe
L{

d 1/2
1/2
f (t)} = L{0 Dt f (t)} = s1/2 fe(s) s1/2 f (0)
0I
dt t

(2.8)

2.4.2.1 Fourier transform of fractional derivatives


The Fourier transform defined as
1 Z
0

F {f (x)} = f () =
f (x)e(ix ) dx0
2
has the useful property that the n-th derivation (n integer number) of the function
f (t) is transformed into a multiplication in the Fourier space:
(

dn
f (x) = (i)n f()
dxn
)

This statement can be generalized for fractional derivatives, as shown by Das [Das11].
The Fourier transformation of the fractional derivative leads to the following expression:
F () 0 D f (t)
The equation shows that a fractional order derivative leads to an decay in the frequency domain with the order of the fractional derivative. The connection to the
decay of the kernel in the Basset History force is given by the definition of the
fractional derivative.
1/2

0 Dt f (t) =

df ()
d1/2 f (t)
1Zt
dt
d
=
0 (t )1/2
dt1/2

(2.9)

We can conclude that a decay of the Basset History kernel with t1/2 leads to a
decay with 1/2 in the frequency domain.

2.5 General solution for the Basset equation


By using the complex wave ansatz, the interesting initial motion of the sphere was
eliminated from the solution. In the following, the transient case is of interest and
therefore, a general solution for the Basset equation will be derived.

44

2.5 General solution for the Basset equation

2.5.1 Laplace Transformation of the Basset equation

To solve the Basset equation, we first perform a Laplace-Transformation of the


equation in the form given in 2.1. We start with the decomposition of the fluid
acceleration v into an acceleration amplitude F and time dependent part g(t), which
leads to the formula v(t)

= F g(t). The non-dimensional equation is observed by


2 +1

rescaling the times with = 92 a p f2 f and the velocities with = pp+ 1 f F .
2 f
Therewith we get the rescaling formulas for the velocity u = u0 and time t = t0 .
With the definition of a function q 0 (t0 ) = g( t), we can rewrite the forcing term to
get F g(t) = F q 0 (t0 ). Finally, we observe:

du 9
f
d1/2 u
f
9
+
u=
+
1
dt
2 a p + 2 f dt1/2
2a2 p + 12 f

du0 9
d1/2 u0
f
9
f
+
+ 2
u0 =
1
0
1/2
01/2
dt
2 a p + 2 f
dt
2a p + 12 f

p f
v
p + 12 f
p f 0
F q
p + 12 f

du0
d1/2 u0
+

+ u0 = q 0
dt0
dt01/2
In the last step we introduced the parameter =
primes and perform the Laplace transformation:

p
9
.
2 p + 12 f

(2.10)
Now lets drop the

du
d1/2 u
L
+ 1/2 + u = L {q}
dt
dt
(

e
s ue(s) + s1/2 ue(s) + 1 ue(s) u(0) u(0) s1/21 = q(s)

ue(s) =

e
u(0) (1 + s1/21 )
q(s)
+
s + s1/2 + 1
s + s1/2 + 1

e
The initial condition u(0) is time-independent, whereas the function q(s)
can be
time-dependent. We have to find the following functions (convolution theorem for
Laplace transformation):

u(t) = f (0) u0 (t) +

Z t
0

q(t ) u ()d

(2.11)

with
(

u0 (t) = L

(1 + s1/21 )
s + s1/2 + 1

and

u (t) = L1

1
s + s1/2 + 1

(2.12)

45

2 Theory
To get the solution for the motion of the sphere, we have to perform the inverse
Laplace transformation for these two functions. Before we do this calculation, we
discuss the physical meaning of these functions. The function u0 (t) describes the
decay of the initial condition u(0). If the initial conditions is zero, the convolution
of the function u (t) and the (non dimensional) forcing q(t) gives the velocity. This
means the function u (t) is the response function for this system. Next, we make
a short excursion to discuss the meaning of the response function in terms of the
dielectric response. Thereafter we go back to our problem and solve the equation
2.11.

2.5.2 Excursion: Dielectric relaxation and the response function


We started with the equation 2.10 and found that the solution can be written in
form of a contribution due to the initial conditions and a contributions in form
of a response function. The concept of a response function is new in the field
of fluid dynamics, but there a similar systems, in which the response functions is
discussed and interpreted for long times (see P. Debye [Deb34]). As shown by R.
Hilfer [Hil02], the equation 2.10 can be used to describe the dielectric relaxation
of glass forming materials. The velocity u(t) is replaced by the polarization P (t),
the forcing is replaced by the electric field E(t) and the response function is the
dielectric susceptibility e . R. Hilfer introduced two fitting parameters, which we
call 1 and 2 . Also the order of the fractional derivative was a fitting parameter.
Now the differential equation for the polarization reads:

dP
d P
+ 2 + P = 0
1
dt
dt

(2.13)

The solution of this equation leads to an equation for response function e :


P (t) = 0

Z t
0

E(t )e ()d.

(2.14)

The symbol 0 defines the electric permittivity of free space. In the Laplace space
the response function is given by (see Hilfer [Hil02]):
e (s) =

46

1
.
1 + (2 s) + 1 s

2.5 General solution for the Basset equation


This result was used to fit the experimental data. The real 0 and imaginary part
00 of the susceptibility were measured for a wide frequency range at different temperatures. The results are given in figures 2.3, respectively 2.4, in which the crosses
represent the data points and the lines show the according fit. Note the good agreement between the model with only three fitting parameters and the data.

Figure 2.3: Real part of the susceptibility (taken from [Hil02])

Figure 2.4: Imaginary part of the


susceptibility (taken from [Hil02])

Summary We saw that the Basset equation 2.10 can be used to model the dielectric
response of glass forming materials. With a small number of fitting parameters, the
behaviour of the material for a wide range of frequencies and temperatures can be
modelled. This result shows the connection between the dielectric susceptibility e
and the response function u defined in 2.12. We conclude that the motion of the
particle is the response to the external forcing. The dependence of the response
function on the control parameters is our main interest.

2.5.3 Solution for oscillatory motion


After the short excursion, we go back to our problem and solve the equation. Therefore we have to perform the inverse Laplace transformation. The solution was already given by Gorenflo and Mainardi [GM08]. An own derivation can be found in
the appendix. The following result is obtained
1 Z x0 t
e Idx0
u (t) =
0
1
x0 2 sin( 12 )
1 Z x0 t
=
e
dx0 .
1
0
(1 x0 )2 + x0 2 + 2(1 x0 )x0 2 cos( 12 )

(2.15)
(2.16)

47

2 Theory
Figure 2.5 shows the result for the integration with = 1. The response function
1

u (nondim.)

0.8

0.6

0.4

0.2

0
0

4
6
time (nondim.)

10

Figure 2.5: Result for the inverse Laplace transformation


decreases monotonically from 1 to 0. Note that the response function depends only
the density and not on the length or velocity scales. Before we can numerically
integrate the solution, we have to recover the dimensions of the equation. We use
the initial condition u(0) = 0 and get the following equation:
u = u(0) u00 +
=0+

Z t
0

Z t
0

q 0 (t ) u0 ()d

q 0 (t ) u0 ()d

The response function has no units and the unit of the normalized force is recovered
with the formula q 0 (t0 ) = g(t0 ). The following result is obtained:

1
u=
g (t ) u0 ()d

0
p f

Z t
F 1
p + 12 f
=
g (t ) u0 ()d

0


Z t
p f
1
F
g

(t

)
u0 ()d
=
1

0 p + 2 f
Z t

(2.17)

To perform the integration we have to specify the forcing on the particle. The next
part shows the result for a sinusoidal forcing with the amplitude A.
Example: Sinusoidal forcing We specify the time dependent part of the forcing
as g(t) = sin(t) and the normalized force amplitude as F = A 2 . This ansatz

48

2.5 General solution for the Basset equation


corresponds with a sinusoidal forcing on the particle with amplitude A and (angular)
frequency . The insertion of this ansatz into formula 2.17 leads to:
u=

Z t
0

!
Z
0 12
x
sin( 21 )
1
p f

0t
2
x

A
sin
(t

)
e
dx0 d
1
0 21
0
2
0
2
0

0
p + 21 f
(1 x ) + x + 2(1 x )x cos( 2 )

(2.18)
q

1.10
= 1.768, A = 1 m, = 1 seconds
Figure 2.6 shows the result for = 92 1.10+0.50.9659
and = 2. The graph shows an oscillation, which is superposed with a small drift
towards smaller values. The drift is caused by the fractional term and decreases
fast.

0.3

vel [m/s]

0.2

0.1

0.1

0.2
0

10

time [sec]
Figure 2.6: Result for the transient state

Comparison with quasi-steady state Since we already solved the equation for the
quasi-steady state, we can compare this result with the result from the full solution.

49

2 Theory
First, we rewrite the full solution in terms of and S. We have
9
2
2
9
( + 12 ) 2
S= 2
9
=

2
and get = 0.9659
= 0.878 and S = 11.768
= 2.183. These parameters lead to an
1.1
9
Ap
amplitude ratio of As = 0.0547 and a phase shift of = 0.8784 rad. Figure 2.7
shows the full solution and the quasi-steady solution. The amplitude goes almost
to the value of the quasi-steady within the first oscillation. The amplitude of the
full solution is slightly smaller, which might be caused by round-off errors in the
numerical integration. Nevertheless the amplitude stays constant in both cases.
The phase difference shows also shows the a different behaviour. Due to the initial
conditions the phase differs at the beginning. After a few oscillations the phase of
the two solutions coincide.

0.3
Full solution
Quasi steady

vel [m/s]

0.2

0.1

0.1

0.2
0

time [sec]

Figure 2.7: Result for the transient state

50

10

2.5 General solution for the Basset equation

2.5.4 Summary
We solved the equation of motion for the quasi-steady state and the transient case.
The quasi-steady solution results in a harmonic motion of the particle and the
amplitude and phase of this motion depend on the density and the dimensionless
frequency S. The full solution is a rather complicated sum of a part depending on
the initial conditions and a part, which can be identified as response to the forcing.
The resulting trajectory is a harmonic motion with additional shift at the beginning.
We found that the amplitude of the two solutions reaches the steady value within
the first oscillation. The values for the amplitude slightly differ for the two solutions.
The phase difference of the two solutions also differs at the beginning, but the phase
of the full solution and the quasi-steady solution coincide after a few oscillations.

