Sei sulla pagina 1di 266

University of Southampton Research Repository

ePrints Soton

Copyright and Moral Rights for this thesis are retained by the author and/or other
copyright owners. A copy can be downloaded for personal non-commercial
research or study, without prior permission or charge. This thesis cannot be
reproduced or quoted extensively from without first obtaining permission in writing
from the copyright holder/s. The content must not be changed in any way or sold
commercially in any format or medium without the formal permission of the
copyright holders.

When referring to this work, full bibliographic details including the author, title,
awarding institution and date of the thesis must be given e.g.
AUTHOR (year of submission) "Full thesis title", University of Southampton, name
of the University School or Department, PhD Thesis, pagination

http://eprints.soton.ac.uk

UNIVERSITY OF SOUTHAMPTON

Interactions of a submerged
membrane with water waves and
its use in harnessing nearshore
wave power.
by
Nicolas Choplain

A thesis submitted in partial fulfillment for the


degree of Doctor of Philosophy
in the
Faculty of Engineering and the Environment
Energy and Climate Change Research Group

May 2012

UNIVERSITY OF SOUTHAMPTON
ABSTRACT
FACULTY OF ENGINEERING AND THE ENVIRONMENT
ENERGY AND CLIMATE CHANGE RESEARCH GROUP
Doctor of Philosophy
by Nicolas Choplain

Developed and developing countries need electricity, and this usage is increasing everyday. This constant increase cannot be satisfied with the current ways of electricity
generation that have shown themselves to be out of phase with environmental concerns.
Oceans yield great amount of energy that could be converted into electricity and the
current research deals with one portion of ocean energy, wave energy.
The wave energy converter studied in this thesis is a bottom-mounted, liquid-filled
rectangular duct, covered with a rubber membrane and aligned head to the waves propagation direction. Two types of membrane were tested. The behaviour of this device
beneath waves was investigated with two configurations: one with both its ends closed,
the other one with a power take-off connected at its stern. The pressure at both ends
was characterised by means of pressure transducers and the pressure inside the duct by
means of laser sensors measuring the membrane displacement.
Results from experiments carried out on the closed version of the duct pointed out
a resonant behaviour of the system for wave frequencies at which bulges, propagating in
the rubber membrane, could travel an integer number of times along the ducts length.
This resonance was characterised by pressure magnitudes at the stern up to 2.8 times
that acting on the membrane from the incident wave. Moreover, the membrane displacement was for the first time mapped and the profile obtained showed characteristic
nodes and antinodes.
The performance of this device in harnessing wave power was evaluated by connecting a linear dashpot at its stern. Capture widths of up to 2.2 times the device width
were obtained and the bandwidth of maximum power capture not limited to a single
frequency.
The pressure behaviour in both configurations was explained with a one dimensional
theory of bulges propagating in distensible tubes with good agreement for the thicker
tested membrane. On the contrary, this was not the case for the thinner membrane,
suggesting that this model could not be used for configurations where bulge wavelengths
are much shorter than that of the incoming wave.

Contents
Abstract

iii

List of Figures

xi

List of Tables

xxiii

Declaration Of Authorship

xxv

Acknowledgements

xxvii

Nomenclature

xxix

1 Introduction
1.1 Motivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
1.2 Thesis layout . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2 Wave energy glossary and relevance of the study
2.1 Introduction . . . . . . . . . . . . . . . . . . . . . .
2.2 Wave characteristics . . . . . . . . . . . . . . . . .
2.2.1 Introduction . . . . . . . . . . . . . . . . . .
2.2.2 Definitions . . . . . . . . . . . . . . . . . . .
2.2.3 Kinetic and potential energies . . . . . . . .
2.3 Wave power background . . . . . . . . . . . . . . .
2.3.1 Introduction . . . . . . . . . . . . . . . . . .
2.3.2 Wave power resource distribution . . . . . .
2.3.3 Harnessing concepts of wave power . . . . .
2.3.3.1 Introduction . . . . . . . . . . . .
2.3.3.2 Oscillating water column . . . . . .
2.3.3.3 Overtopping devices . . . . . . . .
2.3.3.4 Oscillating bodies . . . . . . . . . .
2.3.4 Conclusions . . . . . . . . . . . . . . . . . .
2.4 Relevance of the study . . . . . . . . . . . . . . . .
2.4.1 Introduction . . . . . . . . . . . . . . . . . .
2.4.2 The Anaconda WEC . . . . . . . . . . . . .
2.4.2.1 Principle . . . . . . . . . . . . . .
2.4.2.2 The bulge wave theory . . . . . . .
v

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

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

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

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

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

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

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

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

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

1
1
3
5
5
5
5
6
7
9
9
10
11
11
11
12
12
13
14
14
14
14
15

vi

CONTENTS

2.5

2.6

2.4.2.3 Current developments . . . . . . . . . . . . . . . .


2.4.2.4 Conclusions . . . . . . . . . . . . . . . . . . . . . .
2.4.3 Nearshore wave power . . . . . . . . . . . . . . . . . . . . .
2.4.3.1 Introduction . . . . . . . . . . . . . . . . . . . . .
2.4.3.2 Exploitable wave power . . . . . . . . . . . . . . .
2.4.3.3 Nearshore bottom-mounted wave energy converters
2.4.3.4 Conclusions . . . . . . . . . . . . . . . . . . . . . .
Starting point of the project . . . . . . . . . . . . . . . . . . . . . .
2.5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.5.2 Aim and objectives . . . . . . . . . . . . . . . . . . . . . . .
2.5.3 Novelty and advantages of the studied device . . . . . . . . .
Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

3 Literature review
3.1 Introduction . . . . . . . . . . . . . . . . . . . . .
3.2 Use of rubber for marine applications . . . . . . .
3.3 Breakwaters . . . . . . . . . . . . . . . . . . . . .
3.3.1 Introduction . . . . . . . . . . . . . . . . .
3.3.2 Submerged elastic membrane breakwaters
3.3.3 Submerged fluid-filled membranes . . . . .
3.3.4 Bottom-mounted fluid-filled membranes . .
3.3.5 Conclusions on breakwaters studies . . . .
4 Experimental equipment
4.1 Introduction . . . . . . . . . . . . . . .
4.2 Narrow wave flume . . . . . . . . . . .
4.2.1 Dimensions . . . . . . . . . . .
4.2.2 Wave-maker . . . . . . . . . . .
4.2.3 Absorbing beach . . . . . . . .
4.3 Wave basin . . . . . . . . . . . . . . .
4.3.1 Dimensions . . . . . . . . . . .
4.3.2 Wave-maker . . . . . . . . . . .
4.3.3 Absorbing beach . . . . . . . .
4.4 Rubber . . . . . . . . . . . . . . . . . .
4.4.1 Introduction . . . . . . . . . . .
4.4.2 Youngs modulus . . . . . . . .
4.4.3 Structural damping . . . . . . .
4.4.4 Summary . . . . . . . . . . . .
4.5 Power take-off system . . . . . . . . . .
4.5.1 Introduction . . . . . . . . . . .
4.5.2 Components . . . . . . . . . . .
4.5.3 Power calculation . . . . . . . .
4.6 Instrumentation . . . . . . . . . . . . .
4.6.1 Data acquisition and equipment

18
19
21
21
21
24
25
26
26
26
27
28

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

31
31
31
33
33
34
43
44
54

. . . . . . . . .
. . . . . . . . .
. . . . . . . . .
. . . . . . . . .
. . . . . . . . .
. . . . . . . . .
. . . . . . . . .
. . . . . . . . .
. . . . . . . . .
. . . . . . . . .
. . . . . . . . .
. . . . . . . . .
. . . . . . . . .
. . . . . . . . .
. . . . . . . . .
. . . . . . . . .
. . . . . . . . .
. . . . . . . . .
. . . . . . . . .
driving system

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

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

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

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

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

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

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

57
57
57
57
59
59
59
59
59
61
62
62
63
64
65
65
65
67
68
70
70

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

vii

CONTENTS

4.7
4.8

4.6.2 Wave probes . . . . . . . . .


4.6.3 Water pressure transducers .
4.6.4 Air pressure transducers . .
4.6.5 Digital camera . . . . . . .
4.6.6 Laser sensors . . . . . . . .
Tested model . . . . . . . . . . . .
Conclusions . . . . . . . . . . . . .

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

5 One dimensional model for bulges propagation in the


5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . .
5.2 Load-deflection of rectangular membrane and free bulge
5.3 Bulge equation . . . . . . . . . . . . . . . . . . . . . .
5.4 Closed ends configuration . . . . . . . . . . . . . . . .
5.5 Duct with a power take-off system . . . . . . . . . . . .
5.5.1 Bulge pressure . . . . . . . . . . . . . . . . . . .
5.5.2 Power calculation . . . . . . . . . . . . . . . . .
5.6 Conclusions . . . . . . . . . . . . . . . . . . . . . . . .

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

71
71
72
73
73
74
76

duct
. . . .
speed
. . . .
. . . .
. . . .
. . . .
. . . .
. . . .

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

77
77
77
79
80
84
84
87
87

6 Narrow flume experiments: static deflection, free bulge speed and


pressure amplification
89
6.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89
6.2 Experimental layout . . . . . . . . . . . . . . . . . . . . . . . . . . 89
6.3 Static tests: 1mm thick membrane . . . . . . . . . . . . . . . . . . 91
6.3.1 Membrane deflection . . . . . . . . . . . . . . . . . . . . . . 91
6.3.2 Free bulge speed . . . . . . . . . . . . . . . . . . . . . . . . 93
6.4 Membrane beneath waves: 1mm thick membrane . . . . . . . . . . 94
6.4.1 Experimental conditions . . . . . . . . . . . . . . . . . . . . 94
6.4.2 Pressure amplification . . . . . . . . . . . . . . . . . . . . . 96
6.4.3 Membrane displacement and pressure in the duct . . . . . . 99
6.4.4 Bulge pressure components . . . . . . . . . . . . . . . . . . . 102
6.4.4.1 Separation of components . . . . . . . . . . . . . . 102
6.4.4.2 Components at resonance . . . . . . . . . . . . . . 104
6.4.5 Wave analysis and loss in the rubber . . . . . . . . . . . . . 111
6.4.6 Transverse motion of the membrane . . . . . . . . . . . . . . 116
6.4.7 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . 120
6.5 Thin membrane results . . . . . . . . . . . . . . . . . . . . . . . . . 120
6.5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . 120
6.5.2 Experimental conditions . . . . . . . . . . . . . . . . . . . . 121
6.5.3 Static deflection and free bulge speed . . . . . . . . . . . . . 121
6.5.4 Pressure variation at the stern . . . . . . . . . . . . . . . . . 125
6.5.5 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . 125
7 Extractable power
131
7.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 131
7.2 Experimental layout and test conditions . . . . . . . . . . . . . . . 131

viii

CONTENTS

7.3

7.4

7.5
7.6
7.7

7.2.1 Description . . . . . . . . . . . . . . . . . . . . . . .
7.2.2 Power take-off in experimental conditions . . . . . . .
Matching impedances configuration . . . . . . . . . . . . . .
7.3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . .
7.3.2 Power capture . . . . . . . . . . . . . . . . . . . . . .
7.3.3 Pressure variation in the duct . . . . . . . . . . . . .
7.3.4 Bulge components . . . . . . . . . . . . . . . . . . .
Mismatch impedances configuration . . . . . . . . . . . . . .
7.4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . .
7.4.2 Power capture . . . . . . . . . . . . . . . . . . . . . .
7.4.3 Pressure variation in the duct and bulge components
Water-air interface displacement and pressure in the PTO .
Wave analysis . . . . . . . . . . . . . . . . . . . . . . . . . .
Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . .

