Impulse Response

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

Experiments in Fluids 28 (2000) 152159  Springer-Verlag 2000

The use of impulse response functions for evaluation of added mass and
damping coefficient of a circular cylinder oscillating in linearly stratified fluid
E. V. Ermanyuk

152
Abstract The damped horizontal oscillations of a circular
cylinder in linearly stratified fluid are studied experimentally.
The cylinder is fixed to the lower end of a physical pendulum
with variable restoring moment. The impulse response function of the pendulum in time domain is recorded and
converted to the frequency response function using Fourier
transform. The density stratification is shown to have a
strong effect on frequency-dependent hydrodynamic
coefficients (added mass and damping). The data obtained
are compared with available theoretical predictions. The
applicability of a simplicistic method implying approximation
of impulse response functions by analytical functions is
discussed.

List of symbols
t
u
o
g
Do
H
D
N
r(t)
" R(u)"
h(u)

time
frequency
fluid density
gravity acceleration
density variation over depth
depth of fluid
diameter of the cylinder
BruntVaisala frequency
impulse response function
amplitude of frequency response function
phase of frequency response function

Received: 15 September 1998/Accepted: 1 April 1999


E. V. Ermanyuk
Lavrentyev Institute of Hydrodynamics
Siberian Division of Russian Academy of Science
Av. Lavrentyev 15, Novosibirsk, 630090, Russia
E-mail: ermanyuk@hydro.nsc.ru
The present study has been supported by the grant for young scientists
of Siberian Division of Russian Academy of Science. The author
express thanks to Prof. A. A. Korobkin and to Prof. I. V. Sturova for
valuable discussions and comments. Thanks are due to Dr. V. A.
Kostomakha and Dr. N. V. Gavrilov for their help in experiments and
data processing. Further, the author express thanks to the reviewers
for helpful and contructive comments. The manuscript of this paper
was prepared during the authors stay as a visiting researcher at the
Research Institute for Applied Mechanics, Kyushu University. The
friendly help and guidance of Prof. M. Ohkusu and Prof. M. Kashiwagi
is gratefully acknowledged.

k
j

added mass
damping coefficient

1
Introduction
Internal waves generated by an oscillating body in a density
stratified fluid have been a topic of much interest during past
three decades (see, Mowbray and Rarity 1967; Hurley 1969;
Appleby and Crighton 1986; Ivanov 1989; Makarov et al. 1990;
Hurley 1997; Hurley and Keady 1997). The cases of simple
body geometry (circular and elliptical cylinders) and exponential (at small vertical scales linear) density distribution over
depth are the most commonly studied ones. As a result of these
studies, the details of the spatial structure of internal waves
emitted by harmonic oscillators are quite well understood.
However, the integral quantities (such as added mass and
damping coefficient) characterizing the fluid-body system as
a mechanical oscillator have been considered only by Hurley
(1997) (inviscid case) and Hurley and Keady (1997) (approximate viscous solution). When the fluid is assumed to be
inviscid, the energy dissipation is associated solely with the
radiation of internal waves by an oscillating body. Hurley
(1997) reveals some curious features of fluid-body interaction
in this case.
He found that for the frequency range below the
BruntVaisala frequency, added mass is zero and damping is
non-zero, and, conversely, for frequencies higher than the
BruntVaisala frequency, added mass is non-zero and damping is zero. Moreover, the dependencies of dynamic coefficients
on frequency are of universal character for any direction of
body oscillation. The study presented in Hurley (1997) has
been generalized by Hurley and Keady (1997) to take viscous
effects into account. They argue that for sufficiently large
Reynolds numbers (which are normally encountered in
experiments), the inviscid solutions do hold approximately for
the viscous case. However, when the frequency of oscillations is
lower (higher) than the BruntVaisala frequency, the value of
added mass (damping) in viscous fluid is expected to be
non-zero. Note that no experimental confirmation of the
above-mentioned results has been reported in literature so far.
The goal of the present study is to evaluate experimentally
added mass and damping coefficient for a circular cylinder
oscillating horizontally in a linearly stratified fluid. To do this,
we use impulse response function. This basic tool is well
known in different fields of physics. It is known that for
any linear system its response to a unit impulse in the time
domain is related by Fourier transform with the response to