51

3 setup - censored

53

4 Results
The section is organized as follows: The first section covers the classical problem
of an oscillating sphere in quasi-steady state. We discuss our results and compare
them to the previous work by Lesperance [LCTR05], Abbad [AS04a] and Weinstein
[WKB08]. The second section focuses on the influence of the Reynolds number Ref
and the dimensionless frequency S. We will discuss in which parameter range the
results of the quasi-steady state solution are valid. Thereby all the discussion is in
the comoving frame.

4.1 Quasi-steady state


We start with the case of low Reynolds numbers and check the dependence of the
sphere motion on the dimensionless frequency S and the density ratio . In the introduction, we saw that these parameters are suitable to describe the time variation
of the boundaries and the material properties. In the theory we solved the equation
of motion for the quasi-steady state and the transient case. Before we look at the
results of the experiment, we compare the two solutions for typical parameters of
the experiment. In the second section we check, if the assumptions for our solutions
are fulfilled in our experiment. Then we go on to the results of the experiment and
compare them with the results reported in the literature. Thereafter we go on to
the high Reynolds number experiments.

4.1.1 Full solution versus Quasi-steady state


As shown in the theory part, the full solution and the quasi-steady solution differ
only slightly. Figure 4.1 shows the resulting velocity versus time plots for typical
parameters of the experiments. Since the quasi-steady solution does not include the
transient part, the two solutions differ for the first oscillations. For longer times the
phase coincide and the amplitude of the full solution is only slightly smaller. Figure

55

4 Results
4.2 shows the position versus time plot. Again, the results differ at the beginning,
but the difference between the two results become constant after a few oscillations.
We can conclude that the sphere reaches a quasi-steady oscillation within the first
oscillation. If the shift in the position is removed, the quasi-steady and full solution
agree within the accuracy of the experimental data. We will therefore use the
quasi-steady solution to fit the data and discuss the error due to this assumption
afterwards.
Full solution
Quasi steady

x 10

1.5
1

vel [m/s]

0.5
0
0.5
1
1.5
2
0

time [sec]

10

Figure 4.1: Particle velocity from the full and quasi-steady solution for S = 0.0279.
Parameters: As = 5mm a = 2 mm, = 100 106 m2 sec1 and = 2 rad sec1 .

4.1.2 Assumptions
The steady state solution given in Chapter 2 is only valid, if the assumptions (steady
state, known initial condition and no wall effects) and constraints on dimensionless
parameters (Sl > 1 and Rep < 1) are fulfilled. Before we look at the results, we
have to check, if these assumptions are violated.
The steady state assumption is fulfilled, since we cut off at least one oscillation
before fitting the data. This is in good agreement with the full solution, which
showed that the sphere reaches the quasi-steady state within the first oscillation

56

4.1 Quasi-steady state


(see last section). The initial conditions was that the velocity of the sphere (in the
shaking direction) is zero. This condition is not exactly fulfilled, but the noise was
much smaller than the signal. Therefore we can assume that the initial condition
u = 0 was fulfilled approximately.
The condition on the container size needs a bit more attention. The walls of the
container will become important, if the ratio of the container size D to the sphere
diameter d defined as = Dd is small. Then the wall effects will cause measurement
errors and the Faxen corrections have to be included. The critical value for the sphere
to container size has been estimated by several authors. Tritton [Tri88] states that a
value of = 100 will cause an error less than 2%, whereas Lindgren [Lin99] estimates
the error for = 50 with 4 %. Both estimates are valid for very small Reynolds
 1. For our apparatus the value of is between 30 for the
numbers Rep = ua

biggest sphere and 50 for the smallest sphere. Since the non-inertial terms will scale
as 2 = Rep (see Veysey [VIG07]), this number should be small. The value of 2
is in the interval 4.40 105 to 6.46 103 for the experimental data shown in this
section. Therefore the walls will have a very small effect on the oscillating motion.
The conditions on the Reynolds and Strouhal number can be checked easily from the
Full solution
Quasi steady

x 10

ampl [m]

5
4
3
2
1
0
0

time [sec]

10

Figure 4.2: Particle position from the full and quasi-steady solution for S = 0.0279.
Parameters: As = 5mm a = 2 mm, = 100 106 m2 sec1 and = rad sec1 .

57

4 Results
data. The Reynolds number should be much smaller than unity and the Strouhal
number should be greater than unity. The particle Reynolds number takes values
= 9Aa p
between 2.9 104 and 4.3 102 . The Strouhal number defined as Sl = a
9u
is between 1.1 to 7.6. Both conditions are fulfilled.

4.1.3 Measured trajectories


The biggest errors can be divided into two classes: systematical errors due to the
limitations of the experiment and measurement errors. For the measurements with
slow frequencies, only a few oscillations could be recorded due to the settling of
the sphere. Especially the phase detection was sensitive to short data tracks. At
measurements with very high frequencies, we could measure more oscillations, but
here the amplitude of the oscillation had to be reduced to keep the Reynolds number
low. Therefore the signal to noise ratio decreased and the fit to the data became
worse. Again, the phase was more affected by this effect. The measurement errors
are caused by bad data quality. Figure 4.3 shows the data quality of three different
measurement runs, all correspond to the oscillation of a 5 mm sphere driven by
a sled oscillation of 12.5 mm. Due to inadequate illumination, the quality of run
two is worse than the other two. Run two was performed with LEDs leading to
weaker illumination and therefore opened aperture. Also the light was diffused by
translucent paper, causing overall a more dispersed illumination than the brighter
Halogen lamp. Due to the bad quality of measurement run two, the data from this
run were not used. Nevertheless, the usage of the Halogen lamp also caused errors,
since the power varies with the switching frequency of 100 Hertz. This signal is
apparent on all data tracks taken with Halogen lamp illumination. Another problem
was the focus of the lens, which had to be readjusted after each measurement day.
The oscillation and vibration of the sled lead to a shift of the focus. But the error
due to the out-of-focus is much smaller than the error due to the illumination.

58

measured
fit

0.2

0.15
0.1

0.1

displacement (mm)

displacement (mm)

0.15

0.05
0
0.05
0.1

0
0.05
0.1
0.15

0.15
0.2
4

0.05

7
t (s)

10

0.2
8

measured
fit
8.5

(a)

9.5

t (s)

10

10.5

11

11.5

(b)
measured
fit

0.15

0.05
0
0.05
0.1
a: run 1, f = 1 Hz

0.15
0.2
9

b: run 2, f = 1.05 Hz

10

11

t (s)

12

13

14

c: run 3, f = 1.05 Hz
The other parameters are the same.

59

(c)

Figure 4.3: Quality of different measurement runs

4.1 Quasi-steady state

displacement (mm)

0.1

4 Results

4.1.4 Results
The result for the quasi-steady state experiment is shown in figures 6.1 and 6.2 for
the different spheres and sled frequencies. The amplitude of the sled motion was
different, but always so small that the conditions for the solution were fulfilled.
The results for the amplitude are very close to the predicted curve with the Basset
History force. We can confirm that the Basset History force is needed to describe
the behaviour of the particles correctly. Especially for the high values of S the
importance of the Basset History force is evident. The data for the phase scatter
much more than the data for the amplitude. We already explained the difficulties
in the measurement of the phase. Nevertheless the data support the importance of
the Basset History force.
From the full solution we learned that the quasi-steady solution gives a good
prediction for the phase, whereas the amplitude in the full solution is a bit smaller.
The error due to the quasi-steady assumption is much smaller than the variation of
the data points. We can conclude that quasi-steady solution describes the results of
the experiment correctly.

0.04
0.035
Ap/As (nondim)

0.03
0.025

3 mm
4 mm
5 mm
with Basset History force
without Baset History force

0.02
0.015
0.01
0.005
0

10

10

Figure 4.4: Result for the amplitude ratio (logarithmic)

60

4.1 Quasi-steady state

/4

[rad]

/8

3 mm
4 mm
5 mm
with Basset History force
without Baset History force

0.02

0.04

0.06

0.08
S

0.1

0.12

0.14

0.16

Figure 4.5: Result for the phase difference


Estimation of the error The typical size of the error bar is shown in Figures 4.6
and 4.7. The error was calculated with the formula for error propagation from the
individual errors.
We estimated the individual errors as follows. The error for the radius was measured with a slide gauge and is less than 10%. The frequency was checked with the
internal oscilloscope of the motor. The error can be neglected in comparison with
the error for the radius. The viscosity of the oil was measured with an commercial
viscosimeter and is also less than 10%. The amplitude of the sled was measured
very precise (error less than 0.1%), but the amplitude of the sphere has an error
of the order of 10%. A typical sphere amplitude was 1-5 pixels and the error after
fitting was 0.1 pixel. Also the density of the spheres varied. The manufacturer (TIS
GmbH) states the density of the spheres with 1.10 103 kg m3 , but we measured
densities from 1.08 to 1.14 103 kg m3 . For the experiments only spheres with densities from 1.09 to 1.10 103 kg m3 were used, which gives an error for the particle
density of 1%. The density of the silicone oil was measured with a precision scale
and measuring cylinder. Overall the error for the fluid density can be estimated to
be less than 0.1%. The error due to the density has to be added to the error of the
amplitude ratio.

61

4 Results

0.04

Ap/As (nondim)

0.03

with Basset History force


without Baset History force
data

0.02

0.01

0.01

0.02

0.04

0.06

0.08
S

0.1

0.12

0.14

0.16

Figure 4.6: Result for the amplitude ratio with error bars.

/4

[rad]

/8

0
with Basset History force
without Baset History force
data
0.02

0.04

0.06

0.08
S

0.1

0.12

0.14

0.16

Figure 4.7: Result for the phase difference with errors bars.