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

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

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

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

131
133
137
137
138
140
143
147
147
148
150
152
155
158

8 Two dimensional analytical model of a duct covered with a tensioned membrane beneath waves
161
8.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 161
8.2 Problem formulation for the closed ends configuration . . . . . . . . 162
8.2.1 Diffraction problem . . . . . . . . . . . . . . . . . . . . . . . 164
8.2.1.1 Definitions of the potentials . . . . . . . . . . . . . 164
8.2.2 Radiation problem . . . . . . . . . . . . . . . . . . . . . . . 165
8.2.2.1 Definition of the potentials . . . . . . . . . . . . . 165
8.2.2.2 Membrane deformation . . . . . . . . . . . . . . . 167
8.2.3 Reflection and transmission coefficients . . . . . . . . . . . . 168
8.2.4 Conservation of volume in the duct . . . . . . . . . . . . . . 169
8.2.4.1 Uniform potential below the membrane . . . . . . . 170
8.2.4.2 Weighted least-square method . . . . . . . . . . . . 170
8.2.4.3 Matrix approximation . . . . . . . . . . . . . . . . 171
8.2.5 Code Layout . . . . . . . . . . . . . . . . . . . . . . . . . . 171
8.2.6 Case studies . . . . . . . . . . . . . . . . . . . . . . . . . . . 175
8.2.6.1 Diffraction problem . . . . . . . . . . . . . . . . . . 175
8.2.6.2 Complete configuration and comparison of the proposed
methods . . . . . . . . . . . . . . . . . . . . . . . . 175
8.3 Experimental verification . . . . . . . . . . . . . . . . . . . . . . . . 186
8.4 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 190
9 Conclusions and future work
193
9.1 Achievements and perspectives . . . . . . . . . . . . . . . . . . . . . 193
9.2 Future work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 195
A Power take-off test in the narrow flume
A1
A.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . A1
A.2 Experimental set-up and test conditions . . . . . . . . . . . . . . . A2

ix

CONTENTS

A.3 Linearity of the power take-off . . . . . . . . . . . . . . . . . . . . . A3


A.4 Power capture . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . A5
B Analytical model of duct covered
the maths
B.1 Introduction . . . . . . . . . . . .
B.2 Closed ends configuration . . . .
B.2.1 Diffraction problem . . . .
B.2.2 Radiation problem . . . .
B.2.3 Membrane deformation . .

with a tensioned membrane:


.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

B1
B1
B1
B1
B5
B8

C Pictures of the experimental set-ups

C1

D Determination of rubber properties


D.1 Introduction . . . . . . . . . . . . .
D.2 Youngs modulus . . . . . . . . . .
D.3 Incompressibility of rubber . . . . .
D.4 Dynamic damping . . . . . . . . . .

D1
D1
D2
D4
D5

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

List of Figures
2.1
2.2
2.3
2.4
2.5
2.6
2.7

Sketch of a water wave . . . . . . . . . . . . . . . . . . . . . . . . .


Energy absorbed per unit time T . . . . . . . . . . . . . . . . . . .
Global distribution of annual mean wave power, after Cornett (2008)
Wave energy converters classification, from Rahm (2010) . . . . . .
Anaconda WEC, after Checkmate SeaEnergy Ltd. . . . . . . . . . .
Bulge wave propagating in the rubber tube. . . . . . . . . . . . . .
Capture widths as functions of relative wave periods for (a) Z=0.64,
(b) Z=1.05, (c) Z=3.3, (d) Z=7.7. Measurements are shown as
points, one-dimensional theory with = 13 as continuous lines,
after Chaplin et al. (2012) . . . . . . . . . . . . . . . . . . . . . . .
2.8 Challenges facing the development of a new wave energy converter .
2.9 Shoaling of a 10 second energy period wave propagating orthogonal to depth contours for different seabed slopes, after Folley and
Whittaker (2009). . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.10 Variation in percentage power loss from offshore to nearshore site
for an initial sea-state of T =10s and Pi =50 kW/m travelling at an
angle to the depth contours on a seabed slope of 1:100, after Folley
and Whittaker (2009). . . . . . . . . . . . . . . . . . . . . . . . . .
2.11 The offshore Oyster 1 device, after Cameron et al. (2010). . . . . .
2.12 3D view of the tested duct. . . . . . . . . . . . . . . . . . . . . . . .
3.1

3.2
3.3

3.4
3.5

3.6

Example cross-section of a Fabriconda with 10 cells, showing the


structure at its minimum cross-section (left) and at its medium
point (right), after Hann et al. (2011). . . . . . . . . . . . . . . . .
Vertical membrane configuration. . . . . . . . . . . . . . . . . . . .
Reflection coefficient of hinged-hinged membrane breakwater as function of dimensionless tension Tz /gh2 and wave number kh, after
Kim and Kee (1996). . . . . . . . . . . . . . . . . . . . . . . . . . .
Horizontal membrane configuration. . . . . . . . . . . . . . . . . . .
Reflection coefficients of a submerged impermeable membrane breakwater as function of length of membrane s/h and wavenumber k1 h
for d/h=0.2, Tx /gh2=0.1, after Cho and Kim (1998) . . . . . . . .
Response of the membrane (/A) as a function of wavenumber
k1 h and horizontal coordinate x/L for d/h = 0.2, s/h = 0.5 and
Tx /gh2 =0.1, after Cho and Kim (1998). . . . . . . . . . . . . . . .

xi

7
8
10
11
14
16

19
20

22

22
25
28

32
34

35
38

39

39

xii

LIST OF FIGURES
3.7

3.8

3.9
3.10
3.11
3.12

3.13

3.14
3.15

3.16
3.17
3.18
4.1
4.2
4.3
4.4
4.5
4.6

Modal response amplitude as function of wavenumber k1 h and horizontal coordinate x/L for d/h=0.2, s/h=0.5, Tx /gh2 =0.1; upper
plot: first symmetric mode, middle plot: second symmetric mode;
lower plot: third symmetric mode. . . . . . . . . . . . . . . . . . . .
Modal response amplitude as function of wavenumber k1 h and horizontal coordinate x/L for d/h=0.2, s/h=0.5, Tx /gh2 =0.1; upper
plot: first asymmetric mode, middle plot: second asymmetric mode;
lower plot: third asymmetric mode. . . . . . . . . . . . . . . . . . .
Wave interception by a pair of sea balloons, after Ijima and Uwatoko
(1985). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Sketch of the flexible mound device. . . . . . . . . . . . . . . . . . .
Wave energy loss in the flexible mound, after Tanaka et al. (1992a).
Amplitude of membrane normal deflection for p0 /gh = 0.07 and
d/h=0.5: (a) fundamental mode; (b) second mode, (c) third mode
( , i / = 0.9; , i / = 1.0; , i / = 1.1), after Phadke and
Cheung (1999); i , density of the inner fluid. . . . . . . . . . . . . .
Transmission coefficient as function of excitation dimensionless frequency L/: (a) p0 /gh = 0.065, (b) p0 /gh = 0.208; data
from Ohyama et al. (1989); two dimensional linear model from
Phadke and Cheung (1999); two dimensional non linear model
from Phadke and Cheung (2003); three dimensional model from
Das and Cheung (2009) . . . . . . . . . . . . . . . . . . . . . . . . .
Water-filled hemicircular membrane breakwater. . . . . . . . . . . .
Free surface wave amplitudes for normal waves over (a) rigid and
(c) flexible structure for = /6 and over (b) rigid and (d) flexible
structure for = /4, from Dewi et al. (1999). . . . . . . . . . . . .
Variation of shadow transmission coefficient KT with kR , after
Dewi et al. (1999). . . . . . . . . . . . . . . . . . . . . . . . . . . .
Side view of the hemi-cylindrical flexible composite breakwater model,
after Stamos (2000). . . . . . . . . . . . . . . . . . . . . . . . . . .
Rapidly installed breakwater system XM99 (from Briggs (2001)) . .

Narrow flume at the University of Southampton. . . . . . . . . . .


Wavemaker for the narrow flume. . . . . . . . . . . . . . . . . . .
Wave basin at the University of Southampton. . . . . . . . . . . .
Wave paddle for the wave basin. . . . . . . . . . . . . . . . . . . .
Absorbing beach for the wave basin. . . . . . . . . . . . . . . . . .
Measured absorbed reflection coefficients for the absorbing foam
beach in the wave basin. . . . . . . . . . . . . . . . . . . . . . . .
4.7 Stress in the strip of membrane as a function of the elongation
for the quasi-static Youngs modulus measurement. Each symbols
represent different loading. . . . . . . . . . . . . . . . . . . . . . .
4.8 Cross section power take-off. . . . . . . . . . . . . . . . . . . . . .
4.9 Schematic of the power take-off system. . . . . . . . . . . . . . . .
4.10 Power take-off system set-up . . . . . . . . . . . . . . . . . . . . .
4.11 Model for the power take-off. . . . . . . . . . . . . . . . . . . . . .

.
.
.
.
.

41

42
44
45
46

47

49
50

51
52
53
55
58
58
60
60
61

. 63

.
.
.
.
.