harmonical forcing in frequency domain. This idea serves as


a background of the present study.
The fundamentals concerning this approach as applied to
the theory of transient ship motions in surface waves have been
thoroughly considered by Cummins (1962) (for an overview,
see Wehausen 1971). A distinctive physical feature of this
problem is the memory effect which is evident from the
simple observation that wave disturbance generated by the
motion of a body at any time instance persists in future and
therefore acts on the further character of the body motion.
Mathematically, the memory effect is taken into account by
the convolution integral introduced into the description of the
body motion in time domain. The importance of the memory effect has been emphasized in early theoretical study by
Sretenskii (1937). Sretenskii studied damped oscillations of
a body floating on a free surface. The calculated decay of free
oscillations was found to have much more complicated
character than that of a simple damped linear oscillator.
Physically, the wave motion in a stratified fluid has much in
common with surface waves. For a particular type of stratification (two-fluid system with interface between layers) the
approach used by Sretenskii is further developed by Akulenko
and Nesterov (1987) and Akulenko et al. (1988). The analysis is
restricted to the case of vertical oscillations of thin 2D bodies.
The results of these theoretical studies are partially supported
by quantitative measurements of damped oscillations of a body
at the interface of miscible fluids reported by Pylnev and
Razumeenko (1991). The analysis of experimental data in their
paper was based on the approximation of experimental curves
by simple analytical functions.
Larsen (1969) presents a theoretical and experimental study
of damped oscillations of a sphere initially displaced from the
equilibrium position in a linearly stratified fluid. Theory and
experiments are found to be in good agreement. In particular,
Larsen argues that the observed damping can be explained
entirely by the radiation of energy by internal waves, while the
viscous effects are practically negligible. However, no quantitative measurement of frequency dependent dynamic coefficients was presented in his study.
It should be noted that there exists an extensive literature
on the experimental evaluation of inertial and damping force
coefficients in the case of harmonical oscillations of a circular
cylinder in homogeneous fluid. These coefficients are
shown to be dependent on two non-dimensional parameters:
KeuleganCarpenter number and Stokes number (see, Sarpkaya 1986). In the present study, we are concerned with the
influence of stratification which is governed by another basic
parameter, namely, the ratio between the frequency of
oscillations of a cylinder and the BruntVaisala frequency.
This ratio has the physical sense of the Froude number.

2
Experiments
2.1
Experimental installation
The experiments were carried out in a test tank [0.14 m wide,
0.32 m deep and 1 m long]. A scheme of the experimental
installation is shown in Fig. 1. The test tank 1 was filled with
linearly stratified fluid to the depth H:0.28 m. The total

153

Fig. 1. Scheme of the experimental installation

variation of density over depth was Do:0.022 g/cm3. The


corresponding BruntVaisala frequency was N:(9gdo/dz
:0.88 rad/s. The linearity of the density distribution was
checked by the conductivity probe 2.
A weak solution of glycerine in water was used to create the
density stratification. Let us make a brief note on the physical
properties of such a solution. Detailed reference tables showing
the dependence of the glycerinewater solutions dynamic
viscosity on concentration and temperature may be found in
Vargaftik (1963). It is known that the dynamic viscosity of
a weak glycerinewater solution varies almost linearly with
concentration. In the present experiments, the volume concentration of the glycerine in the solution gradually increased
from zero at the free surface up to 8.8% close to the bottom of
the test tank. Correspondingly, the dynamic viscosity varied
from 1.14;10\3 kg/(m s) up to 1.45;10\3 kg/(m s). Here, the
first value corresponds to the dynamic viscosity of pure water
at the temperature T:15C which was fairly constant (to
within <1C) throughout the experiments. It is interesting to
note that for the weak glycerinewater solution the relative
drop of the dynamic viscosity Dg/g due to a slight increase of
temperature DT in the vicinity of T+15C appears to be
independent from concentration, being equal to 2.4% per 1C.
The same value of Dg/g takes place in pure water.
The diameter D of the circular cylinder 3, which was fixed at
the lower streamlined end of the physical pendulum 4, was
equal to 3.7 cm. The centre of the cylinder was submerged to
the depth 0.5H. The volume of the immersed streamlined part
of the pendulum was less than 1% of the cylinder volume so
that its influence on the fluid-body interaction may be safely
neglected. The pendulum had two wedge-shaped supports of
steel, 5. Each wedge contacted a horizontal cylinder made of
steel and oriented normally to the rib of the wedge. Two points
of contact defined the axis of pendulum rotation. This design
allowed to reduce the friction at the supports to a minimum.
The pendulum had a micrometric screw 6 with a nut 7. The
variation of vertical distance z between the gravity center of
0
the nut and the axis of pendulum rotation allowed the restoring
moment of the pendulum to be changed. As result, the
frequency of damped oscillations could be varied. The horizontal beam of the pendulum was equipped with a pre-tensioned
rubber membrane 8. The impulse of force needed to excite free
oscillations was introduced by dropping a steel ball which was