62

4.1 Quasi-steady state

4.1.5 Comparison with the data in the literature


Now we compare our results with the results in the literature. We plot the amplitude
2
p
versus the dimensionless frequency S = a9 for the different density ratios
ratio A
As

= fp . Figure 4.8 shows only a big data cloud at S = 0. The comparison is


better seen in the logarithmic plot (see Figure 4.9). Our data confirms the previous
investigations and extends the range by one order to smaller values of S. The data
points for all experiments agree well with the prediction and we can conclude that
the equation of motion has to include the Basset History force.
Nevertheless the data quality could be improved. If one looks carefully at the
data by Weinstein [WKB08], the deviation from the prediction is quite big for the
heavy spheres. This can be explained by the relative high Reynolds number Ref
(see table 4.1). If the Reynolds number is of the same order as the dimensionless
frequency S the condition on the Strouhal number might be violated (Sl > 1). The
other experimental data show variational data quality, but they all agree with the
predicted curves.
The logarithmic plot 4.9 shows that this experiment extends the data range by
almost one order of magnitude. The low frequency part is particularly interesting
because of the different proposals for the correct form of the Basset History force.
If the long time behaviour is incorrect, the data points should deviate from the
curve for small frequencies (and therefore long times). Since the data points are on
the curve, the decay of the Kernel with t1/2 is confirmed and the classical Basset
History force is correct for small values of the Reynolds number.
source
Weinstein 1
Weinstein 2
Weinstein 3
Abbad 1
Abbad 2
LEsperance 1
LEsperance 2
Own 3 mm
Own 4 mm
Own 5 mm

radius [mm]
0.5, 0.75
0.4, 1.0, 1.6
1.6, 2.4
25
25
1.98
1.98
1,5
2
2,5

[m2 sec1 ]
(0.84 0.95) 104
(0.78 0.88) 104
(0.83 0.84) 104
6.9 104
2.0 104
0.4 104
0.4 104
1.0 104
1.0 104
1.0 104

f [Hz]
20
20
20
0 10
0 10
20 70
20 70
0.14 5
0.2 5
0.2 5

0.112
0.3070
0.606
0.5714
0.8517
0.56
0.214
0.8473
0.8473
0.8473

Rev
0 0.5
0 0.5
0 0.5
0 1.25
0 1.25
05
05
4
4.1 10 6.2 102
2.5 103 1.7 102
2.7 103 1.4 101

Table 4.1: Parameters for the measurements reported in 4.8.

63

4 Results

0.9
0.8
Weinstein 1
Weinstein 2
Weinstein 3
Abbad 1
Abbad 2
LEsperance 1
LEsperance 2
Own data 3mm
Own data 4mm
Own data 5mm
with History force
without History force

0.7

Ap/As

0.6
0.5
0.4
0.3
0.2
0.1
0
0

3
S

Figure 4.8: Comparison with the literature data


Summary and application of the results Our results agree with the results in
the literature and we concluded that the equation of motion given by Maxey and
Riley with the classical Basset History kernel is correct. This result is correct for low
Reynolds numbers and dimensionless frequencies S from (at least) 103 to 10. There
is no evidence that the solution becomes incorrect beyond this frequency range.
This result is important as a starting point for further discussion and experiments.
The effect of the different forces can be estimated with the equation. In the introduction we saw that the importance of the different terms changes with the value of
S. We saw that -in general- Stokes drag, Basset History force and the added mass
term are needed to describe the motion correctly. But for some special cases one
of the terms dominates the other, so that they can be neglected without causing a
big error. If the dimensionless frequency S is very low (S  1), the Basset History
and added mass term can be neglected and the error due to this assumption can
be calculated from the complete solution. For the high frequency case (S  1)
the Stokes drag and Basset term might be neglected without causing a big error.
This observation simplifies the calculations in the two cases dramatically. Also the
simulation of particles in flows get much easier, if the Basset History force can be
neglected.

64

4.2 High Reynolds number regime

10

Ap/As

10

10

10

10 3
10

10

10
S

10

10

Figure 4.9: Comparison with the literature data (logarithmic). Same labels as in 4.8
The correct equation of motion is also important for experiments and applications.
The equation might be used to calculate the motion of small particles in a low
Reynolds number flow, for example small pollen at low wind speeds or sand particles
in water. Also the motion of tracer particles in experiments is an important question.
If the Reynolds number based on the particle radius and relative velocity is small
enough, this equation can be used to calculate the relative motion of the particle
and the fluid. The effect of a higher Reynolds number will be discussed in the next
section.

4.2 High Reynolds number regime


In the previous sections, we discussed experiments, in which the Reynolds number
p
was very low and its effect on the experiment was negligible.
Ref = a2f Ap = aA

Now, we increase the Reynolds number and look at the full parameter space , S and
Ref . Before we go into the details of the results, we look again at the dimensionless
Navier Stokes equations, plug in the force F = As 2 sin(t) and use the time scale

65

4 Results
T = 9 , the velocity scale U = Ap and the length scale L = a to get:
U u0 U 2 0
U 02 0
0 0
u

u
=
u + As 2 sin(t)
+
T t0
L
L
u0 Ref 0
1
9As
u 0 u0 = 02 u0 +
sin(t)
+
0
t
S
S
Ap
u0
1
1
1
sin(t)
+ u0 0 u0 = 02 u0 +
0
t
Sl
S
Fn
Ap
, which we
In the last step we defined the new dimensionless parameter F n = 9A
s
call the force number. In the introduction we related every force to the viscous
forces, whereas now we use the unsteady term as criterion. The Strouhal number
Sl = ReS f = 9Aa p was already introduced before. It is the ratio of the unsteady terms
and the convective terms. For values of Sl  1, the convective terms dominate
the unsteady terms. The dimensionless frequency S was also discussed before. For
S  1 the viscous terms are very important in comparison with the unsteady terms.
If the frequency S is constant, we can replace Sl by Ref .
Ap
The force number F n = 9A
can be seen as the ratio of two Reynolds numbers
s
based on the sled and particle amplitude:

aAp /
Ref
Ap
Ap a/
=
=
=
9As
9As a/
9aAs /
Res
inertial f./viscous f.
=
external f./viscous f.

Fn =

We can say that the force number relates the external force to the inertial force.
For F n  1 the external forces are much bigger than the inertial forces. Another
related parameter is the force time ratio f , which is defined as follows:
external force time
L/U
=
unsteady force time
T
a/As
a
=
=
1/
As

f =

The force time ratio relates the time scales of the unsteady term with the time
scale of the external forcing. For f  1 the external forces can be neglected in
comparison with the unsteady terms.
Now we face the problem that we have to compare the relative importance of four
terms (unsteady, convective, viscous and external forcing). We approach the topic

66

4.2 High Reynolds number regime


as follows. First we vary the amplitude of the sled motion As and focus on the
convective and unsteady terms. We keep the ratio of unsteady and viscous term (S)
constant and check the influence of the convective terms (Ref ). Then the amplitude
As is kept constant and we vary the frequency S. In this case all terms vary. We
prepare experiments with the same parameters (S, Ref ) as before and discuss the
importance of the force number. The amplitude ratio of sled and particle motion is
measured for a wide range of frequencies and compared to the low Reynolds number
solution (Ref  1). The third section gives a short overview of the density effects.
In the last section we summarize the results and discuss the physical implications.

4.2.1 Constant frequency


At this point, we should emphasize that the following experiments are a new feature
and have not been reported in the literature. So far the validity of the Basset
equation was checked for small Reynolds numbers, but we ask, if the solution derived
for small Reynolds numbers (Ref  1) is still valid for flows, in which the convective
terms are important (Ref & 0.1). Our experimental approach was to keep the
frequency constant and vary the amplitude As . With this approach the Reynolds
number Ref varied, whereas the dimensionless frequency S stayed constant.

1.6

data S = 0.0112
data S = 0.0873
theory

1.4
1.2

A /A (normalized, nondim)

1.8

1
0.8

0.05

0.1

Ref

0.15

0.2

0.25

Figure 4.10: Normalized amplitude ratio in dependence of the Reynolds number Ref .
The amplitude ratio was normalized with the theoretical value (Ap /As )0 .
We measured the ratio of the particle and sled motion and calculated the the-

67

4 Results
oretical values with the linear theory presented in the last chapter. If the linear
p
normalized with the
theory (derived for Ref  1) is correct, the amplitude ratio A
As
Ap
theoretical value ( As )0 should be unity.
Figure 4.10 shows the normalized amplitude ratio in dependence on the Reynolds
number Ref for two different values of S. The data points labelled with S = 0.0112
are measured with 4 mm spheres and vibration frequencies of 0.4 Hz. The oscillation
amplitudes increased from 25 mm to 150 mm. The second data set was taken at a
frequency S = 0.0873. These data points were taken with a 5 mm sphere at f = 2
Hz and the amplitude increased from 5 mm to 30 mm. The data shows that the
linear solution is okay for small values of Ref , but if the Reynolds number increases,
the experimental data and the prediction deviate. For Re 0.2, the deviation is
Re
1
= S f was
more than 60% of the expected value. The inverse Strouhal number Sl
between 0.877 and 1.707.
0.04
0.035

Ap/As ( nondim)

0.03

with Basset History force


without Basset History force
data points, horizontally shifted

0.025
0.02
0.015
0.01
0.005
0
0.005
0.01
0

0.02

0.04

0.06

0.08

0.1

Figure 4.11: Normalized amplitude ratio in dependence of the dimensionless frequency S. The data points have been shifted for better visibility. The data points
in each cluster have the same value for S.
For better comparison with the other experiments, we plot the errors in the S
vs amplitude ratio diagram (see Figure 4.11). The error bars were calculated with
the same assumptions as before (see section 4.1.4). Note that the data were shifted
for better visibility. Both data sets show that the error for the amplitude is almost

68

4.2 High Reynolds number regime


constant. The analysis of the error calculation showed that the error in this direction
is dominated by the error from the density variation. The error due to the amplitude
measurement is very small. The error for the dimensionless frequency increases
with both the radius and the frequency . For the measurement with the value
S = 0.0112 the error for S is very small.