64
66
66
67
68

xiii

LIST OF FIGURES
4.12
4.13
4.14
4.15
4.16

Pressure transducers. . . . . . . . . . . . . . . . . . . . . . . .
Wheastone bridge for the pressure transducers. . . . . . . . . .
Laser sensor. . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Laser sensors set-up. . . . . . . . . . . . . . . . . . . . . . . .
Influence of the plastic boxes used for the laser sensors on the
flection coefficient for the experimental set-up as in chapter 7
two sets of laser sensors positions. . . . . . . . . . . . . . . . .
4.17 Cross section sketch of the rubber duct. . . . . . . . . . . . . .
5.1
5.2

5.3

5.4

5.5
6.1

. .
. .
. .
. .
refor
. .
. .

Cross section sketch . . . . . . . . . . . . . . . . . . . . . . . . . .


Theoretical components of the wave pressure in the rubber duct for
a static internal pressure head of 7cm corresponding to a free bulge
speed of 3.06m/s in the configuration of chapter 6. . . . . . . . . .
Theoretical internal velocities for the first four harmonics at 30 instants over one wave period for the duct with an internal static pressure head of 7cm; (a) f =0.306Hz; (b) f =0.612Hz; (c) f =0.918Hz
(d) f =1.224Hz. . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Theoretical internal pressure for the first four harmonics at 30 instants over one wave period for the duct with an internal static pressure head of 7cm; (a) f =0.306Hz; (b) f =0.612Hz; (c) f =0.918Hz
(d) f =1.224Hz. . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Definition sketch of the stern. . . . . . . . . . . . . . . . . . . . .

.
.
.
.

72
72
73
74

. 75
. 75
. 78

. 81

. 82

. 83
. 85

Experimental layout for the closed ends configuration. Dashed lines


correspond to different internal static pressure heads. . . . . . . . . 90
6.2 Experimental values of membrane centreline deflection (+ + +),
load-deflection relation 5.9 () . . . . . . . . . . . . . . . . . . . . 92
6.3 Equivalent membrane submergence depth . . . . . . . . . . . . . . . 92
6.4 Time series of pressure transducers from free bulge speed measurement 93
6.5 Free bulge speed for the 1mm thick membrane, each symbols representing one set of measurement. . . . . . . . . . . . . . . . . . . . 94
6.6 Time series of laser sensors when rubber membrane is beneath waves. 95
6.7 Time series of pressure transducers (at bow and stern) when rubber
membrane is beneath waves. . . . . . . . . . . . . . . . . . . . . . . 96
6.8 Magnitude of the pressure at the stern relative to the wave pressure
acting on the membrane with an internal static pressure head of
7cm. symbols: experimental values; : theory. . . . . . . . . . . . 97
6.9 Magnitude of the pressure at the stern relative to the wave pressure
acting on the membrane with an internal static pressure head of
10cm. symbols: experimental values; : theory. . . . . . . . . . . . 97
6.10 Pressure in the duct at thirty instants over one wave period with an
internal static pressure head of 7cm; wave frequency f =0.95Hz; (a)
theory; (b) inferred from laser sensors measurement of membrane
displacement. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100

xiv

LIST OF FIGURES
6.11 Pressure in the duct at thirty instants over one wave period with an
internal static pressure head of 7cm; wave frequency f =1.2Hz; (a)
theory; (b) inferred from laser sensors measurement of membrane
displacement. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100
6.12 Pressure in the duct at thirty different instants over one wave period with an internal static pressure head of 7cm; wave frequency
f =1.1Hz; (a) theory; (b) inferred from laser sensors measurement
of membrane displacement. . . . . . . . . . . . . . . . . . . . . . . . 101
6.13 Pressure in the duct at thirty instants over one wave period with an
internal static pressure head of 10cm; wave frequency f =1.075Hz;
(a) theory; (b) inferred from laser sensors measurement of membrane displacement. . . . . . . . . . . . . . . . . . . . . . . . . . . . 102
6.14 Separation of bulge wave components from laser measurements. . . 103
6.15 Measured total internal pressure, normalised by the pressure acting
on the membrane (a) and phase (b) for the duct with an internal static pressure head of 7cm, plotted as points. Lines represent
amplitudes and phases computed as the sum of the components
obtained from the least squares method. Wave frequency f =0.95Hz. 105
6.16 Pressure in the duct at thirty different instants over one wave period with an internal static pressure head of 7cm; wave frequency
f =0.95Hz; (a) inferred from laser sensor measurement of membrane
displacement amplitude; (b) sum of the three bulge wave components (and the wave pressure) obtained from a least squares method.105
6.17 Magnitude of Pb+ relative to the wave pressure acting on the membrane with an internal static pressure head of 7cm. symbols: experimental values inferred from a least squares method; : calculated
from (5.21). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106
6.18 Magnitude of Pb relative to the wave pressure acting on the membrane with an internal static pressure head of 7cm. symbols: experimental values inferred from a least squares method; : calculated
from (5.22). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107
6.19 Magnitude of PbA relative to the wave pressure acting on the membrane with an internal static pressure head of 7cm. symbols: experimental values inferred from a least squares method; : calculated
from (5.20). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107
6.20 Measured total internal pressure, normalised by the pressure acting
on the membrane (a) and phase (b) for the duct with an internal
static pressure head of 10cm, plotted as points. Lines represent
amplitudes and phases computed as the sum of the components
obtained from the linear squares method. Wave frequency f =1.075Hz.108
6.21 Pressure in the duct at thirty instants over one wave period with an
internal static pressure head of 10cm; wave frequency f =1.075Hz;
(a) inferred from laser sensor measurement of membrane displacement amplitude; (b) sum of the three bulge wave components (and
the wave pressure) obtained for a least squares method. . . . . . . . 109

LIST OF FIGURES

xv

6.22 Magnitude of Pb+ relative to the wave pressure acting on the membrane with an internal static pressure head of 10cm. symbols: experimental values inferred from a least squares method; : calculated from (5.21). . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109
6.23 Magnitude of Pb relative to the wave pressure acting on the membrane with an internal static pressure head of 10cm. symbols: experimental values inferred from a least squares method; : calculated from (5.22). . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110
6.24 Magnitude of PbA relative to the wave pressure acting on the membrane with an internal static pressure head of 10cm. symbols: experimental values inferred from a least squares method; : calculated from (5.20). . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110
6.25 Ratio of the transmitted wave power, Pt , to the incident wave power
Pi for the duct with an internal static pressure head of 7cm. Black
line and symbols: simple harmonic component of wave gauge signals; Grey line and symbols: first three harmonic components of
wave gauge signals. . . . . . . . . . . . . . . . . . . . . . . . . . . . 111
6.26 Ratio of the reflected wave power, Pr , to the incident wave power Pi
for the duct with an internal static pressure head of 7cm. Black line
and symbols: simple harmonic component of wave gauge signals;
Grey line and symbols: first three harmonic components of wave
gauge signals. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 112
6.27 Ratio of the wave power loss, Ploss , to incident wave power Pi for
the duct with an internal static pressure head of 7cm. Line and
symbols: simple harmonic analysis of wave gauge signals. . . . . . . 113
6.28 Instantaneous elongation in the rubber at 20 instants (plotted as
points) as a function of time normalised by the wave period T .
Wave frequency is f = 0.775Hz. . . . . . . . . . . . . . . . . . . . . 115
6.29 Ratio of the power lost in the rubber, Ploss , to the incident wave
power Pi for the rubber duct with an internal static pressure head
of 7cm. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115
6.30 Cross-section view of the duct and transverse motion of the membrane.117
6.31 Experimental values of membrane centreline deflection (+ + +),
load-deflection relation 5.9 () for y,0 = 0.14 and x,0 =0 . . . . . . 122
6.32 Experimental values of membrane centreline deflection (+ + +),
load-deflection relation 5.9 () for y,0 = 0.28 and x,0 =0. . . . . . . 122
6.33 Experimental values of membrane centreline deflection (+ + +),
load-deflection relation 5.9 () for y,0 = 0.14 and x,0 =0.14. . . . . 123
6.34 Averaged values of free bulge speed measurement, for y,0 = 0.14
and x,0 =0 for the 0.2mm membrane; () calculated values; (+++)
averaged values from pressure and laser measurements. . . . . . . . 124
6.35 Averaged values of free bulge speed measurement, for y,0 = 0.28
and x,0 =0 for the 0.2mm membrane; () calculated values; (+++)
averaged values from pressure and laser measurements. . . . . . . . 124

xvi

LIST OF FIGURES
6.36 Averaged values of free bulge speed measurement, for y,0 = 0.14
and x,0 =0.14 for the 0.2mm membrane; () calculated values; (++
+) averaged values from pressure and laser measurements. . . . . .
6.37 Magnitude of the pressure at the stern relative to the wave pressure
at the free surface elevation with conditions of test 1; left: measured
for the three runs (each symbols for one run); right: one dimensional
theory with c = 1.11m/s. . . . . . . . . . . . . . . . . . . . . . . . .
6.38 As for Figure 6.37, but with conditions of test 2; left: measured for
the three runs (each symbols for one run); right: one dimensional
theory with c = 0.88m/s. . . . . . . . . . . . . . . . . . . . . . . . .
6.39 As for Figure 6.37, but with conditions of test 3; left: measured
for the two runs (each symbols for one run); right: one dimensional
theory with c = 1.11m/s. . . . . . . . . . . . . . . . . . . . . . . . .
6.40 As for Figure 6.37, but with conditions of test 4; left: measured
for the two runs (each symbols for one run); right: one dimensional
theory with c = 0.88m/s. . . . . . . . . . . . . . . . . . . . . . . . .
6.41 As for Figure 6.37, but with conditions of test 5; left: measured
for the two runs (each symbols for one run); right: one dimensional
theory with c = 1.11m/s. . . . . . . . . . . . . . . . . . . . . . . . .
6.42 As for Figure 6.37, but with conditions of test 6; left: measured;
right: one dimensional theory with c = 0.88m/s. . . . . . . . . . . .
6.43 Surface elevation at the middle point along the length of the duct,
relative to the incident wave amplitude. . . . . . . . . . . . . . . . .
6.44 Ratio of the transmitted wave power, Pt , to the incident wave power
Pi for all the tests and measured with the first three harmonic components of wave gauge signals. . . . . . . . . . . . . . . . . . . . . .
6.45 Membrane displacement at thirty instants over one wave period for
configuration of test 1; wave frequency f =0.775Hz; (a) theory; (b)
inferred from laser sensors measurement of membrane displacement.
7.1
7.2

7.3

7.4
7.5
7.6

Experimental layout for the rubber duct equipped with the PTO.
Dashed lines correspond to different internal static pressure heads.
PTO air pressure versus PTO water level for the rubber duct with
an internal static pressure head of 0.02m; f =0.625Hz; (a) test 7;
(b) test 6; (c) test 4; (d) test 5. . . . . . . . . . . . . . . . . . . .
Phase lag between PTO water level and PTO air pressure as a function of wave frequency for the duct with an internal static pressure
head of 0.02m ; measured, from (4.25); (a) test 7; (b) test
6; (c) test 4; (d) test 5. . . . . . . . . . . . . . . . . . . . . . . . .
Ratio of the flow rate in the copper pipes to that of the air in the
chamber, for the duct with an internal static head of 0.02m. . . .
Measured values of the bulk modulus of air, with configuration as
in test 8. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Capture width (in duct widths) for a static pressure head of 3cm;
lines: computed from theory; symbols: measured; colour legend:
test 2 (Z = 1.10) in grey and test 1 (Z = 1.03) in black. . . . . . .