154

initially held by an electric magnet 9. As the characteristic


period of free oscillations of the pendulum in experiments was
of the order of several seconds, the impact of a ball on the
membrane produced practically ideal impulse loading. To
ensure a single impulse, the initial position of a ball was chosen
so that it hit the membrane slightly off-center and then jumped
away.
The motion of the pendulum was measured by a displacement sensor 10. The sensor was made of a resistive wave gauge.
A wave gauge having two vertical wires was placed in a small
vessel containing a conductive liquid. A light shield made of
non-conductive material and connected to the horizontal beam
of the pendulum by a thin needle was free to move between the
wires. Static and dynamic performance of the sensor was tested
and proved to be both linear and practically instant. The drag
at the shield was evaluated from the records of oscillations of
the pendulum in air and found to be negligibly small. The
advantage of the described design is the absence of mechanical
Coulombs friction. Additionally, it eliminates the need for any
electric feeders connected to the moving parts of the installation. These precautions were taken as the present experiments
imply the measurement of motions governed by very small
forces. Let us also note that the diameter of the needle
connecting the shield to the pendulum was minimized in order
to avoid the effects due to surface tension. At this point it is
pertinent to add some comments concerning the effects of the
lubrication flow in the gaps between the ends of the cylinder
and the walls of the test tank. The area of the cylinder
cross-section was approximately equal to the area of the shield
used in the displacement sensor. The thickness of the oscillatory boundary layer at the shield is of order d:(l/u (see, for
example, Hurley and Keady (1997)). In the studied frequency
range, this value is comparable with the width of the gaps
between the ends of the cylinder and the walls of the test tank
which were equal to 0.5 mm. Consequently, we can assume that
the drag due to the lubrication flow in the gaps is comparable
with the drag at the shield. Therefore, it can be safely neglected.
The ends of the test tank were equipped with the wave
breakers 11. Each wave breaker consisted of two perforated flat
plates spaced at 2 cm from each other. The plate placed at the
weather side was perforated with bigger orifices of about 3 cm
in diameter while the second plate had smaller orifices of about
1.5 cm in diameter. The performance of the wave breaker was
found to be very effective.
The geometrical and dynamical parameters of the pendulum
were chosen at the design stage so as to ensure a sufficiently
wide range of the oscillation frequencies. The actual values
were carefully measured. The moment of inertia of the
pendulum about the axis of rotation is J :1.12;106 g cm2 with
0
an accuracy of 0.5%. The total moment of inertia (including
nut 7 of mass m:188 g) is J:J ;mz2 . The mass of the
0
0
pendulum (without nut) is M:706 g.
To measure the pendulums restoring moment coefficient,
a static calibration in situ was performed. To do this, the
pendulum was loaded with standard calibrated weights placed
in a light bowl 12 suspended at distance 60 cm from the axis of
rotation of the pendulum. The angle of static inclination was
measured by the displacement sensor. A typical load used for
calibration varied between 0.5 g for small z (about 3 cm)
0
and 0.01 g for large z (the pendulum was close to neutral
0

equilibrium at z +17 cm). At the second stage of calibration,


0
a known angle deflection was applied to the pendulum by
means of a micrometric mechanism 13. The mechanism
allowed to produce a known vertical displacement with an
accuracy of 0.01 mm. The displacement was applied at
a distance of 60 cm from the axis of rotation of the pendulum.
The measurement of the restoring moment coefficient was
performed for all values of z used in the experiments. The
0
accuracy of this procedure was about 0.5%. The studied range
of the restoring moment coefficient is between 9;10\3 and
0.3 Nm/rad.
It should be noted that the angular inclination of the
pendulum did not exceed 0.5 in the experiments or 0.7
during static calibration. The distance between the axis of
rotation of the pendulum and the centre of the cylinder was
b:60 cm. Thus, the linear horizontal displacement of the
cylinder center in experimental runs was less than 0.14 D.
To complete the description of the experimental installation,
it should be noted that it was located in a place protected from
air currents and vibrations in order to minimize any disturbance which could affect such a sensitive system.

2.2
Mathematical model
The mathematical model of the problem can be readily
formulated following Cummins (1962).
For any stable linear system, if r(t), the response to a unit
impulse, is known, the response of the system to an arbitrary
force f (t) is

x(t):  r(q) f (t9q) dq
(1)
0
Let us assume that the dynamic system is exposed to the action
of a harmonic force
F(iu):f e St
0
Substituting Eq. (2) in Eq. (1) one obtains

(2)

x(t):f e StR(iu)
(3)
0
where the complex frequency response function R(iu) is
defined as Fourier transform of the impulse response function

R(iu):  r(q)e\ SO dq
0
The complex frequency response function can be separated
into real and imaginary parts as R (iu):R (iu)9iR (iu),
c
s
where

R :  r(q) cos uq dq
c
0

R :  r(q) sin uq dq
s
0
Furthermore, one can introduce the amplitude "R":
([R (u)]2;[R (u)]2 and the phase h:arctan(R /R ) of the
s c
c
s
frequency response function. Let us note that the last representation seems to be more convenient for an experimentalist
since the physical sense of the data obtained can be checked. In
particular, when the energy is radiated from the system, the
phase angle must fall in the range between 0 and n.