4.2.2 Constant amplitude


It is also possible to increase the Reynolds number by increasing the frequency, but
keeping the amplitude As constant. Since the amplitude ratio increases monotonically with the dimensionless frequency S, a higher frequency leads to a higher particle
amplitude Ap and therewith higher Reynolds number Ref . Figure 4.12 shows the
results for different frequencies at an sled amplitude of 5 mm. The data deviate from
the prediction for higher S and the linear theory breaks down for smaller spheres at
lower frequencies. Figure 4.13 shows the same data set in a linear plot with error
bars. Like before, the error for the dimensionless frequency increases with the radius
of the sphere and the frequency . The error for the amplitude ratio increases very
slow with the frequency , but still the biggest contribution comes from the error
for the density ratio. An important observation is that the error is smaller than the
deviation from the theory.
0.12

Ap/As (nondim)

0.1

0.08

3 mm
4 mm
5 mm
with Basset History force
without Basset History force

0.06

0.44
0.4

0.38

0.22

0.23

0.17

0.12

0.04
0.06

0.02

0.1

0.08
0.03

0.006
0.005
0.007

0.01

0.017

0.028

0.04

0.018

0.01

10
S

Figure 4.12: Amplitude ratio in dependence of the dimensionless frequency. The


numbers next to the symbols are the Reynolds numbers Ref .

69

4 Results

0.14
0.12

with Basset History force


without Basset History force
data

Ap/As (nondim)

0.1
0.08
0.06
0.04
0.02
0
0.02
0

0.05

0.1

0.15

0.2

0.25

Figure 4.13: Same data as 4.12, but linear plot and added error bars.
To collapse the data, we normalize the amplitude ratio with the predicted value
(Ap /As )0 from the last chapter and plot this versus the inverse Strouhal number
Sl1 (see figure 4.14). The normalized amplitude ratio gives the deviation from
the (linear) theory and figure 4.14 shows that this deviation is proportional to the
inverse Strouhal number. Since the Strouhal number is the ratio of the convective
to the unsteady terms, we can conclude that the data deviation is proportional to
the strength of the convective terms (in comparison with the unsteady terms).

4.2.3 The role of the forcing


We saw before that the parameter space can be defined by S and Ref . Figure 4.15
shows the trajectory of two particles with different diameter, but almost the same
Reynolds number (Re5mm = 0.185/Re3mm = 0.224) and dimensionless frequency
(S5mm = 0.0873/S3mm = 0.0785). The amplitude of the motion was normalized
with the sled amplitude As . Also the time was normalized with the time scale
= 1/. According to the theory both normalized amplitudes should be (almost)
the same in the dimensionless form.
Since the two trajectories do not coincide, the experiment is not described com-

70

4.2 High Reynolds number regime


pletely with only Ref and S. Figure 4.16 shows the parameter space Ref , S and the
colour marks the deviation from the prediction given by the normalized amplitude
p
p
/( A
) . For the same Ref and S we can prepare experiments, which are
ratio A
As
As 0
quite well described by the linear theory or not (see figure 4.15). The difference in
these two experiments is the strength of the forcing. Figure 4.17 shows the normalized amplitude in dependence on the inverse Strouhal number and the colour marks
the force given by the force time ratio f . It compares the external force time scale
L/U = a/As to the unsteady time scale 1/. If the unsteady time scale is much
bigger than the external force time scale (f  1), the motion of the sphere is dominated by the unsteady terms. In this case the linear theory gives a good estimate
for the amplitude ratio. Nevertheless with increasing inverse Strouhal number Sl1
the convective terms become important and the deviation from the linear theory
increases slowly.
If the external forces are of the same order than the unsteady terms (f 1),
the deviation from the linear theory increases fast with increasing strength of the
convective terms (increasing Sl1 ). Only for very small values of Sl1 , the linear
theory can be used. For Sl1 0.7 and f 0.2 the linear theory gives a prediction

Ap/As (normalized, nondim)

10

0
0

0.5

1
1.5
2
1/Slf= a Ap / *1/S

2.5

Figure 4.14: Amplitude ratio in dependence of the inverse Strouhal number Sl1 =
Ref
p
= aA
. The amplitude ratio was normalized with the theoretical value
S
S
(Ap /As )0 .

71

4 Results

normalized displacement (non dim)

0.2

3 mm
5 mm

0.15

0.1

0.05

0.05
180

190

200

210

220
230
t* (non dim)

240

250

260

Figure 4.15: Particle trajectories for different force number f n, but same Ref and
S.
that is by a factor of two too small.

72

4.2 High Reynolds number regime

0.5

Ap/As (normalized, nondim)

8
7

Ref (nondim)

0.4

6
0.3

5
4

0.2

3
0.1
0
0

2
1
0.05

0.1
0.15
S ( nondim)

0.2

0.25

10

0.25

0.2

0.15

0.1

0.05

0
0

a/ As

Ap/As (normalized, nondim)

Figure 4.16: Reynolds number Ref in dependence on the dimensionless frequency


S. The colours indicates the deviation from the prediction, which is given by the
p
p
/( A
).
normalized amplitude ratio A
As
As 0

0.5

1
1.5
2
1/Sl = a A / *1/S
f

2.5

Figure 4.17: Normalized amplitude ratio in dependence on the inverse Strouhal number Sl1 . The colours indicates the strength of the forcing given by Aas .

73

4 Results

0.6

displacement (mm)

0.4
0.2
0
0.2
0.4
6.5

7.5

t (s)

8.5

Figure 4.18: Particle trajectory at Ref = 0.229.


Change of the trajectory Figure 4.15 showed the trajectory for different experiments at different forcing and one could observe that the trajectory becomes nonharmonic at high forcing. Now we discuss this effect in more detail.
Figure 4.18 shows the trajectory of the particle motion for a sled frequency of 5
Hz and an amplitude of 5 mm. The particle had a radius of 3 mm and the Reynolds
number based on the forcing Ref was 0.229. The plot clearly shows a deviation from
the sinus motion of the particles at low Reynolds numbers. The same conclusion can
be made from the spectrum 4.19. Due to the resolution of the spectrum, the first
harmonic and the sled frequency do not coincide exactly. Figure 4.19 shows that the
second and third harmonic give a notable contribution to the spectrum. This can
only be explained by non linear effects. Linear forces like the added mass, Basset or
Stokes force give only a contribution to the base frequency (first harmonic). Since
the Reynolds numbers Ref is & 0.1, the convective terms give already a notable
contribution and these non-linear terms can explain the appearance of higher order
harmonics.
Summary We saw that the Reynolds and the Strouhal number are not enough to
describe the motion of the particle. In addition to the strength of the convective and
unsteady terms, we have also to look at the external forces. We can conclude that
the data deviate from the linear prediction because of two effects. If the convective
term becomes large in relation to the unsteady term, the linear theory is not valid.

74

4.2 High Reynolds number regime

Intensity [dB]

100

1 st
2 nd
3 rd

200

300
sled frequency

400
0

10

20

f [Hz]

30

40

50

Figure 4.19: Spectrum of the particle motion shown in 4.18. The dashed line shows
the sled frequency.
The strength of the deviation is controlled by the force time ratio f . If f  1, the
external forces can be neglected and the deviation from the linear theory increases
slowly with the strength of the convective terms (increasing Sl1 ). If the external
forces can not be neglected (f 1), the linear theory is valid only if the convective
terms can be neglected. If both the external force and the convective terms need to
be included, the deviation from the theory increases very fast.
In the next section we check the dependence on the density of the particle and
then we summarize and discuss our results.

4.2.4 High Reynolds number and density variation


It is also possible to test the validity of the linear theory in dependence of the
Reynolds number and density ratio. As noted before, the variation of the density

ratio = fp introduces another dimension to the parameter space (S, Ref , ). Figure 4.20 shows the results for 3 mm steel and plastic spheres. All measurements
were done the same amplitude (5 mm). The numbers next to the data points are
the Reynolds numbers for that measurement. Figure 4.20 shows that the solution is
valid in the same frequency range for plastic and steel spheres. Given an amplitude
of 5 mm, the solution will be valid until Scrit 0.06. Another observation is that
the solution for the steel sphere is valid for higher Reynolds numbers Ref than for
the plastic spheres.

75

4 Results
From our experiment we can conclude that spheres with a diameter of 3 mm will
follow the prediction until a dimensionless frequency of S 0.06. Figure 4.21 shows
the errors for this experiment. In comparison with the previous experiments the
relative error for the amplitude ratio is smaller, since the density of the steel spheres
is much bigger and therefore easier to measure. Also the amplitude of the motion
is bigger, which leads to better signal to noise ratio. Nevertheless the plot shows
a large error bar, but the relative error is small. The error for the dimensionless
frequency is the same as for the plastic spheres.

4.2.5 Summary and application of the results


We saw that a particle in a flow with low Reynolds number Ref & 0.1 can have
different trajectories. If the force time ratio F is small, the deviation between
the linear theory and the measurement is small and we can say that the particle
follows the flow. If the force time ratio is high, the particle trajectory becomes
non-harmonic and the deviation between the predicted amplitude and the measured
amplitude becomes very large.
It was shown that the internal parameters like the particle amplitude Ap and
frequency or the viscosity are not enough to describe the motion of the particle. Also external parameters like the sled amplitude As were needed to predict
the correct result. The measurement of the Reynolds number alone might not be

0.8
0.7
0.6

Ap/As (nondim)

0.5
0.4

steel
plastic
steel with Basset History force
steel without Basset History force
plastic with History force
plastic without History force

1.15
0.94
0.76

0.48

0.3

0.34

0.2
0.1
0

0.22
0.005

0.06

0.01

0.12

0.007

0.1
0.2
0.02

0.03

0.04

0.05
S

0.06

0.07

0.08

Figure 4.20: Validity of the quasi-steady solution in dependence of the density ratio
and Reynolds number. Plot for the amplitude

76

4.3 Discussion

Ap/As (, normalized, nondim)

0.8
0.6

steel with Basset History force


plastic with Baset History force
data points

0.4
0.2
0
0.2
0.02

0.03

0.04

0.05
S

0.06

0.07

0.08

Figure 4.21: Same data as in Figure 4.20, but with error bars.
sufficient to characterize a flow in nature or experiments.
An application of our results might be possible for the response of rain droplets
to horizontal wind. The size of the droplet is comparable to the size of our particles.
Even though the Reynolds number based on the droplets radius and the relative
velocity (in the horizontal direction) might be small, the particle might be displaced
substantially by the wind, if the force time ratio is high. In general, a flow with
higher force time ratio leads to bigger particle amplitudes for the same Reynolds
number. This is explicitly important for experiments with tracer particles. The
experimentalist should ensure that both force time ratio and Reynolds numbers are
small, otherwise the particle will not follow the fluid. This leads to a problem, since
the fluid velocity might not be measurable directly.
Another conclusion is that the density seems not to be a control parameter.
Spheres with the same diameter, but different densities showed a clear deviation
from the theory at the same critical values. This effect is counter-intuitive and
needs further experimental verification.