125

126

126

126

127

127
127
128

128

129

. 132

. 134

. 135
. 135
. 137

. 138

LIST OF FIGURES
7.7

7.8

7.9

7.10

7.11

7.12

7.13

7.14

7.15

7.16

Capture width (in duct widths) for a static pressure head of 2cm;
lines: computed from theory; symbols: measured; colour legend:
test 6 (Z = 1.08) in grey and test 7 (Z = 1.02) in black. . . . . . .
Pressure in the duct at thirty instants over one wave period with an
internal static pressure head of 2cm; wave frequency f =0.625Hz; (a)
theory for Z = 1.02; (b) inferred from laser sensors measurements
of membrane displacement. . . . . . . . . . . . . . . . . . . . . . .
Pressure in the duct at thirty instants over one wave period with an
internal static pressure head of 2cm; wave frequency f =0.625Hz; (a)
theory for Z = 1.08; (b) inferred from laser sensors measurements
of membrane displacement. . . . . . . . . . . . . . . . . . . . . . .
Pressure in the duct at thirty instants over one wave period with an
internal static pressure head of 3cm; wave frequency f =0.75Hz; (a)
theory for Z = 1.03; (b) inferred from laser sensors measurements
of membrane displacement. . . . . . . . . . . . . . . . . . . . . . .
Pressure in the duct at thirty instants over one wave period with an
internal static pressure head of 2cm; wave frequency f =1.05Hz; (a)
theory for Z = 1.08; (b) inferred from laser sensors measurements
of membrane displacement. . . . . . . . . . . . . . . . . . . . . . .
Measured total internal pressure, normalised by the pressure acting
on the membrane (a) and phase (b) for the duct with an internal static pressure head of 2cm, plotted as points. Lines represent
amplitudes and phases computed as the sum of the components obtained from the linear squares method. Wave frequency f =0.625Hz,
with conditions as in test 6 . . . . . . . . . . . . . . . . . . . . . .
Measured total internal pressure, normalised by the pressure acting
on the membrane (a) and phase (b) for the duct with an internal static pressure head of 2cm, plotted as points. Lines represent
amplitudes and phases computed as the sum of the components obtained from the linear squares method. Wave frequency f =1.05Hz,
with conditions as in test 6 . . . . . . . . . . . . . . . . . . . . . .
Pressure in the duct at thirty instants over one wave period with an
internal static pressure head of 2cm; wave frequency f =0.625Hz; (a)
inferred from laser sensor measurement of membrane displacement
amplitude; (b) sum of the three bulge wave components (and the
wave pressure) obtained for a least squares method. . . . . . . . .
Pressure in the duct at thirty instants over one wave period with an
internal static pressure head of 2cm; wave frequency f =1.05Hz; (a)
inferred from laser sensor measurement of membrane displacement
amplitude; (b) sum of the three bulge wave components (and the
wave pressure) obtained for a least squares method. . . . . . . . .
Magnitude of Pb+ relative to the wave pressure acting on the membrane with an internal static pressure head of 2cm and conditions as
in test 6; symbols: experimental values inferred from a least squares
method; : calculated from (5.34). . . . . . . . . . . . . . . . . .

xvii

. 139

. 140

. 141

. 142

. 142

. 144

. 144

. 145

. 145

. 146

xviii

LIST OF FIGURES

7.17 Magnitude of Pb relative to the wave pressure acting on the membrane with an internal static pressure head of 2cm and conditions as
in test 6; symbols: experimental values inferred from a least squares
method; : calculated from (5.35). . . . . . . . . . . . . . . . . .
7.18 Magnitude of PbA relative to the wave pressure acting on the membrane with an internal static pressure head of 2cm and conditions as
in test 6; symbols: experimental values inferred from a least squares
method; : calculated from (5.33). . . . . . . . . . . . . . . . . .
7.19 Capture width (in duct widths) for a static pressure head of 3cm;
lines: computed from theory; symbols: measured for the three runs
of test 3. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
7.20 Capture width (in duct widths) for a static pressure head of 2cm;
lines: computed from theory; symbols: measured for the run of test
4. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
7.21 Capture width (in duct widths) for a static pressure head of 2cm;
lines: computed from theory; symbols: measured for the run of test
5. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
7.22 Pressure in the duct at thirty instants over one wave period with an
internal static pressure head of 3cm; wave frequency f =0.625Hz; (a)
theory for Z = 1.93; (b) inferred from laser sensors measurements
of membrane displacement. . . . . . . . . . . . . . . . . . . . . . .
7.23 Measured total internal pressure, normalised by the pressure acting
on the membrane (a) and phase (b) for the duct with an internal static pressure head of 3cm, plotted as points. Lines represent
amplitudes and phases computed as the sum of the components obtained from the linear squares method. Wave frequency f =0.625Hz,
with conditions as in test 3. . . . . . . . . . . . . . . . . . . . . .
7.24 Pressure in the duct at thirty instants over one wave period with an
internal static pressure head of 3cm; wave frequency f =0.625Hz; (a)
inferred from laser sensor measurement of membrane displacement
amplitude; (b) sum of the three bulge wave components (and the
wave pressure) obtained for a least squares method. . . . . . . . .
7.25 Pressure (left column) in the PTO and amplitude (right column) of
the water-air interface displacement. Lines are the calculated values from the theory in section 5.5.2 and points represent measured
values for (a) test 1; (b) test 2; (c) test 3. . . . . . . . . . . . . . .
7.26 Pressure (left column) in the PTO and amplitude (right column) of
the water-air interface displacement. Lines are the calculated values from the theory in section 5.5.2 and points represent measured
values for (a) test 7; (b) test 6; (c) test 4 and (d) test 5. . . . . .
7.27 Wave analysis for the duct in with an internal static pressure head
of 3cm and configurations as in (a) test 1; (b) test 2; (c) test 3 and
(d) test 5. Left column: power carried by the first harmonic; right
column: power carried in the first three harmonics of the signals
from the wave gauges. . . . . . . . . . . . . . . . . . . . . . . . .

. 146

. 147

. 148

. 149

. 149

. 150

. 151

. 151

. 153

. 154

. 156

LIST OF FIGURES

xix

7.28 Wave analysis for the duct in with an internal static pressure head
of 2cm and configurations as in (a) test 7; (b) test 6; (c) test 4 and
(d) test 5. Left column: power carried by the first harmonic; right
column: power carried in the first three harmonics of the signals
from the wave gauges. . . . . . . . . . . . . . . . . . . . . . . . . . 157
8.1
8.2
8.3
8.4
8.5
8.6
8.7

8.8
8.9
8.10
8.11

8.12

8.13

8.14

8.15

8.16

Definition sketch of a closed duct with elastic membrane on its top.


Code developed for the analytical model. . . . . . . . . . . . . . . .
Reflection coefficient for h/d = 2 and s/d = 2. from Mei and
Black (1969), diffraction part of the present method. . . . . .
As for Figure 8.3, but s/d = 4. . . . . . . . . . . . . . . . . . . . . .
As for Figure 8.3, but s/d = 6. . . . . . . . . . . . . . . . . . . . . .
As for Figure 8.3, but h/d = 2.78 and s/d = 4.43. . . . . . . . . . .
Reflection coefficient for h/d = 2 and s/d = 2. diffraction only,
Tx for the following methods of conservation of volume:   
uniform potential; + + + least-squares; N N N matrix approximation; no correction. . . . . . . . . . . . . . . . . . . . . . . . .
As for Figure 8.7, but s/d = 4. . . . . . . . . . . . . . . . . . . . . .
As for Figure 8.7, but s/d = 6. . . . . . . . . . . . . . . . . . . . . .
As for Figure 8.7, but h/d = 2.78 and s/d = 4.43. . . . . . . . . . .
Symmetric (left column) and asymmetric (right column) of the
diffracted potentials (upper plots) and velocities (lower plots) at
the interface x = s; s = 2m, h/d = 2 with h = 0.4m and
Tx /gh2 = 0.15 and wave frequency f = 1Hz. . . . . . . . . . . . . .
Symmetric (left column) and asymmetric (right column) of the radiated potentials (upper plots) and velocities (lower plots) at the interface x = s; s = 2m, h/d = 2 with h = 0.4m and Tx /gh2 = 0.15
and wave frequency f = 1Hz. . . . . . . . . . . . . . . . . . . . . .
Combined free surface dynamic condition (8.6) for the symmetric
(left plot) and asymmetric (right plot) diffracted potentials in region
(1); s = 2m, h/d = 2 with h = 0.4m and Tx /gh2 = 0.15 and wave
frequency f = 1Hz. . . . . . . . . . . . . . . . . . . . . . . . . . . .
Combined free surface dynamic condition (8.6) for the symmetric
(left plot) and asymmetric (right plot) radiated potentials in region
(1); s = 2m, h/d = 2 with h = 0.4m and Tx /gh2 = 0.15 and wave
frequency f = 1Hz. . . . . . . . . . . . . . . . . . . . . . . . . . . .
Combined free surface dynamic condition (8.6) for the symmetric
(left plot) and asymmetric (right plot) diffracted potentials in region
(2); s = 2m, h/d = 2 with h = 0.4m and Tx /gh2 = 0.15 and wave
frequency f = 1Hz. . . . . . . . . . . . . . . . . . . . . . . . . . . .
Combined free surface dynamic condition (8.6) for the symmetric
(left plot) and asymmetric (right plot) radiated potentials in region
(2); s = 2m, h/d = 2 with h = 0.4m and Tx /gh2 = 0.15 and wave
frequency f = 1Hz. . . . . . . . . . . . . . . . . . . . . . . . . . . .