Assuming that the angle u of pendulum inclination is small


(u:bx, where x is horizontal displacement of the center of the
cylinder), one can write the equation of the horizontal motion
of the cylinder center in the frequency domain as follows:
(J/b2;k(u))x( ;j(u)xR ;cx:f e St
(4)
0
Here, k(u) is the added mass of the cylinder, j(u) is the
damping coefficient, c is restoring force coefficient, overdot
denote the differentiation with respect to time. Let us note that
c is a linear function of z . This property was used to check the
0
static calibration data. Combining Eqs. (3) and (4), and using
the linearity of the system, one can write the expressions for
added mass and damping coefficient

" R(0)"
J
c
19
cos(h(u)) 9
k(u):
"R(u)"
b2
u2

(5)

c " R(0)"
j(u):
sin(h(u))
u "R(u)"

(6)

Here "R(0)" denotes the amplitude of the frequency response


function at zero frequency. Once the system is proved to be
linear, the normalization of "R(u)" by "R(0)" allows us to use
Fourier transforms of the experimental impulse response
functions at any value of the impulse loading. Thus, there is no
need to perform any direct measurement of the actual value of
the impulse.
Theoretically, the above-mentioned procedure allows us to
evaluate the dependencies k(u) and j(u) for 0:u:R from
one record of damped oscillations. In practice, the reliable
estimates may be obtained within a finite frequency range in
the vicinity of the resonant frequency u , where u corres*
*
ponds to the maximum of "R(u)". In the present problem, the
evaluation of this frequency range is of special interest as one
can expect a quite different physical behavior of the hydrodynamic coefficients at the frequencies when damping is
conditioned by wave or viscous phenomena.
Along with the method described above, we also apply
a simple approach to the evaluation of hydrodynamic coefficients using the least squares approximation of the experimental impulse response function by an exponentially
decaying sine function. Mathematically, the method is based
on the minimization of the integral

 (r(t)9r (t, A, k, X))2 dt,
(7)
a
0
where r (t, A, k, X):Ae\kt sin(Xt); the values A, k, X are
a
unknown variables. Once there variables are determined, then,
using a rough assumption that Eq. (4) does hold true for
transient responses, we can determine the added mass and
damping coefficient as follows:
J
c
9
k(X):
X2;k2 b2

(8)

2kc
j(X):
X2;k2

(9)

The choice of the approximating function in Eq. (7) is,


strictly speaking, arbitrary. There is no rigorous theoretical
proof that an exponentially decaying sine function may serve
as a meaningful approximation for an actual response function

of a complicated dynamical system with memory. Moreover,


it is rather well known that the law of damped oscillations
of a wave-emitting body has different asymptotics at small
and large time-scales (see, Huskind 1947; Wehausen 1971;
Akulenko et al. 1988). For this reason, it is not evident that the
approximations of the experimental records of damped
oscillations by simple analytical functions can give reliable
estimates of dynamically important frequency-dependent
parameters such as added mass and damping. However, with
some variations, this technique has been widely used by many
authors (see, for example, Pylnev and Razumeenko 1991) as an
effective tool for data analysis. The advantage of this method is
that the comparison of experimental records with their
approximations gives an explicit illustration of the qualitative
difference between the response of a simple linear damped
oscillator and the response of a complicated system under
study. A quantitative comparison between the results obtained
following the two methods mentioned is one of the goals of the
present study.

2.3
Data aquisition system and data analysis
An IBM computer with sampling software and an 8-bit A/D
convertor was used to record the time histories of damped
oscillations of the pendulum (impulse response functions). The
sampling frequency was set at 12 Hz. Although the resolution
of the A/D conversion was relatively low, the position of the
cylinder centre could be measured with the resolution of
0.02 mm what proved to be sufficient for the purposes of the
present study. The recording system was initiated before the
action of the impulse. For the initial condition of the time
history (system at rest) the value of the pendulum inclination
was set zero. Then this part of the time history was cut away
and the standard algorithm of the fast Fourier transform was
applied to the impulse response function. In order to get
a reasonably good frequency resolution, the history of the
motion was recorded during a time interval of about 1520
periods of damped oscillations. The data analysis was conducted according to expressions given by formulas (5)(8) with
the static calibration data used as input for the value of the
restoring force coefficient c. Finally, the plots of k and j versus
u (or the value of these coefficients at the frequency X, when
the method of least square approximations was in use) were
obtained and the experiment was repeated at another value of
the restoring force coefficient.
Let us make a brief note concerning the experimental
evaluation of "R(0)" introduced in Sect. 2.2. The algorithm of
the fast Fourier transform applied to the digitally represented
curve r(t) yields "R" and h for a set of discrete values u . For
i
i.1 the values "R(u )" represent a smooth curve having
i
a resonant peak at a certain frequency u :u . The value
k
*
"R(0)" corresponds to zero frequency u :0. However, an
i0
experimental record r(t) normally contains a small additive
term due to zero drift or, simply, due to the discrete nature of
analog-to-digital conversion. As a result, the value "R(0)" falls
out of a smooth curve "R(u )". To evaluate "R(0)", one
i
can make an extrapolation. In the present study we take
"R(0)"+"R(u )" as the first approximation. Practice shows
1
that this is sufficient. The unaccuracy introduced by this
estimate is noticeable only in the immediate vicinity of u . So,
1

155

for the final dependencies k(u ) and j(u ) the points correi
i
sponding to the lowest frequencies, where i:0. . . 3, were not
taken into account.