4.3 Discussion
In the motivation, we identified two main questions: what is the correct equation
of motion and how does this equation depend on the control parameters? The first
question is answered with the experiments at low Reynolds numbers, whereas the

77

4 Results
second question was answered with high Reynolds number experiments.
Low Reynolds number experiments In the introduction the different approaches
for the equation of motion have been discussed. There are at least five different
proposals for the correct form of the equation. The previous experiments showed
that the equation of motion given by Maxey and Riley with the Basset Kernel for the
History force is correct and we also used this equation. The data base was expanded
towards lower frequencies and we saw that the quasi-steady solution is also valid for
smaller frequencies. The correct form of the Basset History term was questioned
by several authors, but from our experiments we can conclude that Basset History
term is correct. The solution was tested over a wide range of frequencies and there
is no evidence for a break down of the solution at very high or low frequencies.
Nevertheless the equation is only valid for small Reynolds numbers, as shown in the
second part.
Besides, the full solution was compared to the quasi-steady solution, but the
difference was very small for our experiments. We can assume that the sphere
reaches the quasi-steady motion within the first oscillation.
High Reynolds number experiments We concluded that the Basset History force
is correct for small Reynolds number, but the correct form of the equation for higher
Reynolds numbers remained unclear. Several approaches based on numerical solutions of the Navier Stokes equations have been proposed, but the physical justifications for these approaches were not given.
Our experiments showed the dimensionless frequency S and the Reynolds number
Ref are not enough to describe the motion of the particle. These parameter describe
only the dependence on the unsteady and convective terms in relation to the viscous
terms, but we saw that the external forcing is also important. The external forcing
can be measured by the force time scale f , which has to be small, so that the
linear prediction is valid. If the external force is large, the trajectories show a clear
non-harmonic shape.
We concluded that the linear theory is only valid, if the Reynolds number is small
(Ref < 0.1) and the force number time scale is also small (f < 0.04). This result
is very important, since the role of the forcing was not emphasized in the literature
before. As long as these two parameters are small enough, the theory is valid for all
values of the dimensionless frequency S.

78

4.3 Discussion
Measurement errors At this point, we should also take a critical look at our
experimental procedure. We estimated and discussed the errors before (see section
4.1.4).
In the case of the low Reynolds number experiments, the error bars for the dimensionless frequency S were much smaller than the error for the amplitude ratio. This
result is consistent with the experiment. In the low Reynolds number regime, the
amplitude of the particle motion was very small and the overall error due to fitting
was of the order 1%. Besides, the error due to the density measurement, which was
much bigger, was added. The error due to the density could be reduced with better
measurements and every experiment should be done with the same sphere, since the
density of the spheres varied, too. Nevertheless, even with a known density ratio,
the amplitude ratio is still affected by the error due to the measurement of the particle trajectory. One way to improve the data quality is a better illumination. We
were able to improve the data quality (see section 4.1.3), but maybe an even better
measurement is possible with a brighter lamp and better mirrors and lenses.
The high Reynolds number experiments are characterized by larger errors for the
dimensionless frequencies S. As discussed before this error increased both with S and
a. Since the frequency is already known very precisely, only a better measurement
for the particle radius and the viscosity could improve our data quality. The error
for the amplitude ratio is also larger. Since the error for the density stays constant,
the additional errors is caused by the amplitude error. A bigger amplitude leads to
a larger error, since the error scales with the amplitude. Nevertheless, the relative
error is better for higher amplitude because of the better signal to noise ratio.
Beside these errors, we have to discuss the systematical errors. As mentioned
before, the lenses are slightly defocused after one measurement day. Most likely this
comes from the vibration of the sled. The error due to the defocusing was estimated
from the videos. Compared to the other errors, this error can be neglected. Another
important question is the vibration of the camera itself. For very high frequencies
of the sled motion, the whole table vibrates and the camera (including the lenses
and mirrors) might vibrate as well. Unfortunately, we were not able to measure this
error, since no suitable measurement procedure is available. Targets on the sled will
vibrate and targets standing outside the measurement section cannot be focused.
We had to assume that the camera did not vibrate, but we cannot finally evaluate
whether this assumption is fulfilled.
Furthermore, we assumed that the whole fluid column follows exactly the sled

79

4 Results
motion. At every measurement run, we assured that no air is in the container, but
the internal motion of the fluid could not be measured. This assumption needs also
further attention.
Beside the correction of the mentioned sources of errors, the experiment has several
possibilities for further improvements. The next section gives a short overview of
possible modifications and interesting setups.

80

5 Outlook
The correct description of motion of a particle inside a fluid remains unclear. The
parameters of the problem can be grouped into three aspects: the material properties of fluid and particle, the time variation of the boundary conditions and the
forcing. We started with the definition of dimensionless parameters to characterize
the different aspects and defined the three parameters density ratio , dimensionless
frequency S and the Reynolds number Re. For velocity measurement the definition
includes the velocity, whereas the definition was based on the forcing for force measurements.
The first two experiments focused on the dependence on S and . In the literature
several experiments were reported and we expended the parameter space for the
dimensionless frequency towards lower frequencies. Our experiment, as well as all
experiments conducted before, showed that the equation of motion should include
the Basset History force with the classical Basset Kernel. Future experiments could
use more viscous fluids to expand the data range even further. Also the approach
by Garbin et al. [GDO+ 09] is very interesting. The ultrasonic excitation of bubbles
could be used to expend the data range towards higher frequencies.
The transient solution was tested, but the effect was small in the range of the
parameters explored. Therefore the parameter range of the experiment should be
increased towards higher frequencies. Also experiments with slower decreasing response functions could be undertaken. However, due to the fractional contribution
of the Basset History force all response functions are self-similar and the influence
of the transient part is weak.
In the third part of the results, we reported experiments at higher Reynolds numbers. We measured the deformation of the particle trajectory due to non linear
effects and the parameter space (, S and Re) was explored for the first time. Subsequent studies should be done to explore the parameter space completely.
From this starting point, a wide range of different aspects can be investigated.
The theory was restricted to Newtonian fluids, spherical particles and still fluid. The

81

5 Outlook
fluid could be replaced by a non Newtonian fluid or some additives could be used.
The resulting changes of the particle trajectory might lead to a further understanding
of these material properties.
Also the experiment could be repeated with non spherical particles. This will
lead the new kinds of memory force terms, as shown by Lawrence and Weinbaum
[LW88]. A numerical study with a slightly eccentric fluid spheroid moving in an
viscous medium was reported by Abbad, Souhar and Cabillina [ASC06]. The effect
of the shape of the particle is very important for industrial applications, for example
the mixing or the sedimentation of ingredients (see for example Dietrich [Die82]).
The sedimentation of particles leads to another interesting question: the interaction
between particles. The experiment could be performed with more than one particle
to examine the interaction. Also the interaction between the walls and the particles
are a recent research topic (for example Ten Cate et al. [TCNDVdA02]).
Our experiment was only performed with heavy particles, but the motion of bubbles is also very interesting. As noted in the introduction, the wake of a bubble
shows different characteristics. Another question is the possible deformation of the
bubble due to the unsteady flow. As noted by Garbin et al. [GDO+ 09], the motion
of microsized bubbles in blood is an important question for medical applications.
The physics of bubble motion is also interesting because of the interface between
bubble and fluid. The bubble can be deformed or vary the size because of the fluid
motion, but it is also possible that the bubble acts as rigid particle (see Abbad
and Souhar [AS04b] respectively Garbin et al. [GDO+ 09]). Another aspect is the
texture of the particle. Porous particles might show a different behaviour.
Of course, the sphere could also be replaced by a cylinder. On this topic experiments can already be found in the literature (see for example Sarpkaya [Sar86]). The
periodic forcing of the cylindric steel legs is a serious problem in the construction
of offshore platforms. The experimental setup might be interesting to achieve high
Strouhal and Reynolds numbers at the same time.
Last but not least, the fluid could be stirred during the experiment to see the
influence of a non uniform flow.

82

6 Appendix

6.1 Solution in the inertial frame


We use the equation according to Tchen [Tch47] and L0 Esperance [LCTR05]. Since
we are only interested in the motion in the shaking direction, the gravitational term
is neglected.
v(t
1 ) u(t
1)
4a3
2a3
a Zt
4a3
dt1
p v =
f u
f (v u)
6a (v u) +
3
3
3

t t1
Z t
1
9
9 a
v(t
1 ) u(t
1)

0 = p v f u + f (v u)
+
(v

u)
+
dt
1
2
2 a2
2 a2
t t1
Z t
p + 1/2f
9
9 1
v(t
1 ) u(t
1)

v 3/2 u +
(v

u)
+
0=
dt
1
f
2 a2
2 a
t t1
Z t
v(t
1 ) u(t
1)
1
9
9 1

dt1
0 = (1 + )v 3/2 u +
(v u) +
2
2
2a
2 a
t t1
"

Now we use the following ansatz:


u = As eit
u = As ieit
v = Ap eit+0
v = Ap ieit+i0
Before putting this into the equation, we plug the ansatz into the integral term
and get:
Z t itt

e dt1

i
= ... = eit i
t t1

(see Tchen 1947)

83

6 Appendix

9
1
0 = (1 + )Ap ieit+i0 3/2 As ieit +
(Ap eit+i0 As eit )
2
2 a2
v(t
1 ) u(t
1)
9 1 Z t

dt1
+
2 a
t t1

1
9

0 = (1 + )Ap iei0 3/2 As i +


i(Ap ei0 As )
(Ap ei0 As ) +
2
2 a2
2 a
In the next step we sort the terms according to their dependence on the amplitude
As or Ap :

9
9
1
9
9
1
i0
i0
3/2 As i +
iAs = ( )Ap ie +
iAp ei0
As +
Ap e +
2
2
2a
2 a
2
2a
2 a
Now we rewrite the result in terms of a complex number:

3/2 i + 29 a2 + 92 a i
Ap i0

e =
As
(1 + 12 )i + 92 a2 + 92 a i

(1 + 12 )i (1 + 21 )i
3/2 i + 29 a2 + 92 a i


+
=
(1 + 12 )i + 92 a2 + 92 a i (1 + 12 )i + 29 a2 + 92 a i

=1+
=1+

3/2i (1 + 21 )i
(1 + 12 )i + 92 a2 + 92
2(1 1)i

(21 + 1)i + 9 a2 + 9

Since we plot the results in terms of the dimensionless frequency S =


the result in terms of S and :
Ap i0
2( 1)i

e =1+
As
(2 + )i + 9 a2 + 9 a i
2( 1)iS

=1+
a2
(2 + )iS + + 9 a i 9

84

a2
9
a2
9

a2
,
9

we show

6.1 Solution in the inertial frame

2( 1)iS
Ap i0

e =1+
As
(2 + )iS + + 32 (1 + i) S
=1+
=1+

2( 1)iS
(2 + )iS + +

3 ei 4 2 S
2
2

2( 1)iS

(2 + )iS + + 3 Sei 4

This result coincides with the result by L0 Esperance [LCTR05].