162
172
173
173
174
174

176
176
177
177

178

179

180

180

181

181

xx

LIST OF FIGURES
8.17 Dynamic condition (8.18) for the symmetric (left plot) and asymmetric (right plot) for the first modal (l = 1) radiated potential in
region (2); s = 2m, h/d = 2 with h = 0.4m and Tx /gh2 = 0.15
and wave frequency f = 1Hz. . . . . . . . . . . . . . . . . . . . .
8.18 Dynamic condition (8.18) for the symmetric (left plot) and asymmetric (right plot) for the first modal (l = 1) radiated potential in
region (3); s = 2m, h/d = 2 with h = 0.4m and Tx /gh2 = 0.15
and wave frequency f = 1Hz. . . . . . . . . . . . . . . . . . . . .
8.19 Kinematic condition (8.27) for the symmetric (left plot) and asymmetric (right plot) with no volume correction; s = 2m, h/d = 2
with h = 0.4m and Tx /gh2 = 0.15 and wave frequency f = 1Hz. .
8.20 Kinematic condition (8.27) for the symmetric (left plot) and asymmetric (right plot) with the WLS volume correction; s = 2m,
h/d = 2 with h = 0.4m and Tx /gh2 = 0.15 and wave frequency
f = 1Hz. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
8.21 Kinematic condition (8.27) for the symmetric (left plot) and asymmetric (right plot) with the MatApp volume correction ( = 1015 );
s = 2m, h/d = 2 with h = 0.4m and Tx /gh2 = 0.15 and wave frequency f = 1Hz. . . . . . . . . . . . . . . . . . . . . . . . . . . .
8.22 Comparison of the change of volume induced by the symmetric and
efficiency of the proposed methods; (a) no correction of volume, (b)
WLS method; (c) Unip method; (d) MatApp ( = 1015 ); s = 2m,
h/d = 2 with h = 0.4m and Tx /gh2 = 0.15. . . . . . . . . . . . .
8.23 Comparison of the error on the energy conservation in the system
for the proposed methods; (a) no correction of volume, (b) WLS
method; (c) Unip method; (d) MatApp ( = 1015 ); s = 2m,
h/d = 2 with h = 0.4m and Tx /gh2 = 0.15. . . . . . . . . . . . .
8.24 Ratio of the reflected power, Pr , to incident wave power Pi for the
duct with an internal static pressure head of 2cm. Line: present
model; symbols: first harmonic analysis of wave gauge signals. . .
8.25 Response of the membrane (/A) as a function of wave frequency
f and horizontal coordinate x, parameters such as in test 8. . . .
8.26 Modal response amplitude as function of wave frequency f and horizontal coordinate x; upper plot: first symmetric mode, middle plot:
second symmetric mode; lower plot: third symmetric mode. . . . .
8.27 As for Figure 8.26 for asymmetric modes. . . . . . . . . . . . . . .
8.28 Membrane displacement (normalised by the incident wave amplitude A) for configuration of test 8 and f =1.05Hz. . . . . . . . . .
8.29 Response of the membrane (/A) as a function of wave frequency
f and horizontal coordinate x, parameters such as in test 8, with
the longitudinal tension used 2.7 times greater than that due to
Poissons ratio effect only. . . . . . . . . . . . . . . . . . . . . . .
8.30 As for Figure 8.29 with a the tension multiplied by a factor of 11.

. 182

. 182

. 183

. 183

. 184

. 185

. 186

. 187
. 187

. 188
. 189
. 190

. 191
. 192

A.1 Power take-off set-up in the narrow flume. . . . . . . . . . . . . . . A2

LIST OF FIGURES

xxi

A.2 Air pressure in the chamber as a function of the water-air interface elevation for several dimensionless impedances Z. The wave
frequency is f = 0.95Hz and incident wave amplitude A =0.025m. . A3
A.3 Phase lag between PTO water level aPTO and PTO air pressure
pPTO as a function of dimensionless impedance Z for several wave
frequencies; measured, from (4.25). . . . . . . . . . . . . . A4
A.4 Capture widths for the duct equipped with the PTO and placed
in the narrow flume for various dimensionless impedances Z; Grey
line: calculated from (5.43), line and symbols: measured. . . . . . . A5
C.1 The duct, covered with the rubber 1mm membrane, in the wave
basin. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
C.2 Bow of the duct, with the connection for the water hose, in the wave
basin . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
C.3 Stern of the duct, with the connection to the bent tube, in the wave
basin. Heavy masses are visible and were used to hold the bottom
plate still. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
C.4 Power take-off system. . . . . . . . . . . . . . . . . . . . . . . . .
C.5 Ramp and bow of the duct, in the narrow flume. . . . . . . . . . .
C.6 The duct on the artificial seabed, covered with the 1mm membrane,
in the narrow flume. . . . . . . . . . . . . . . . . . . . . . . . . .
D.1 Sketch of the apparatus for testing rubber properties. . . . . . . .
D.2 Three different loads applied to the strip of rubber; (a) F = 0N,
(b) F = 78.86N, (c) F = 98.48N. . . . . . . . . . . . . . . . . . .
D.3 Stress-strain relationship for successive loadings with two minutes
intervals. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
D.4 Stress-strain relationship for successive loadings with thirty seconds
intervals. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
D.5 Volume of the rubber strip for the different stresses applied. . . .
D.6 Damped free oscillations of a rubber strip. . . . . . . . . . . . . .

. C1
. C2

. C2
. C3
. C3
. C4
. D1
. D2
. D3
. D3
. D4
. D5

List of Tables
2.1

Comparison of wave power calculations for West Orkney site, after


Folley and Whittaker (2009). . . . . . . . . . . . . . . . . . . . . . . 23

3.1
3.2

Average wave coefficients of flexible models, after Stamos et al. (2003). 53


Average wave coefficients of rigid models, after Stamos et al. (2003). 54

4.1

Measured characteristics of test rubber . . . . . . . . . . . . . . . . 65

6.1
6.2

Closed ends experiment: test conditions . . . . . . . . . . . . . . . 94


Comparison of experimental resonant frequencies fres with the theoretical ones calculated as fn = nc/2L where c is the theoretical free
bulge speed for the corresponding internal static pressure heads. . . 98
Experimental conditions for the closed ends experiment with the
0.2mm membrane. . . . . . . . . . . . . . . . . . . . . . . . . . . . 121

6.3
7.1

PTO experiment: test conditions . . . . . . . . . . . . . . . . . . . 133

A.1 PTO experiment in narrow flume: test conditions . . . . . . . . . . A2

xxiii

Declaration of Authorship
I, Nicolas Choplain, declare that the thesis entitled Interactions of a submerged
membrane with water waves and its use in harnessing nearshore wave power and
the work presented in the thesis are both my own, and have been generated by
me as the result of my own original research. I confirm that:
this work was done wholly or mainly while in candidature for a research
degree at this University;

where any part of this thesis has previously been submitted for a degree or
any other qualification at this University or any other institution, this has
been clearly stated;

where I have consulted the published work of others, this is always clearly
attributed;

where I have quoted from the work of others, the source is always given.
With the exception of such quotations, this thesis is entirely my own work;

I have acknowledged all main sources of help;


where the thesis is based on work done by myself jointly with others, I have
made clear exactly what was done by others and what I have contributed
myself;

Signed: . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Date

xxv

Acknowledgements
The work presented in this thesis would not have been possible without the support
of many people that ought to be given credits here. First of all, I would like to
thank Professor John R. Chaplin for giving me the opportunity to work on this
project, guiding me trough it and bearing with my many questions. Most of
this work being experimental, I owe a huge debt of gratitude to the laboratory
staff who helped me, one way or the other, during this study: Karl Scammell,
Rhys Jenkins, Harvey Skinner, Mike Rose, Dave Lynnock and Earl Peters. I also
benefited from interesting discussions or practical help from the Anaconda team
and I wish to thank its members for that: Francis Farley, Grant Hearn, Valentin
Heller, Steve Rimmer and Martyn Hann. I wish to thank my housemates for
making the different houses I lived in more than just a house: Patrick Wiemann,
John Sandoval, Diana Caicedo, Martha Climenhaga and Ramesh Satkurunath.
The friendship that is born from this sharing is truly valuable. I feel lucky for
meeting many people during my stay in the UK and I especially want to thank
a couple of them for different reasons: Salome Giao, Polyvios Polyviou, Tassos
Papafragkou, Nicholaos Bakker, Davide Magagna, Jack Giles, Pascal Galloway,
Mark Leybourne, Ghayth Abed, Peter Smith, Konstantinos Nestoridis and the
football mates. Back home, I wish to thank my pillar friends: Benjamin Msika,
Pascal Jacq, Flavien Gouillon, Adrien Biggi, Jerome Boisard, Sylvain Crequet,
Aurelien Collombet and Lauriane Pautot-Lemaire. Thanks for giving sense to the
word friendship. A special and sincere consideration goes to Loic and Aurelie
Jullion. Lastly, I would like to express my gratitude to my family who always
supported me through those years of study, and not only PhD. Life makes that
we do not see as much as we would like to, but their thoughts and gestures always
warm my heart up. My mother and father, for leaving me the choice to do what
I like and support me all the way long, mentally, emotionnaly, physically and
financially. Having such wonderful parents is a gift that I am more than lucky to
have received. Finally, Konstantina, with her constant love, support and positive
attitude made me believe it could happen. Thank you all!
xxvii

Nomenclature
Variables of importance are detailed in the table below. Other notations are used
in this thesis for mathematical simplification and are detailed when in use.
a

Half-width of the duct (m)

ai

Equivalent wave amplitude (m)

Wave amplitude (m)

Ai

Incident wave amplitude (m)

Ar

Reflected wave amplitude (m)

Depth of the duct (m)

Free bulge speed (m)

cg

Group velocity (m/s)

cp

Phase speed (m/s)

chys

Free bulge speed with hysteresis in the rubber (m/s)

Membrane submergence (m)

d1

Equivalent submergence depth (m)

Distensibility (Pa1 )

Displacement of the water-air interface in the power take-off (m)

E
E

Youngs modulus (Pa)

Ep

Potential energy of the membrane (J)

Ek

Kinetic energy of the membrane (J)

Wave frequency (Hz)

Wave energy (J/m2 )

fres

Resonant frequency of the fluid-membrane system (Hz)

fslug

Resonant frequency of the slug of water (Hz)