2.4
Dimensionless parameters

156

The dynamic interaction of an oscillating circular cylinder with


a linearly stratified fluid depends on the physical properties of
the fluid, the law of the body motion and the geometric setup of
the problem. For a cylinder oscillating harmonically with the
amplitude a, it is convenient to introduce the following set of
the most important non-dimensional parameters: a/D the
KeuleganCarpenter number, D(u/l the Stokes number,
u/N the normalized frequency which represents a version of
the Froude number.
In the present study, we consider only the case of low
KeuleganCarpenter numbers (a/D:0.14). The tests have
demonstrated that the magnitude of the experimental impulse
response functions is related linearly to the magnitude of the
impulse excitation. Correspondingly, the estimated values of
the added mass and damping coefficient do not depend on a/D
both for homogeneous and stratified fluid to within the
accuracy of the experiments.
The fact that the spatial structure of internal waves depends
both on D(u/l and u/N is clear from the theoretical
considerations and experimental data discussed in Makarov
et al. (1990) and in Hurley and Keady (1997). However, for
the range of parameters studied in the present experiments,
we can expect that the effects predicted for the inviscid case
in Hurley (1997) are dominant. The laboratory study of
damped oscillations of a sphere in a linearly stratified
fluid presented in Larsen (1969) also suggests the
dominant role of the inviscid scenario. Therefore, the
normalized frequency u/N is considered in the present study
as the basic dynamically important parameter. The dimensionless added mass and damping coefficients are introduced as
follows
C :k/o W
I
0
C :j/o WN
H
0

(10)

where o is the fluid density at the depth corresponding to the


0
cylinder center (in experiments o :1.011 g/cm3) and W is the
0
volume of the submerged cylinder.
For homogeneous fluids, the fluid-body interaction
is governed by viscous effects; the actual value of the
BruntVaisala frequency is zero what makes normalization
(10) inappropriate. In that case, we use normalization (10)
formally by adopting the value of N which took place in the
experiments with the stratified fluid. Thus, the denominator
o WN is common both for stratified and homogeneous
0
fluid what is convenient for the explicit quantitative comparison.
Mention should be also made of the nondimensional ratio
D/H characterizing the geometry of the problem. In the present
experiments D/H:0.132. This value is considered sufficiently
small to match, in practical terms, the assumption of infinite
fluid depth adopted in Hurley (1997) and Hurley and Keady
(1997). Some comments concerning the influence of the
limited fluid depth are given in Sect. 2.5.

2.5
Results
As one of the major objectives of the present study is to
compare the dynamics of body oscillations in stratified and
homogeneous fluid, a series of experiments in homogeneous
fluid (water) has been conducted. The experimental response
functions for this case are found to have very slight deviation
from an exponentially decaying sine function. The memory
effect is weak. The frequency-dependence is detectable only for
the damping coefficient, while the added mass coefficient is
constant being equal to the theoretical value for an infinite
ideal fluid (C :1) to within the experimental accuracy. The
I
complete quantitative results and discussion for this series are
presented below along with the data obtained for the stratified
fluid.
Two distinct types of the impulse response functions were
observed in the experiments with the stratified fluid. To
illustrate the difference between the dynamic system considered and a simple linear damped oscillator, the experimental records (solid lines) are shown in Figs. 2 and 3 along
with the least square approximations of these curves by
exponentially decaying sine functions (dashed lines). As
the experimental system is found to be linear, the response
functions r(t) and their approximations r (t) are normalized
a
by the maximum magnitude of the experimental response
r
for each test and plotted versus nondimensional time

tN/2n.
When the frequency u corresponding to the peak value of
*
the Fourier image of the impulse response function is below the
BruntVaisala frequency N (i.e. u /N:1), the damping is
*
essentially conditioned by the radiation of internal waves.
A typical experimental record for this case is shown in Fig. 2
for z :15 cm, c:0.154 N/m. As it is easy to see from this
0
figure, there is an important difference between the behavior of
the experimental system and the behavior of a simple linear
damped oscillator. The time interval between the moments
corresponding to r (t):0 is constant. In contrast, the experia
mental response r(t) develops over time in such a way that the
succeeding intervals between the moments corresponding to
r(t):0 gradually decrease, what implies a strong dependence
of k and j on frequency. Qualitatively, one can readily say that
k(R)9k(0) as the inertial properties of such a system at t;0
and t;R are governed by the added mass coefficients
corresponding to u;R and u;0, respectively (see, Huskind
1947). The variation of the characteristic time taken by each
cycle of the damped oscillations implies that the cylinder
motion generates the wave field which consists of a wide
spectrum of elementary wave disturbances. As a result, the
impulse response function recorded at u /N:1 contains very
*
rich information on the wave phenomena in a wide range of
non-dimensional frequencies u/N. In other words, the system
has long memory in that case.
Let us note that for very small values of the restoring force
coefficient c:0.022 N/m a critical damping is observed, i.e.
the pendulum, being disturbed, reaches a certain maximum
inclination and then approaches its equilibrium position
monotoneously. The logarithmic plot of such a curve shows
that for t;R the disturbance decays as t\1.
For u /N91 the damping is mainly conditioned by viscous
*
effects. However, the presence of stratification affects the