6.1.1 Analysis of the result in the inertial frame

As mentioned, the result is given in terms of a complex number. We divide the


number into norm and phase. The norm gives us the amplitude ratio between the
sled and particle amplitude, whereas the phase gives us the phase difference between
them.
The norm of a complex fraction
z=

a + ib
ac + bd
bc ad
= 2
+i 2
2
c + id
c +d
c + d2

is given by:
r = |z| =

v
u
u
t

ac + bd
c2 + d2

!2

bc ad
+ 2
c + d2

!2

We rewrite the result in this form:


Ap i
2iS( 1)

e =1+

As
iS( + 2) + 3 S ei 4 +
Ap i
2iS( 1)
e =1+

As
iS( + 2) + 3 S 22 (1 + i) +

85

6 Appendix

iS( + 2) + 3 S 22 (1 + i) +
2iS( 1)
Ap i
e =
+



As
iS( + 2) + 3 S 22 (1 + i) + iS( + 2) + 3 S 22 (1 + i) +


Ap i 2iS( 1) + iS( + 2) + 3 S 22 + i3 S 22 +
e =


As
iS( + 2) + + 3 S 22 + i3 S 22

2
2 
Ap i 3 S 2 + + i 2S( 1) + S( + 2) + 3 S 2

e =


As
+ 3 S 22 + i S( + 2) + 3 S 22



3 S 22 + + i 3S + 3 S 22
Ap i

e =


As
+ 3 S 22 + i S( + 2) + 3 S 22

2
+
wherea = 3 S
2

2
b = 3S + 3 S
2

2
c = + 3 S
2

2
d = S( + 2) + 3 S
2

therewith one gets for the norm:


|z| =

|z| =

2
(3 S 2

v
u
u
t

ac + bd
c2 + d2

!2

bc ad
+ 2
c + d2

!2



2
+ )( + 3 S 22 ) + (3S + 3 S 22 )(S( + 2) + 3 S 22 )


( + 3 S 22 )2 + (S( + 2) + 3 S 22 )2

2
2
2
2 2 1/2
(3S + 3 S 2 )( + 3 S 2 ) (3 S 2 + )(S( + 2) + 3 S 2 )
+

( + 3 S 22 )2 + (S( + 2) + 3 S 22 )2

86

6.1 Solution in the inertial frame


The phase for a complex fraction is
tan =

=
=
<

bcad
c2 +d2
ac+bd
c2 +d2

bc ad
ac + bd





(3S + 3 S 22 )( + 3 S 22 ) (3 S 22 + )(S( + 2) + 3 S 22 )
=




(3 S 22 + )( + 3 S 22 ) + (3S + 3 S 22 )(S( + 2) + 3 S 22 )
(For the sake of simplicity we gave only the result for < > 0. We used the matlab
function atan2 in order to obtain the correct result.)

Figure 6.1 and 6.2 show the results for the amplitude ratio and the phase difference
in the inertial frame of reference.
2.5

Ampl. ratio (nondim)

=5
=2
= 0.5
= 0.2

1.5

0.5

0 6
10

10

10

10
S (nondim)

10

10

10

Figure 6.1: Amplitude ratios for different density ratios. Inertial frame of reference.

87

6 Appendix

=5
=2
= 0.5
= 0.2

/8

[rad]

/8

10

10

10

10
S (nondim)

10

10

10

Figure 6.2: Phase difference for different density ratios. Inertial frame of reference.

6.2 Solution for the co-moving frame

Again, we use the Ansatz:


u = As eit
u = As ieit
v = Ap eit+0
v = Ap ieit+i0

The equation is used according to Abbad and Souhar (see [AS04a]):


1
du
du
= (p f ) V v(t)
6f a u f V
Fh
dt
2
dt
!
p + 12 f du
p f
a
a
U

du
0
=
As 2 eit 6f U 6f
+
f
dt
f
V
V
2 dt
p V

with V = 4/3a3 ,

88

0 =

a2
,
f

2f
a2

= S 1/2

2
9

and

S=

a2
.
9

6.2 Solution for the co-moving frame


The insertion of these parameters leads to:
q

p +
f

2 f

du
p f
a
U
a
+
=
As 2 eit 6f

6
f

3
3
dt
f
4/3a
4/3a

p +
f

2 f

U
du
p f
9
9
=
As 2 eit 2 f U 2 f
+
dt
f
2a
2a

Now we define the parameter :=

p + 21 f
f

= 1 +

1
2

a2
2f

4
2

2f
a2

a4
f2

du

dt

du

dt

and get:

p = f 1/2f
p f
f 1/2f f
=
= 3/2
f
f
Insertion leads to:
q

9S

9
9
9S
du
du
= ( 3/2) As 2 eit 2 f U 2 f
U+ 2

dt
2a
2a
2
dt
In the next step we sort terms according to order u(t)
q

9S

9f 2 du 9
9
+
+

+
f
f
2a2
dt
2a2
2a2

9S
U = ( 3/2) As 2 eit
2

Now we use the ansatz to get:


q

9S

9f 2
9
9
2 2 it+

+
A
i

e
+

+
f
p
f
2a2
2a2
2a2

9S

9f 2
9
9
+
Ap i2 ei + 2 f + 2 f
2
2a
2a
2a

Ap i
e =
As
+

9f
2a2

9S 
2

9S
Ap ieit+ = ( 3/2) As 2 eit
2
s

9S
Ap iei = ( 3/2) As
2

( 3/2)
i2 +

2a2 f

2a2 f

9S
2

89

6 Appendix
The last step is to write the result in terms of S and :
Ap i
e =
As
Ap i
e =
As

a 9
( 3/2) 9
a2
9S

9
i2 + 2af2 2 i2 + i 2a92 f +

2a2 f

9S
2

( 3/2)S
i2 +
9
a2

9f
9 2a2
a2

9S 2
i
2

a 9
a 9
+ i 9
+ i 9

2a2 f
2a2 f
2

9S
2

Ap i
( 3/2)S
q
q
e =
1
9S
As
i2 S + i2 9S
+
i
+
i
2
2
2
Ap i
( 3/2)S
q

e = 
1
As
(1
+
i)
+
i iS + 9S
2
2
i( 3/2)S
Ap i
q

e =
i 4 2
As
iS + 9S
e
+
2
2

1
2

i(1 + 12 3/2)S
Ap i

e =

As
(1 + 12 )iS + 32 S ei 4 + 12
2iS(1 1)
Ap i

e =

As
iS(21 + 1) + 3 S ei 4 + 1
2iS(1 1)
Ap i


e =

As
iS(21 + 1) + 3 S ei 4 + 1
Ap i
2iS( 1)

e =

As
iS( + 2) + 3 S ei 4 +
with S =

a2
9

and =

f
p

6.2.1 Analysis of the result in the co-moving frame

We do the same steps as in the last section and get for the norm of the result:
r = |z| =

90

v
u
u
t

ac + bd
c2 + d 2

!2

bc ad
+ 2
c + d2

!2

6.2 Solution for the co-moving frame


Now we rewrite the result to get the parameters a, b, c, d:
Ap i
2iS( 1)

e =

As
iS( + 2) + 3 S ei 4 +
i2S( 1)
Ap i
e =


As
iS( + 2) + 3 S 22 (1 + i) +
a=0
b = 2S( 1)

2
c = + 3 S
2

2
d = S( + 2) + 3 S
2

therewith one gets for the norm:


|z| =

2

(2S( 1))(S( + 2) + 3 S 22 )

+
( + 3 S 22 )2 + (S( + 2) + 3 S 22 )2

2
)
2

(2S( 1))( + 3 S


( + 3 S 22 )2 + (S( + 2) + 3 S 22 )2

2 1/2

The phase for a complex fraction is


=
tan = =
<

bcad
c2 +d2
ac+bd
c2 +d2

bc ad
bc 0
=
ac + bd
0 + bd

+
3
S
c
2
= arctan( ) = arctan

d
S( + 2) + 3 S 22

(For the sake of simplicity we gave only the result for < > 0. We used the matlab
function atan2 in order to obtain the correct result.)
Figure 6.3 and 6.4 show the results for the amplitude ratio and the phase difference
in the comoving frame .

91

6 Appendix

1.4

Ampl. ratio (nondim)

1.2
1

=5
=2
= 0.5
= 0.2

0.8
0.6
0.4
0.2
0 6
10

10

10

10
S (nondim)

10

10

10

Figure 6.3: Amplitude ratios for different density ratios. Comoving frame.

[rad]

/2

=5
=2
= 0.5
= 0.2

/4

0 6
10

10

10

10
S (nondim)

10

10

10

Figure 6.4: Phase difference for different density ratios. Comoving frame.