Pressure amplification factor

Acceleration due to gravity (m/s2 )

Water depth (m)

Wave height (m)

Imaginary unit; index of summation

Index of summation

Wave number (m1 ); index of summation

k1

Water wave number (m1 )

k2

Bulge wave number (m1 )

Kl

Lost wave energy coefficient

Kr

Reflected wave energy coefficient

Kt

Transmitted wave energy coefficient

Length of the slug of water in the rigid tube (m)

Length of the membrane (m)

Lp

Length of a copper pipe (m)

mass per unit length of the membrane (kg/m); index of summation

mt

sum of the mass of the membrane per unit area and its added mass (kg/m2 )

Number of natural modes for the membrane displacement

Index of summation

Number of eigenvalues in the expansion series

Pressure (Pa)

p0

Internal static pressure head (m)

pD

Wave dynamic pressure (Pa)

pa

Atmospheric pressure (Pa)

pb

Bulge pressure (Pa)

pw

Wave pressure (Pa)

Phys
Pi
Ploss

Power lost in the rubber due to hysteresis (W)


Incident wave power (W)
Lost wave power (W)

Pf

Mean power absorbed by the power take-off


from flow rate measurement only (W)

Pr

Reflected wave power (W)

Pt

Transmitted wave power (W)

Flow rate in the copper pipes (m3 s1 )

Radius of a stainless steel pipe in the power take-off (m)

Gas constant (J/kg K)

R1

Resistance level for characteristic resistor 1 ()

R2

Resistance level for characteristic resistor 2 ()

R3

Resistance level for characteristic resistor 3 ()

R4

Resistance level for characteristic resistor 4 ()

xxx

Rm

Radius of curvature of the membrane deflection (m)

Half length of the membrane (m)

Instantaneous cross-section of the rubber tube (m2 )

S0

Undisturbed cross-section of the rubber tube (m2 )

Sd

Cross-section of the duct (m2 )

St

Cross-section of the rigid tube (m2 )

Time (s)

Wave period (s)

Tu

Circular tension in the rubber (N/m)

Tx

Tension in the rubber in the x-direction (N/m)

Ty

Tension in the rubber in the y-direction (N/m)

Tz

Tension in the rubber in the z-direction (N/m)

Fluid particle velocity (m/s)

ux

Fluid particle velocity in the x-direction (m/s)

uy

Fluid particle velocity in the y-direction(m/s)

uz

Fluid particle velocity in the z-direction(m/s)

Instantaneous rubber thickness (m)

v0

Initial rubber thickness (m)

Instantaneous volume of air in the PTO (m3 )

VA

Voltage level for characteristic point A (V)

VB

Voltage level for characteristic point B (V)

Membrane centreline deflection (m)

Position along the x-axis (m) (wave propagation direction)

Position along the y-axis (m)

Position along the z-axis (m) (perpendicular to still water surface)

Dimensionless impedance

ZPTO

Impedance of the power take-off (m3 s1 )

Zd

Impedance of the duct (m3 s1 )

Z1

Impedance pf one copper pipe (m3 s1 )

Lagrangian (J)

Hysteresis coefficient (s)

Loss angle

Strain

x

Instantaneous strain rubber in the x-direction

y

Instantaneous strain in the rubber in the y-direction

x,0

Initial strain in the rubber in the x-direction

xxxi

y,0

Initial strain in the rubber in the y-direction

Water surface displacement (m)

Angle of wave incidence ( )

Wavelength (m)

Poissons ratio

Ratio of the duct cross-section to that of the bent tube

air

Dynamic viscosity of air (Pa.s)

Mass per unit area of the membrane (kg/m2 )

Normalised (by the wave amplitude) membrane displacement

Density of water (kg/m3 )

Density of rubber (kg/m3 )

Stress (Pa)

Instantaneous stress in the rubber in the x-direction (Pa)

Instantaneous stress in the rubber in the y-direction (Pa)

x,0

Initial stress in the rubber in the x-direction (Pa)

y,0

Initial stress in the rubber in the y-direction (Pa)

Ratio of the flow rate in the copper pipes to that of the air in the chamber

Flow potential (m2 s1 )

Phase of incident wave

Phase of reflected wave

eq

Phase lag between pressure and displacement in the power take-off

Angular frequency (s1 )

xxxii

To my parents and family


A mes parents et ma famille

xxxiii

Chapter 1
Introduction
1.1

Motivation
I have a great many reasons why I think wave power can be put
to practical and profitable use, but I dont wish to occupy too much of
your valuable time, or, perhaps, there are not many of you particularly
interested in the subject

This statement made within the discussion of an article by Stahl (1892) seems obsolete. Nowadays, almost no one would disagree that the power contained in the
oceans, in many forms and particular in waves, could contribute to the ongoing
change in electricity production from fossil fuels to greener ways.
This has led to the development of various theories and corresponding devices
to harness this power over the last decades. Almost all of these structures were
solid, yielding high costs of operation and maintenance, and hampering the successful wave energy conversion process.
Because they are made of rubber (and other components) and not solid material,
distensible devices are thought to contribute to the development of more reliable
systems. Rubber is cheap, environmental friendly and can adapt to even rough
sea states. One of the promising devices is the Anaconda, a rubber tube placed
beneath the surface and aligned head to waves. It is filled with water and its resonant conditions are used to maximise the magnitude of the pressure propagating
in it to extract the corresponding power. It is aimed to be deployed in deep waters
where wave power is expected to be of higher order of magnitude. As for any other
1

Chapter 1 Introduction

floating or submerged (but not on the seabed) structure, moorings can not only be
hazardous but also add to the cost of operation and maintenance in case of failure.
The goal of the current work was to investigate the applicability of the theory
developed for the Anaconda on a bottom-mounted version and provide the necessary information on a potential commercial development. The appropriate model
was a long rectangular duct, with solid walls, aligned head to waves, resting on
the seabed and covered with a rubber membrane. Two configurations were tested,
one with closed ends and one with a power take-off system connected at its stern.
The work presented in this thesis was driven by the desire to compare theoretical results from the one dimensional model of pressure propagation in distensible
tubes, such as the one in the Anaconda, to the results of a series of experiments
carried on both versions of the duct. Values of pressure propagation in the duct,
membrane displacement and power captured by the power take-off (for the configuration with it) are then presented. Discrepancies between the above theory and
experimental data for one of the tested membrane are explained and a discussion
follows on the limitation of using this one dimensional model for specific configurations.
The main assumption in the latter model is the absence of interactions between
the membrane and the wave propagating over it. In order to further study this
hypothesis, a two dimensional analytical model was developed. It is based on an
eigenfunction expansion method, and deals with the interactions of the studied
geometry where the membrane is stretched in the wave propagation direction. Its
results are applicable to the present work, and also to a wide range of fields, especially for breakwater studies.
A novel technique for the measurement of the membrane displacement (hence
the pressure in the duct) beneath waves, based on a set of laser sensors, placed in
waterproof boxes just beneath the water surface is also presented. Results from
this technique fill the gap of the current available experimental characterisation of
such fluid-filled system behaviour in waves, especially at resonance.

Chapter 1 Introduction

1.2

Thesis layout

The work carried is presented in this thesis according to the following layout:
Chapter 2 supplies the reader with the basic concepts behind wave energy
conversion. The two main reasons on why this work is relevant are given,
together with the precise definition of the aims and objectives for the present
study.
Chapter 3 is devoted to the literature review of previous studies dealing

with similar device. A technical background is also provided, and areas


where the present work can address gaps or improvement in the knowledge
are emphasised.

Chapter 4 details the experimental facilities, techniques and instrumentation


used to fulfil the experimental part of the present investigation.
Chapter 5 describes the theory for the free bulge speed calculation, pressure
in the duct for the closed ends and PTO-equipped versions, based on a
one dimensional model for wave propagation in compliant and liquid-filled
structures.
Chapter 6 presents the experiments carried on the closed ends version of the

duct in a narrow flume and compares the values of membrane deflection, free
bulge speed in the membrane and pressure in the duct with those from the
theory from chapter 5. Discrepancies between results from this model and
that of a series of experiments carried out with the thinner of the two tested
membranes are explained.

Chapter 7 presents the experiment carried on the duct equipped with the
power take-off connected at its stern, and starts with an analysis of the power
take-off system and its reliability as a linear dashpot. This done, power data
are presented for various configurations and there follows a similar study of
pressure characterisation as in chapter 6.
Chapter 8 presents the two dimensional model developed within the course
of this study. It concerns the interactions of water waves with a bottommounted duct, aligned head to waves and covered with a stretched membrane. The challenging part of the fluid volume conservation is detailed and
solutions proposed with the aim of obtaining a best compromise solution.

Chapter 1 Introduction
Outputs from this model are compared as far as possible with experimental
data from the wide flume experiments.
Chapter 9 closes this thesis by summarising the work done and its potential
contribution to science, and outlines the future work that could be carried.

Chapter 2
Wave energy glossary and
relevance of the study
2.1

Introduction

The author has continuously noticed during this work that the idea of harnessing
wave power receives an enthusiastic welcome from individuals not from this field.
This interest was either based on their environmental convictions, their technological interests or simply by curiosity. However, the author also noticed that the first
reaction was that people were associating wave energy with the energetic impact
of a breaking plunging wave seen during their beach holidays.
This is why the present chapter is here to give a brief background on wave characteristics and wave energy conversion. From that point, the justification of the
study follows, based on the consideration of the Anaconda wave energy converter
(WEC) and the new trend in capturing nearshore wave power. The aims and
objectives of the present work closes up this chapter.

2.2
2.2.1

Wave characteristics
Introduction

Due to the differential heating of the atmosphere by the sun, wind motion is
triggered in the atmosphere. When wind blows over oceanic area, forces (pressure
5

Chapter 2 Wave energy glossary and relevance of the study

and friction) disturbs the water surface and gravity tries to restore this disturbance
resulting in oscillations about the equilibrium state. Waves are created and are
the result of a transfer of energy from wind to the surface. For a wavelength more
than a couple of centimetres, ocean waves are categorised as gravity waves.