157

Fig. 2. Example of experimental impulse response function at


u /N:1 (z :15 cm, c:0.154 N/m, u /N:0.74)
*
0
*

Fig. 4. Non-dimensional added mass coefficient (C vs. u/N)


I

Fig. 3. Example of experimental impulse response function at


u /N91 (z :10 cm, c:0.394 N/m, u /N:1.11)
*
0
*
Fig. 5. Non-dimensional damping coefficient (C vs. u/N)
H

initial part of the time history. A typical record of damped


oscillations for this case is shown in Fig. 3, for z :10 cm,
0
c:0.394 N/m. The difference between the experimental curve
(solid line) and its least squares approximation (dashed line) is
quite appreciable only for the magnitude of the first oscillation.
It is well known that the harmonic oscillations of a body in
the linearly stratified fluid do not produce propagating internal
waves when u/N91. However, the process of damped oscillations is both non-stationary and time-dependent. As a result,
a certain portion of the energy is radiated by internal waves
even in the case when u /N91. The comparison of the curves
*
shown in Fig. 3 suggests that the effective generation of a wave
disturbance, which serves as an additional damping factor,
takes place only during the first period of damped oscillations.
The difference between the curves decreases with increasing
u /N; it practically vanishes when u /N.2. The overall
*
*
character of the impulse response function shown in Fig. 3
implies that it contains mainly the information about the
viscous effects while the information on the effects due to
stratification is mainly confined within the first cycle of
oscillations. In this case the experimental system has a short
memory.
The dependencies C (u/N) and C (u/N) obtained from the
I
H
experimental records of impulse response functions according

to the methods presented in Sect. 2.2 are shown in Figs. 4 and 5,


respectively, along with theoretical curves by Hurley (1997)
(for u/N:1, C :0 and C :(19(u/N)2, for u/N91, C :0
H
I
H
and C :(19(N/u)2. The values obtained in the experiments
I
with homogeneous fluid are marked by crosses. Other symbols
correspond to the tests conducted at different values of u /N
*
as specified in Table 1. The theoretical value of the nondimensional added mass coefficient for ideal unbounded homogeneous fluid (C :1) is plotted in Fig. 4 by a dashed line. The
I
results obtained in the experiments with homogeneous fluid
agree qualitatively with the asymptotical solution for a cylinder
oscillating in viscous fluid at a/D1 and D(u/l1 (see,
Landau and Lifshitz 1959). The asymptotic theory predicts
C :1 and j:u0.5. A quantitative disagreement between
I
asymptotic theory and experiments for j(u) may be attributed
to the relatively low value of D(u/l in the experiments
(25:D(u/l:40). The effects of the lubrication flow in the
gaps between the tank walls and the cylinder ends are believed
to be negligibly small as discussed in Sect. 2.1.
In the case of the stratified fluid, the qualitative and
quantitative agreement between theoretical predictions by

Table 1. Legend for Figs. 4 and 5

Stratified fluid

158

Method of least square approximations


Homogeneous water

u /N
*

Symbol

0.57
0.74
0.85
0.93
1.11
1.4

Hurley (1997) and the present experiments is found to be very


good. The comparison of the data obtained in the experiments
with stratified and homogeneous fluid shows that the presence
of stratification can drastically increase the damping of
oscillations and reduce the added mass at u/N:1. It is
interesting to note also that for u/N91 the damping coefficient
C takes somewhat lower values than in the case of homogeneH
ous water. However, it should be kept in mind that these effects
are dependent both on u/N and D(u/l. A significant
variation of the second parameter can essentially change the
character of the fluid-body interaction (see, Hurley and Keady
1997).
A careful examination shows that the tests conducted at
u /N+0.8 give more reliable information within a wider
*
frequency range than the other tests. As one can derive from
multiplying Eq. (4) by xR , integrating and taking an average over
one period of oscillation, the mean power radiated with
internal waves is related to the damping coefficient by
(11)
P:1 ja2u2
2
Substituting Hurleys (1997) result for j, we obtain
P:1 oWa2u2(N29u2. For the constant amplitude of
2
oscillation, the last expression takes a maximum value at
u/N:(2/3+0.8. In the experiments, for the maximum
radiated power we obtain a maximum information about the
system in the frequency domain.
The results of the analysis performed with the use of least
square approximations are also shown in Figs. 4, 5 by black
squares. As one can see, the results provided by two different
methods are in good agreement. For the experiments conducted at u /N:1 this fact seems surprising because the impulse
*
response function and its approximation in this frequency
range show quite different behaviors as illustrated in Fig. 2. It
is interesting to compare the Fourier transforms of these
curves. For the curves shown in Fig. 2 the corresponding
non-dimensional dependencies "R(u/N)"/"R(u /N)" and
*
"R (u/N)"/"R(u /N)" for the experimental response function
a
*
and its approximation are shown in Fig. 6 by solid and dashed
lines, respectively. The curves are different in details as the
added mass and damping coefficient of the actual physical
system are frequency-dependent while the Fourier-image of the
exponentially decaying sine function reflects the properties
of the approximating system having constant dynamic parameters. In spite of this important difference, the curves
"R(u/N)"/"R(u /N)" and "R (u/N)"/"R(u /N)" have common
*
a
*
asymptotics for u/N;0 and u/N;R; the location and