92

6.3 Inverse Laplace Transform of the Basset equation

6.3 Inverse Laplace Transform of the Basset equation


This section shows, how the inverse Laplace transform of the Basset equation can
be performed. The solution for the inverse Laplace transform was already given by
Gorenflo and Mainardi [GM08]. We start with the definition of the integral we want
to solve.
(

u0 (t) = L

1 + s1/21
s + s1/2 + 1

and

u (t) = L1

1
s + s1/2 + 1

We solve the integral by the application of complex integration. We show the steps
for the first integral u (t) explicitly, the solution for the second integral can be
derived analogous.
First, we explicitly write down the Laplace integral we want to solve:
1
)
s + s1/2 + 1
1
1 Z +i st
e
ds
=
2i i
s + s1/2 + 1
1 Z
1
=
est
ds
2i H() s + s1/2 + 1

u (t) = L1 (

+i
. . . ds by the Hankel integral
Thereby we have replaced the Bromwich integral i
R
H() . . . ds. This was possible, because the integral kernel has no zeros and we can
deform the Bromwich path (line at from + on imaginary axis, with
greater than real part of zeros) to a Hankel path (loop around origin with radius
 0). The sketch defines our integration path: The result will be given by the
sum of the three parts defined by:

1
u (t) =
2i

Z
A

Z
AB

Z
B

The first integration is done along the line s = xei with x going from  (as
s goes from ). By defining the function F (s) = s+s11/2 +1 the resulting
integral is:
Z
A

est F (s)ds =

Z 

est F (s)ds =

Z 

Z 

ee

i xt

F (xei )ei dx

ext F (xei )dx

93

6 Appendix

Im
Bromwich path
deformation
of integral
path

Hankel path

Re

Figure 6.5: Integration path: the direction is marked by the small arrows

The next integral is evaluated at the circle s = ei with  0. We get:


Z
AB

est F (s)ds =

= i

ee F (ei )iei d


i

ee F (ei )ei d

Then we take the limit  0:


lim i
0

ee F (ei )ei d = lim i[ei ]

0

=0

94

6.3 Inverse Laplace Transform of the Basset equation


The solution for the integral
Then we get:
u (t) =
=
=
=

R
B

is analogue to the solution for the first integral.

Z
Z
1
+
+
2i A
AB
B
 Z 

Z 
1

ext F (xei )dx + 0


ext F (xei )dx
2i

1 Z  xt
e F (xei )dx
i Z


1  xt
1
i
dx
e Im
|
s=xe

s + s1/2 + 1
Z

In the last step the real part is zero, because we integrate first from to  and
then back, so both contributions cancel. Now we evaluate the Imaginary part I:
1
i
I = Im
|
1/2 + 1 s=xe
s
+

= Im
1
xei + x 2 [cos( 21 ) + i sin( 12 )] + 1


Under consideration of the rule for the imaginary part of a fraction Im(z) = Im( a+bi
)=
c+d
bcad
we get
c2 +d2
1

I=

x 2 sin( 21 )
1

(1 x)2 + 2 x + 2(1 x)x 2 cos( 12 )

Finally, the limit  0 is taken and the result is obtained: The following result is
obtained
1 Z xt
e Idx
u (t) =
0
1
x 2 sin( 12 )
1 Z xt
=
e
dx.
1
0
(1 x)2 + 2 x + 2(1 x)x 2 cos( 12 )

(6.1)
(6.2)

(A discussion of this result can be found in the theory chapter.)

95

Bibliography
[AHP88] T.R. Auton, J.C.R. Hunt, and M. PrudHomme. The force exerted
on a body in inviscid unsteady nonuniform rotational flow. Journal
of Fluid Mechanics, 197(1):241257, 1988.
[All00] H.S. Allen. The motion of a sphere in a viscous fluid. Philosophical
Magazine Series 5, 50(306):519534, 1900.
[Arn11] H.D. Arnold. Limitations imposed by slip and inertia terms upon
stokess law for the motion of spheres through liquids. Physical
Review Series I, 32:233233, 1911.
[AS04a] M. Abbad and M. Souhar. Effects of the history force on an oscillating rigid sphere at low reynolds number. Experiments in fluids,
36(5):775782, 2004.
[AS04b] M. Abbad and M. Souhar. Experimental investigation on the history
force acting on oscillating fluid spheres at low reynolds number.
Physics of Fluids, 16:3808, 2004.
[ASC06] M. Abbad, M. Souhar, and O. Caballina. Note on the memory force
on a slightly eccentric fluid spheroid in unsteady creeping flows.
Physics of Fluids, 18:013301, 2006.
[Bas88] A.B. Basset. On the motion of a sphere in a viscous liquid. Philosophical Transactions of the Royal Society of London. A, 179:4363,
1888.
[BE11] E. Bodenschatz and M. Eckert. Prandtl and the gottingen school.
A Voyage Through Turbulence, page 40, 2011.
[Bou85] V.J. Boussinesq. Sur la rsistance quoppose un liquide indfini en
repos. In Comptes rendus de lAcadmie des sciences, volume 100,
page 9357, 1885.

97

Bibliography
[Bre05] C.E. Brennen. Fundamentals of multiphase flow. Cambridge Univ
Press, 2005.
[Bue66] Y.A. Buevich. Motion resistance of a particle suspended in a turbulent medium. Fluid Dynamics, 1(6):119119, 1966.
[Cau23] B.A.L. Cauchy. Recherches sur lquilibre et le mouvement intrieur
des corps solides ou fluides, lastiques ou non lastiques. 1823.
[CK02] C.F.M. Coimbra and M.H. Kobayashi. On the viscous motion of a
small particle in a rotating cylinder. Journal of Fluid Mechanics,
469:257286, 2002.
[CL56] S. Corrsin and J. Lumley. On the equation of motion for a particle
in turbulent fluid. Applied Scientific Research, 6(2):114116, 1956.
[CLL+ 04] CFM Coimbra, D. LEsperance, RA Lambert, JD Trolinger, and RH
Rangel. An experimental study on stationary history effects in highfrequency stokes flows. Journal of Fluid Mechanics, 504:353363,
2004.
[CR98] C.F.M. Coimbra and R.H. Rangel. General solution of the particle momentum equation in unsteady stokes flows. Journal of Fluid
Mechanics, 370(1):5372, 1998.
[Das11] S. Das. Functional fractional calculus. Springer Verlag, 2011.
[Deb34] P. Debye. Part i. dielectric constant. energy absorption in dielectrics
with polar molecules. Trans. Faraday Soc., 30:679684, 1934.
[Die82] W.E. Dietrich. Settling velocity of natural particles. Water Resources Research, 18(6):16151626, 1982.
[DN74] G. Doetsch and W. Nader. Introduction to the Theory and Application of the Laplace Transformation. Springer Berlin, 1974.
[dSV43] B. de Saint-Venant. Note a joindre au memoire sur la dynamique
des fluides. Comp. Rend, 17:1240, 1843.
[ERFM12] P. Ern, F. Risso, D. Fabre, and J. Magnaudet. Wake-induced oscillatory paths of freely rising or falling bodies. Annual Review of
Fluid Mechanics, 44(1), 2012.

98

Bibliography
[Gat83] R. Gatignol. The faxen formulae for a rigid particle in an unsteady
non-uniform stokes flow. J. Mc. Thor. Appl, 1(2):143160, 1983.
[GDO+ 09] V. Garbin, B. Dollet, M. Overvelde, D. Cojoc, E. Di Fabrizio, L. van
Wijngaarden, A. Prosperetti, N. de Jong, D. Lohse, and M. Versluis. History force on coated microbubbles propelled by ultrasound.
Physics of Fluids, 21:092003, 2009.
[GHG00] I Gonzlez, T.L Hoffmann, and J.A. Gallego. Precise measurement
of particle entrainment in a standing-wave acoustic field between 20
and 3500 hz. Journal of Aerosol Science, 31(12):14611468, December 2000.
[GM08] R. Gorenflo and F. Mainardi.
Fractional calculus: integral
and differential equations of fractional order. Arxiv preprint
arXiv:0805.3823, 2008.
[He00] R. Hilfer (ed.). Applications of fractional calculus in physics, volume
463. World Scientific Singapore, 2000.
[Hil02] R. Hilfer. Experimental evidence for fractional time evolution in
glass forming materials. Chemical physics, 284(1-2):399408, 2002.
[HK94] T. L. Hoffmann and G. H. Koopmann. A new technique for visualization of acoustic particle agglomeration. Review of scientific
instruments, 65(5):15271536, 1994.
[JBD03] M. Jenny, G. Bouchet, and J. Duek. Nonvertical ascension or fall
of a free sphere in a newtonian fluid. Physics of Fluids, 15:L9, 2003.
[JDB+ 04] M. Jenny, J. Dusek, G. Bouchet, et al. Instabilities and transition
of a sphere falling or ascending freely in a newtonian fluid. Journal
of Fluid Mechanics, 508(1):201239, 2004.
[JHI+ 07] G. Jourdan, L. Houas, O. Igra, J. L Estivalezes, C. Devals, and E. E.
Meshkov. Drag coefficient of a sphere in a non-stationary flow: new
results. Proceedings of the Royal Society A: Mathematical, Physical
and Engineering Science, 463(2088):33233345, 2007.

99

Bibliography
[KES98] I. Kim, S. Elghobashi, and W.A. Sirignano. On the equation for
spherical-particle motion: effect of reynolds and acceleration numbers. Journal of Fluid Mechanics, 367:221253, 1998.
[KK78] S.K. Karanfilian and T.J. Kotas. Drag on a sphere in unsteady
motion in a liquid at rest. Journal of Fluid Mechanics, 87:8596,
1978.
[LB93a] P.M. Lovalenti and J.F. Brady. The force on a sphere in a uniform flow with small-amplitude oscillations at finite reynolds number. Journal of Fluid Mechanics, 256:607607, 1993.
[LB93b] P.M. Lovalenti and J.F. Brady. The hydrodynamic force on a
rigid particle undergoing arbitrary time-dependent motion at small
reynolds number. Journal of Fluid Mechanics, 256(1):561605, 1993.
[LCTR05] D. LEsprance, C.F.M. Coimbra, J.D. Trolinger, and R.H. Rangel.
Experimental verification of fractional history effects on the viscous dynamics of small spherical particles. Experiments in fluids,
38(1):112116, 2005.
[LD09] E. Loth and A.J. Dorgan. An equation of motion for particles of
finite reynolds number and size. Environmental fluid mechanics,
9(2):187206, 2009.
[Lie27] H. Liebster. ber den widerstand von kugeln. Annalen der Physik,
387(4):541562, 1927.
[Lin99] E. R. Lindgren. The motion of a sphere in an incompressible viscous fluid at reynolds numbers considerably less than one. Physica
Scripta, 60(2):97110, August 1999.
[LS24] H. Liebster and L. Schiller. Kinematographische messungen der
fallbewegung von kugeln in zaher flussigkeit, auch in nachster nahe
einer wand. Physikalische Zeitschrift, 25:670, 1924.
[LW86] C.J. Lawrence and S. Weinbaum. The force on an axisymmetric
body in linearized, time-dependent motion: a new memory term.
Journal of Fluid Mechanics, 171(1):209218, 1986.