2.2.2

Definitions

A simple representation of a monochromatic (having a single wavelength and frequency) gravity wave is given in Figure 2.1. It is characterised by:
. its frequency f and angular frequency = 2f .
. its amplitude A and height H = 2A.
. its wavelength that is the distance between two wave crests (or troughs)
. its wavenumber k = 2/ with
= kcp ,

(2.1)

where cp is called the phase speed and represents the speed of a crest (or
a trough).
Wave period is defined as
1
f

=
.
cp

T =

(2.2)
(2.3)

The linear wave theory is used in this study. Linear waves have amplitudes much
smaller than the water depth they propagate in. This criteria is characterised as
Ak << 1 (orA/h << 1).
In that case, the dispersion relationship links the wavenumber and the frequency
as
2 = gk tanh(kh).

(2.4)

Chapter 2 Wave energy glossary and relevance of the study

Direction of
propagation

crest
A

j
O i

dz

trough

Figure 2.1: Sketch of a water wave

from which k can be obtained and yielding the phase speed


cp =

g tanh kh
.
k

(2.5)

Finally the surface displacement and potential of the progressive monochromatic wave shown in Figure 2.1 is defined as
(x, t) = A cos(kx t).

2.2.3

(2.6)

Kinetic and potential energies

Waves encompass two forms of energy: kinetic energy associated with the water
particles motion and potential energy as a result of a displaced mass of water
from its equilibrium position in a gravitational field.
The total wave energy per unit area is usually defined as
= g H 2 .
E
8

(2.7)

However for wave power absorption purpose, it is convenient to know how much
energy a device can absorb per unit time as shown in Figure 2.2. With Lc taken
as a unit length, Dean and Dalrymple (1991) defines this available energy as the
rate at which work is done by the dynamic pressure pD on a vertical section in
direction of wave propagation.

Chapter 2 Wave energy glossary and relevance of the study

Incident
waves

Area of energy
absorbed during

Lc

the time T

cgT
Figure 2.2: Energy absorbed per unit time T

This instantaneous power is


I=

pD ux dz

(2.8)

with
pD = p + gz

gz),
= (
t

(2.9)
(2.10)

where is the flow potential and ux = /x the fluid particle velocity. Averaging
this instantaneous work over one wave period T and for the monochromatic wave
in Figure 2.1 gives the average energy flux
1
1
2kh
I = gA2
(1 +
)
2
k
2
sinh
2kh
{z
}
| {z } |{z} |

cp

(2.11)

{z
cg

where cg is called the group velocity and is defined as the speed at which energy
is transported.

Chapter 2 Wave energy glossary and relevance of the study

A dimensional analysis of I gives


g ].1
[I] = [E].[c

(2.12)

= M.T 2 .L.T 1

(2.13)

= M.L.T 3

(2.14)

2
= |M.L{z
.T 3} .L1
power

(2.15)

showing that the energy flux has the dimension of a power per unit length. The
definition in (2.11) will then be used to characterise the mean water wave power
per unit crest length and I will therefore be denoted P .
The main advantage of using a definition of power relative to a unit crest length
is that it gives information on the width of crest from which a device will harness
wave power. This width is called capture width (CW) and is defined as
CW =

2.3
2.3.1

mean power absorbed by a WEC


mean power per unit crest length

(2.16)

Wave power background


Introduction

Waves carry energy as it was theoretically explained in the previous section. However, real sea states are far from the ideal monochromatic case shown in Figure
2.1. A typical wave elevation is due to waves of different amplitudes, frequencies
and directions. Offshore industry and wave energy developers require mapping of
the wave field in terms of wave characteristics, namely wave height and frequency.
It is quite straightforward with a monochromatic wave but not with the sea state.
However, several techniques exist for that purpose, based on the type of representation of the sea-state (stochastic or deterministic) and the corresponding data
analysis of in-situ measurement. Wave models are also continuously developed
and improved either for large scale representation, or specific target site study.

10

Chapter 2 Wave energy glossary and relevance of the study

Figure 2.3: Global distribution of annual mean wave power, after Cornett
(2008)

2.3.2

Wave power resource distribution

The theoretical global resource for wave energy is estimated to be 8, 000-80, 000
TWh/year, corresponding to a power of 0.91 9.13 TW (Boud (2003)). It is usu-

ally estimated that the exploitable part of this power is in about 10% for practical
and economical reasons. To give an order of magnitude for comparison, the an-

nual world electricity consumption is around 16, 000 TWh/year corresponding to


a mean power of 1.826TW: so if all this portion of wave power were converted into
electricity, it could contribute up to 50% to the world electricity consumption.
This figure will be substantially lower, depending on technical developments of
wave energy converters and this is why wave energy should be seen as a contributor together with any other forms of renewable energies, marine or not, rather
than the only solution.
Wave climate, by its origin, is closely related to wind patterns. This is why
the best wave climates (for wave energy purpose) are found between the latitudes
of 30o and 60o , where the prevailing westerlies blow over a long oceanic
distance gifting the west coasts of Europe, North America, Australia and South
Africa (Figure 2.3). With high wave energy resource the wave climate in Europe
is particularly energetic: it has been estimated than the wave power resource for
the north-eastern Atlantic (including the North Sea) is about 320GW (Mrk et al.
(2010)).

11

Chapter 2 Wave energy glossary and relevance of the study

a)

Airflow

b)

Turbine
Column

Basin water level

Water level

Overtopping
device

OWC

Turbine
Water
flow

c)

Attenuator
PA
OWSC

PA

High
pressure

Low
pressure

Submerged pressure differential


Figure 2.4: Wave energy converters classification, from Rahm (2010)

2.3.3

Harnessing concepts of wave power

2.3.3.1

Introduction

Although over 1,000 patents exist today regarding ways of extracting wave power,
a generic classification can be used to categorise the different concepts used for
that purpose. Differentiation between devices can be either based on intended
locations of use or modes of operations. The latter is used here to give the main
concept on wave energy conversion techniques and the differentiation criteria used
are the ones developed in Falcao (2010) and shown in Figure 2.4 as a) oscillating
water column, b) overtopping devices and c) oscillating bodies.

2.3.3.2

Oscillating water column

An oscillating water column is made of a solid structure enclosing a slug of water


whose heave motion is driven by the incoming wave pressure. Above it, a turbine rotates due to the reciprocating flow associated with this displacement and
generates electricity.

12

Chapter 2 Wave energy glossary and relevance of the study

2.3.3.3

Overtopping devices

Another way to extract wave energy is to apply the principles of hydropower to


wave energy conversion. Water wheels have always used the energy of flowing
or falling water to power any kind of work, as for milling flour in gristmills for
instance. Overtopping devices use this concept: they collect water of incident
waves in a elevated reservoir to drive low head turbine(s).

2.3.3.4

Oscillating bodies

Oscillating bodies represent any other devices that do not fit into the two previous
categories. As suggested by its name and unlike the two other types, an oscillating
device is in motion and wave-driven. It is this motion, mostly relative to other
component of the structure, or an immobile reference, that is used to extract wave
power.
This category can be sub-categorised into different categories regarding device
geometry and/or mode of motion. Looking at the oscillating bodies on Figure 2.4
from left to right, this sub-categorisation is:
Oscillating Wave Surge Converters (OWSC). In deep water, as waves

pass over the surface, the particles of water do not move forward with the
waves: they gyrate in circles. As the depth increases, the radius of the circular motion decreases. At a depth equal or greater than half the wavelength,
they are quite still. On the contrary, in shallow waters, the particle trajectories are compressed into ellipses due to the interaction between waves and
bottom. This leads to a back and forth motion of the water particles (bottom waves). OWSC are driven by the surging action of waves, resulting in a
pitching motion of their constitutive plate relatively to the bottom. This is

also why they are set in shallow waters. They will be more detailed in section
2.4.3.3 as they are aimed to harness nearshore power as in the present work.
Next two devices are point-absorbers (PA). The characteristic length of
such device is small compared to the wavelength. First device designated PA
converts the motion of the waves into rotational motion, causing an internal
fluid to run unidirectionally through a specially designed pump to generate
electricity. The second PA, and more widely used, consists of a floating buoy
whose relative displacement due to wave motion to a static reference is used

Chapter 2 Wave energy glossary and relevance of the study

13

to drive a hydraulic pump. The advantage of the latter is that they are
not oriented into a particular direction so can absorb wave energy from all
directions of propagation.
The penultimate device of Figure 2.4 is an attenuator. Attenuators are long
structures placed parallel to wave propagation. They are usually composed
of several sections that each follow the wave profile. The whole structure
length being of same order of magnitude or greater than the wavelength
these sections move relatively to each other. Connecting these sections with
hydraulic rams allows the conversion of this mechanical movement into electricity using hydraulic generators.
The last illustrated device is a submerged pressure differential, using
the pressure variation below travelling waves. Low pressure (wave trough)
lifts the movable part of the device and high pressure (wave crest) pushes it
back, compressing the enclosed air in the chamber.

2.3.4

Conclusions

Through the years, many WECs have been fashioned by scientists, engineers or
inventors, using the concepts detailed in section 2.3.3. Yet, among the different
forms of marine renewable energies, wave energy seems late compared to offshore
wind energy for instance. The reasons are because of the rough and unpredictable
environment this field is dealing with and not because techniques and knowledge
are not mastered. However, because of the recent sea tests for various prototypes,
together with the cost reduction for material used and installation processes, there
are reasons to be optimistic for wave energy farms in the long-term.
Because of the vast majority of devices already existing, it is the belief of the
author that the study, development and experimentation of a new device should
be explained. Therefore, this section is to detail the reasons why this study is
relevant. This is achieved by considering the Anaconda project and the recent
developments of devices capturing nearshore wave power.

14

2.4
2.4.1

Chapter 2 Wave energy glossary and relevance of the study

Relevance of the study


Introduction

This section is to present the two main arguments justifying the present work: a
promising device, the Anaconda, that is aimed to harness offshore wave power and
the reconsideration of capturing nearshore wave power.

2.4.2

The Anaconda WEC

2.4.2.1

Principle

The Anaconda is a new way of extracting energy from ocean waves, based on
bulge waves travelling along a distensible rubber tube (Farley and Rainey (2006);
Chaplin et al. (2007a,b, 2012)) (see Figure 2.5), filled with water, oriented in the
direction of wave travel and anchored head to waves (attenuator). Oscillations in
pressure beneath the waves generate travelling bulges in the tube. The bulge waves
grow in the down-wave direction, converting wave energy into internal oscillatory
flow that can be used to drive a turbine. One end of the tube is closed (the
bow) and the other one is linked to a power take-off (PTO). Tuning is achieved by
matching the speed of free bulge waves (a function of the tubes material properties
and geometry) to the phase speed of water waves.

Figure 2.5: Anaconda WEC, after Checkmate SeaEnergy Ltd.