Fig. 6. Non-dimensional amplitudes of frequency response functions


for experimental response (solid line) and its approximation (dash
line) ("R(u/N)"/"R(u /N)" and "R (u/N)"/"R(u /N)"). Test
*
a
*
conditions correspond to Fig. 2

magnitude of the resonant peaks also coincide. Analogous


behavior is observed for other samples of the experimental
response functions. This fact may be of interest for theoretical
studies of the systems having memory effects.
An interesting property of Hurleys (1997) result is that the
dependencies C (u/N) and C (u/N) are the same for any
I
H
direction of oscillations of a cylinder. The present experiments
are focussed on the study of the horizonal oscillations.
However, there exists a non-direct proof that the dependencies
C (u/N) and C (u/N) are the same for any direction of
I
H
oscillations. The internal wave pattern of St. Andrew cross for
pure horizontal or vertical oscillations of a cylinder in linearly
stratified fluid is well known (see Mowbray and Rarity 1967;
Hurley 1969; Appleby and Crighton 1986; Ivanov 1989;
Makarov et al. 1990). The characteristic feature of this pattern
is that the waves propagate within four stripes which are
inclined at the angle a:arcsin(u/N) to the horizontal.
Gavrilov and Ermanyuk (1997) present an experimental study
of internal waves generated by a circular cylinder whose center
describes a circular path of small radius. Such a motion may be
represented as a sum of horizontal and vertical oscillations of
equal amplitude shifted in phase by 90. It is shown that in this
case the waves propagate within only two stripes. For the
clockwise direction of the cylinder motion the internal waves
are confined within a band which goes through the first and the
third quadrants of the Cartesian coordinate system; the origin
of the system is taken in the center of the cylinder trajectory. It
means that the wave fields produced by vertical and horizontal
oscillations are practically identical so that they can cancel
each other or be summed up depending on their relative phase
shifts in the different regions of space. It is known that for
purely horizontal or vertical oscillations exactly the quarter of
total energy is radiated along each stripe of the wave pattern.
This fact is evident from the symmetry of the problem. Thus,
the cancellation of waves within two stripes observed in
Gavrilov and Ermanyuk (1997) implies that the total energy
radiated with waves is identical for vertical and horizontal
directions of oscillations. The identity of the damping coefficients follows from Eq. (11). Let us note that the damping
coefficient j(u) is related with the value k(u)9k(R) by the

Kramers-Kronig relation (see, Kotik and Mangulis 1962;


Wehausen 1971). For u;R the stratified fluid acts as
a homogeneous one. Consequently, for an unbounded stratified fluid the value k(R) is identical for any direction of
oscillations. Therefore, the dependencies k(u) and j(u) are
universal for the vertical and horizontal directions of oscillations. Moreover, as far as the linear effects are concerned, the
perturbation introduced by the cylinder oscillations of amplitude a at the angle to the horizon t can be represented as
a sum of perturbations due to horizontal and vertical oscillations with the amplitudes a cos t and a sin t, respectively.
Correspondingly, we can conclude that the total energy
radiated with waves (which depends on the amplitude
squared) is the same for any direction of oscillations as
a2(cos2 t;sin2 t):a2. Finally, the above discussion implies
that the values of the dynamic coefficients k(u) and j(u) do
not depend on the direction of the cylinder oscillations in the
unbounded stratified fluid with N:const.
Let us make a brief note concerning the influence of the
limited fluid depth. In the present experiments the ratio
D/H:0.132 is sufficiently small so that the effect of flow
contraction on the characteristics measured in the homogeneous fluid is negligible. However, in the stratified fluid internal
wave rays emitted by the oscillating cylinder at u/N-1 can be
reflected by free surface and bottom and fall back on the
cylinder (reflection at the butt-ends of the test tank was
cancelled by the wave breakers). A simple consideration of
internal wave ray geometry shows that this phenomena takes
place when cos a-D/H, what implies u/N.(19(D/H)2. For
the present experimental conditions it means that the effect
of limited depth is very weak as it may manifest itself only in
the immediate vicinity of the BruntVaisala frequency when
0.99-u/N-1.
Experiments are under way to study the influence of the
limited depth on the characteristics of a body oscillating in
a stratified fluid.