100

Bibliography
[LW88] C.J. Lawrence and S. Weinbaum. The unsteady force on a body
at low reynolds number; the axisymmetric motion of a spheroid.
Journal of Fluid Mechanics, 189(1):463489, 1988.
[MA92] R. Mei and R.J. Adrian. Flow past a sphere with an oscillation in the
free-stream velocity and unsteady drag at finite reynolds number.
Journal of Fluid Mechanics, 237(1):323341, 1992.
[Mei94] R. Mei. Flow due to an oscillating sphere and an expression for
unsteady drag on the sphere at finite reynolds number. Journal of
Fluid Mechanics, 270(1):133174, 1994.
[MK92] R. Mei and J.F. Klausner. Unsteady force on a spherical bubble
at finite reynolds number with small fluctuations in the free-stream
velocity. Physics of Fluids A: Fluid Dynamics, 4:63, 1992.
[ML98] J. Magnaudet and D. Legendre. The viscous drag force on a spherical
bubble with a time-dependent radius. Physics of fluids, 10:550554,
1998.
[MLA91] R. Mei, C. Lawrence, and R. Adrian. Unsteady drag on a sphere
at finite reynolds number with small fluctuations in the free-stream
velocity. Journal of Fluid Mechanics, 233:613631, 1991.
[MP00] N. Mordant and J.F. Pinton. Velocity measurement of a settling
sphere. The European Physical Journal B-Condensed Matter and
Complex Systems, 18(2):343352, 2000.
[MR83] M.R. Maxey and J.J. Riley. Equation of motion for a small rigid
sphere in a nonuniform flow. Physics of Fluids, 26:883, 1983.
[MR10] E.E. Michaelides and A. Roig. A reinterpretation of the odar and
hamilton data on the unsteady equation of motion of particles.
AIChE Journal, 2010.
[NA93] R. Natarajan and A. Acrivos. The instability of the steady flow
past spheres and disks. Journal of Fluid Mechanics, 254(1):323344,
1993.

101

Bibliography
[Nav23] C. Navier. Mmoire sur les lois du mouvement des fluides. Mmoires
de lAcadmie Royale des Sciences de lInstitut de France, 6:389440,
1823.
[OH64] F. Odar and W.S. Hamilton. Forces on a sphere accelerating in a
viscous fluid. Cambridge University Press, 1964.
[Ose10] C.W. Oseen. ber die stokessche formel und ber eine verwandte
aufgabe in der hydrodynamik. Arkiv fr matematik, astronomi och
fysik, 6:29, 1910.
[PKM95] W.C. Park, J.F. Klausner, and R. Mei. Unsteady forces on spherical
bubbles. Experiments in fluids, 19(3):167172, 1995.
[Poi28] S. D. Poisson. Mmoire sur lquilibre et mouvement des corps lastiques. LAcadmie des sciences, 1828.
[Sar86] T. Sarpkaya. Force on a circular cylinder in viscous oscillatory flow
at low KeuleganCarpenter numbers. Journal of Fluid Mechanics,
165:6171, 1986.
[SBB+ 09] G. Schlffel, M. Bastide, S. Bachmann, C. Mundt, and F. Seiler. Investigation of the acceleration of aluminum particles behind a shock
wave using instantaneous laser doppler velocimetry. Shock Waves,
19(2):125134, 2009.
[Sch28] J. Schmiedel. Experimentelle Untersuchungen ber die Fallbewegung
von Kugeln und Scheiben in reibenden Flssigkeiten. Universitt
Leipzig, 1928.
[SG00] H. Schlichting and K. Gersten. Boundary-layer theory. Springer
Verlag, 2000.
[SOC07] Y.D. Sobral, T.F. Oliveira, and F.R. Cunha. On the unsteady forces
during the motion of a sedimenting particle. Powder Technology,
178(2):129141, 2007.
[Sto46] G.G. Stokes. Report on recent researches in hydrodynamics. Brit.
Ass. Rep, 1:120, 1846.

102

Bibliography
[Sto51] G.G. Stokes. On the effect of the internal friction of fluids on the
motion of pendulums. Transactions of the Cambridge Philosophical
Society, 1851.
[Tan56] S. Taneda. Experimental investigation of the wake behind a sphere
at low reynolds numbers. Journal of the Physical Society of Japan,
11:11041108, 1956.
[Tch47] C.M. Tchen. Mean value and correlation problems connected with
the motion of small particles suspended in a turbulent fluid. PhD
thesis, TU Delft, 1947.
[TCNDVdA02] A. Ten Cate, C.H. Nieuwstad, J.J. Derksen, and H.E.A. Van den
Akker. Particle imaging velocimetry experiments and latticeboltzmann simulations on a single sphere settling under gravity.
Physics of Fluids, 14:4012, 2002.
[TK80] S. Temkin and S. S Kim. Droplet motion induced by weak shock
waves. Journal of Fluid Mechanics, 96:133157, 1980.
[TKT91] Y. Tsuji, N. Kato, and T. Tanaka. Experiments on the unsteady
drag and wake of a sphere at high reynolds numbers. International
journal of multiphase flow, 17(3):343354, 1991.
[TM82] S. Temkin and H. K. Mehta. Droplet drag in an accelerating and
decelerating flow. Journal of Fluid Mechanics, 116(1):297313, 1982.
[Tri88] D.J. Tritton. Physical fluid dynamics. Oxford, Clarendon Press,
1988.
[VBVWL05] C. Veldhuis, A. Biesheuvel, L. Van Wijngaarden, and D. Lohse. Motion and wake structure of spherical particles. Nonlinearity, 18:C1,
2005.
[VIG07] John Veysey II and N. Goldenfeld. Simple viscous flows: From
boundary layers to the renormalization group. Reviews of Modern
Physics, 79(3):883, 2007.
[VPS10] I.S. Vodopyanov, A.G. Petrov, and M.M. Shunderyuk. Unsteady
sedimentation of a spherical solid particle in a viscous fluid. Journal
of Fluid Dynamics, 45(2):254263, 2010.

103

Bibliography
[WB07] L. Wakaba and S. Balachandar. On the added mass force at finite
reynolds and acceleration numbers. Theoretical and Computational
Fluid Dynamics, 21(2):147153, 2007.
[Wie22] C. Wieselsberger.
Weitere feststellungen ber die gesetze
des flssigkeits-und luftwiderstandes. Physikalische Zeitschrift,
23:219224, 1922.
[Wil15] W.E. Williams. On the motion of a sphere in a viscous fluid. Philosophical Magazine, 29:526, 1915.
[WKB08] J.A. Weinstein, D.R. Kassoy, and M.J. Bell. Experimental study
of oscillatory motion of particles and bubbles with applications to
coriolis flow meters. Physics of Fluids, 20:103306, 2008.

104

Thanks
First, I would like to thank Prof. Dr. Eberhard Bodenschatz, my advisor, for the
opportunity to conduct this research in his group. He gave me the idea for the
experiment and inspired me with trust to set up the experiment. Also I would like
to thank Prof. Dr. Andreas Tilgner for being the co-referee.
I am also very grateful to Dr. Haitao Xu for his assistance throughout the whole
thesis project. His experience was a great support at all times. The discussions
with him about the experimental setup, the conduction of the experiments and the
results were always fruitful. Especially, I am grateful to him for writing independent
analysis codes, that were much faster and preciser than my codes. These programs
simplified the data analysis substantially. I also would like to thank Walter Pauls,
Mathieu Gibert and Ewe Wei Saw for discussions on the setup and the results.
The mechanical setup of the motor was done by Andreas Gehrke, Nils Winkelmann
and Daniel Wollborn under the supervision of Udo Schminke. I am very grateful
to them for their work and their patience with my numerous design changes that
caused lots of work for them. The electrical setup of the motor was done by Andreas
Renner and Ortwin Kurre, who I would like to to thank as well. Without them the
different signals might still interfere and stop the motor. The initial operation of
the motor and the improvement of the motor performance took a large part of the
thesis period. I am thankful to the Bosch Rexroth service team, especially Michael
Eggert, Wolfgang Walter and Mathias Wahler for their help. I am grateful to the
Heidenhain technical staff and Rainer Schtig from Rheinwerkzeug for their support.
Another company I would like to thank is HEGLA, which gave us the opportunity
to check the performance of their system. The comparison lead to new ideas and
was very useful for the project.
The setup of the experiment itself involved also almost everyone from our group.
Special thanks got to Joachim Hesse, Andre Heil, Andreas Kopp, Dr. Artur Kubitzek and Wolfgang Schubert. Last but not least, I want to thank my brother
Christian and Haitao Xu for correcting the various drafts of the thesis.

105

Eidesstattliche Erklrung:
Hiermit erklre ich, dass ich diese Abschlussarbeit selbstndig verfasst habe, keine
anderen als die angegebenen Quellen und Hilfsmittel benutzt habe und alle Stellen,
die wrtlich oder sinngem aus verffentlichten Schriften entnommen wurden, als
solche kenntlich gemacht habe. Darberhinaus erklre ich, dass diese Abschlussarbeit nicht, auch nicht auszugsweise, im Rahmen einer nichtbestandenen Prfung an
dieser oder einer anderen Hochschule eingereicht wurde.
Gttingen, January 18, 2013

(Sebastian Lambertz)

Você também pode gostar