3
Summary
The damped horizontal oscillations of a circular cylinder
submerged in stratified fluid with a linear density profile are
studied experimentally. The experimental time-histories of
responses to an impulse excitation (impulse response functions) are analysed by Fourier-transforming the problem from
time- to frequency-domain. It is observed that the impulse
response functions have different behavior depending on the
ratio between the frequency of oscillations and BruntVaisala
frequency. When the time-history of damped oscillations is
governed by the wave phenomena, the memory effects are
strong so that the evaluation of frequency-dependent dynamic
coefficients can, in principle, be performed based on a single
experimental realization of the impulse response function.
The measurements demonstrate large effects of stratification
on the added mass and damping coefficients. The results of
experiments confirm theoretical results by Hurley (1997). For
the frequency of oscillations below the BruntVaisala frequency the damping is much greater than in homogeneous
fluid, being primarily conditioned by the radiation of internal
waves. The added mass of a circular cylinder reduces to very
low values and amounts to a few percent of its value for

homogeneous fluid. As the frequency of oscillations increases


above the BruntVaisala frequency, the values of added mass
and damping coefficient asymptotically approach the values
typical for homogeneous fluid.
A simple analysis of experimental records implying the use
of least square approximations of the impulse response
functions by exponentially decaying sine functions is tested.
The results are found to be in good agreement with a strict
(from the theoretical point of view) analysis using the Fourier
transform. It is shown that the Fourier images of the impulse
response functions and their approximations have common
peak values and common asymptotics at u;0 and u;R
while being different in details.

References
Akulenko LD; Nesterov SV (1987) Oscillations of a rigid body at
interface of two fluid. Izvestiya Akademii Nauk SSSR, Mekhanika
Tverdogo Tela, No. 5: 3440 (in Russian)
Akulenko LD; Mikhailov SA; Nesterov SV; Chaikovsky AA (1988)
Numerical-analytical investigation of rigid body oscillations on
interface of two fluid. Izvestiya Akademii Nauk SSSR, Mekhanika
Tverdogo Tela, No. 4: 5966 (in Russian)
Appleby JC; Crighton DG (1986) Non-Boussinesq effects in the
diffraction of internal waves from an oscillating cylinder. Quart
J Mech Math 39: 209231
Cummins WE (1962) The impulse response function and ship
motions. Schiffstechnik 9: 101109
Gavrilov NV; Ermanyuk EV (1997) Internal waves generated by
circular translational motion of a cylinder in a linearly stratified
fluid. J Appl Mech Tech Phys 38: 224227
Hurley DG (1969) Emission of internal waves of vibrating cylinders.
J Fluid Mech 36: 657672
Hurley DG (1997) The generation of internal waves by vibrating elliptic
cylinders. Part 1. Inviscid solution. J Fluid Mech 351: 105118
Hurley DG; Keady G (1997) The generation of internal waves by
vibrating elliptic cylinders. Part 2. Approximate viscous solution.
J Fluid Mech 351: 119138
Huskind MD (1947) Methods of hydrodynamics in the problems of
ship seakeeping in waves. Trudy TSAGI, No. 603: 174 (in Russian)
Ivanov AV (1989) Internal waves generation by an oscillation source.
Izvestiya Akademii Nauk SSSR, Fizika Atmosfery I Okeana, 25:
8489 (in Russian)
Kotik J; Mangulis V (1962) On the KramersKronig relations for ship
motions. Int Shipbuilding Prog 9: 351368
Landau LD; Lifshitz EM (1959) Fluid mechanics. Oxford: Pergamon
Larsen LH (1969) Oscillations of a neutrally buoyant sphere in
a stratified fluid. Deep-Sea Res 16: 587603
Makarov SA; Nekhlyudov VI; Chashechkin Yu D (1990) Space
structure of two-dimensional monochromatic internal wave pakets
in exponentially-stratified fluid. Izvestiya Akademii Nauk SSSR,
Fizika Atmosfery I Okeana, 26: 744754 (in Russian)
Mowbray DE; Rarity BSH (1967) A theoretical and experimental
investigation of the phase configuration of internal waves of small
amplitude in a density stratified fluid. J Fluid Mech 28: 116
Pylnev Yu V; Razumeenko Yu V (1991) Investigation of damped
oscillations for a float of a special shape deeply imbedded into
homogeneous and stratified fluid. Izvestiya Akademii Nauk SSSR,
Mekhanika Tverdogo Tela, No. 4: 7179 (in Russian)
Sarpkaya T (1986) Force on a circular cylinder in viscous oscillatory
flow at low Keulegan-Carpenter numbers. J Fluid Mech 165: 6171
Sretenskii LN (1937) On the damping of vertical oscillations of the
gravity center of floating bodies. Trudy TSAGI, No. 330: 112 (in
Russian)
Vargaftik NB (1963) Handbook on the thermophysical properties of
gases and liquids. Moscow: Gos. Izdat. Fiz-mat.lit (in Russian)
Wehausen JV (1971) The motion of floating bodies. Ann Rev Fluid
Mech 3: 237268

159

You might also like