15

Chapter 2 Wave energy glossary and relevance of the study

The main interest of this section on the Anaconda is not to detail all the theoretical
and experimental work that has been carried out so far by the Anaconda project
team. A set of interesting papers on the subject (Farley and Rainey (2006); Chaplin et al. (2007a,b)) has been written for that purpose. However, the aim here is
to briefly describe the theory behind it and to present some results agreeing with
the promising potential of this device.

2.4.2.2

The bulge wave theory

Bulges like the one shown in Figure 2.6 are waves propagating into the rubber
tube in the same way blood flows in arteries. They are driven by the change of
pressure difference across the tube (or channel) walls. The general assumption
is that this pressure difference is negligible over the cross-section and hence, the
wave propagation in the tube (or channel) is driven by the gradient of pressure
in the longitudinal direction (x with notations from Figure 2.6). Hence, all the
kinetic energy of the fluid is due to this longitudinal motion and the fluid particle
velocity in the channel is reduced to
u = ux i,

(2.17)

where ux will therefore be noted u for clarity and with i the unit vector in the
x-direction.
The pressure acting on the compliant wall is noted pw and the pressure inside
the duct p = pb + pw . For the situation where no other forces are involved in the
fluid motion and neglecting the non-linear terms of the momentum equation,

u
p
= .
t
x

(2.18)

On the other hand, assuming the flow to be uniform over the cross-section, the
linearised form of the continuity equation

S
+
(Su) = 0
t
x
leads to

u
S
= S0 .
t
x

(2.19)

(2.20)

16

Chapter 2 Wave energy glossary and relevance of the study

Water wave

Wave front
pw(x,t)

propagation direction
pb(x,t)+pw(x,t)

Bulge wave

Figure 2.6: Bulge wave propagating in the rubber tube.

Differentiating (2.20) with respect to t and using the definition of the distensibility
(with units of Pa1 )
D=

1 dS
S0 dpb

(2.21)

leads to the bulge wave equation obtained by Farley and Rainey (2006)
2 pb
1 2
=
{pb + pw }.
t2
D x2

(2.22)

The distensibility represents the ability of the tube to respond to changes of internal pressure with a proportional change of its cross-sectional area from its undisturbed value S0 .
This wave equation assumes no losses in the fluid or the rubber. Lighthill (1978)
detailed the wave attenuation by friction along the walls of the channels due to
viscous tangential stresses. This energy dissipation occurs in a boundary layer
(Stokes boundary layer) which Lighthill (1978) gives an approximate thickness a
a function of the fluid kinematic viscosity and frequency of fluid motion . Given
the dimensions of the problem, and especially the width of the duct, this dissipation is assumed negligible, on the contrary to what can happen when dealing with
smaller geometries, such as blood propagation into arteries.

17

Chapter 2 Wave energy glossary and relevance of the study

More important are the losses in the rubber wall by hysteresis, a notion described
3
later in section 4.4.3. Those losses are represented by the extra term tp3b added
to the left hand side of (2.22) such as
2 pb
3 pb
1 2

=
{pb + pw },
t2
t3
D x2

(2.23)

where the loss factor being defined by the stress-strain relationship


= E( + ),

(2.24)

with E the Youngs modulus of the rubber and = tan1 () the loss angle due
to hysteresis in the rubber.
The bulge pressure pb can be written, according to Farley and Rainey (2006),
pb = gAF,

(2.25)

with A the wave amplitude and F the pressure amplification factor that gives the
oscillating pressure inside the tube compared to the oscillating pressure in the sea
wave outside. In the absence of hysteresis, the pressure amplification factor is
k
k2
1
F = 1 sin(x
) k1 x,
2
2
kk

(2.26)

which is valid for small k where


k = k1 k2 ,
k1 + k2
,
k =
2

(2.27)
(2.28)

with k1 = c1 is the wave number of the external wave and k2 is the bulge natural
wave number associated with the natural bulge wave speed first introduced by
Lighthill (1978) defined as

c=

1
.
D

(2.29)

The free bulge speed is affected by the hysteresis and the free bulge speed chys
with the hysteresis taken into account was defined by Chaplin et al. (2012) as
chys = q

c
1
(1
2

+ sec )

(2.30)

18

Chapter 2 Wave energy glossary and relevance of the study

The measured values of the loss angle for the membranes tested in this work
( 2.69 ) do not result in significant modification of the free bulge speed due to
hysteresis. Hence this effect on the free bulge speed will be neglected.
It can be seen from (2.26) that the bulge amplitude grows linearly along the tube.
F is a sin(X)/X for variable k so that the maximum amplification factor occurs
for k = 0 when the natural bulge wave speed matches the external wave speed.
In this condition, the magnitude of oscillation of the bulge pressure is several times
stronger than the magnitude of the incoming wave amplitude. Bulges then carry
power extracted from the water wave above and the objective is to develop and
use a PTO that absorbs a maximum proportion of this power.

2.4.2.3

Current developments

The pressure amplification along the tube occurs when the free bulge speed matches
the above water wave speed. The idea recently developed by Chaplin et al. (2012)
is to use a bent slug of water at the end of the tube, whose natural frequency is
the same as the frequency at which the resonance in the tube occurs, hereinafter
called the resonant water wave frequency. This way, the horizontal motion of the
bulges at the stern of the rubber tube is completely converted into vertical motion
of the slug of water without any reflection. The power take-off is then an air power
take-off similar to what is found in an OWC. The water-air interface at the top of
the slug pushes back and forth air through a turbine.
In their work, Chaplin et al. (2012) model the air PTO as a linear dashpot,
meaning that the air pressure (force acting on the PTO) and the air-water interface velocity are in phase. This is commonly used when modelling air PTO
in wave energy converters, also because turbines such as the Wells turbine are
aimed and designed to be linear. A more complete description on the linear dashpot used is detailed in section 4.5 as the same concept is used for the present work.
The coefficient of proportionality between the air pressure and the air flow in
a linear dashpot is called the impedance. Figure 2.7 shows capture widths for
the PTO described above, placed at the stern of the 7m long and 250mm diameter rubber tube, and for several dimensionless impedances Z (impedance of
the PTO/impedance of the tube) ratio. It can be seen that the maximum capture
width, of order of 1.75 tube diameter, is obtained for a ratio of Z=1.05. This means
that the maximum power absorption is obtained when the tubes impedance (the

Chapter 2 Wave energy glossary and relevance of the study

19

ratio pressure/flow, e.g. c/S0 ) matches the PTO impedance and occurs when
the wave period T equal the period the resonant water wave period T0 . Capture
widths results agree with the 1D theory that is explained later in chapter 5.

Figure 2.7: Capture widths as functions of relative wave periods for (a)
Z=0.64, (b) Z=1.05, (c) Z=3.3, (d) Z=7.7. Measurements are shown as points,
one-dimensional theory with = 13 as continuous lines, after Chaplin et al.
(2012)

2.4.2.4

Conclusions

For any WEC project to be seen as feasible, the different challenges shown in
Figure 2.8 must be satisfied. The Anaconda is a distensible tube made of rubber
whose theory has been introduced in section 2.4.2.2 with a 1D model. This theory

20

Chapter 2 Wave energy glossary and relevance of the study

has been experimentally demonstrated at a laboratory scale of 1:25 as detailed in


Chaplin et al. (2012) (Predictability). The main advantage of the Anaconda lies
in the fact that it is made of rubber. Rubber can be considered:
1. Cheap compared to the materials used for other mechanical devices (Affordability).
2. A natural compound that does not corrode in saline environment (Survivability, Reliability and Operability).
3. A light material, making towing and anchoring processes easier(Installability).

Predictability
Affordability

Reliability

Manufacturability

Challenges

Survivability

Installability

Operability

Figure 2.8: Challenges facing the development of a new wave energy converter

The industrial concern at this point is the Manufacturability challenge. At full


scale, the Anaconda is aimed to be about 200 m long and 7 m in diameter, making it the largest tube of rubber ever designed. Such considerations are still in
development.
Located just beneath water surface, the Anaconda can however be hazardous
for ship passage or marine activities. Moreover, it is aimed to be anchored in
about 40-50 m depth of water. Mooring cables can represent a hazard for marine
life, especially whales migration. Main whales migration routes, located on the
west coast of North America, at the southern tip of South Africa and in Norway
are seen to be interesting spots regarding wave power resource as seen in Figure 2.3.

Chapter 2 Wave energy glossary and relevance of the study

21

This is why the author aims to find out whether the bulge wave concept is applicable on a seabed version, in order to avoid these issues. The bulges being created
by the dynamic wave pressure, the idea is to use it at a location where the seabed
is sensitive to the incoming wave pressure, hence in the nearshore zone. The next
section is to detail why this zone is of interest for wave power industry and the
current developments in term of bottom-mounted WEC.

2.4.3

Nearshore wave power

2.4.3.1

Introduction

Offshore locations have always attracted more interest from wave energy developers
due their highest gross energy resource. Onshore sites, mainly for OWC development, have also been privileged for their easy access (operation and maintenance)
and the fact that no underwater cable is needed to carry back the electricity to
the mainland.
However, nearshore locations (usually defined as water depths of 10-25 m) have
often been of less interest for WEC developers. The main reason is that nearshore
wave climate have been evaluated using the definition of gross energy resource.
This section is to detail the reasons why this definition is not adequate for these
locations where the bathymetry and wave directionality have to be taken into
account.

2.4.3.2

Exploitable wave power

Folley and Whittaker (2009) have recently shown that the nearshore wave climate might be more useful than previously thought. Using the SWAN model,
the nearshore wave climate was predicted using the offshore wave climate as an
input in the model. As an example, they used a wave of 10 second period propagating from offshore location (50 metre depth) to nearshore location, and using
different seabed slopes they showed that the wave power loss from one site to another is relatively small (reproduced in Figure 2.9) as long as the water depth is not
smaller than 8 metre. After that the depth induces wave breaking and energy loss.

22

Chapter 2 Wave energy glossary and relevance of the study

Figure 2.9: Shoaling of a 10 second energy period wave propagating orthogonal


to depth contours for different seabed slopes, after Folley and Whittaker (2009).

Figure 2.10: Variation in percentage power loss from offshore to nearshore


site for an initial sea-state of T =10s and Pi =50 kW/m travelling at an angle
to the depth contours on a seabed slope of 1:100, after Folley and Whittaker
(2009).

23

Chapter 2 Wave energy glossary and relevance of the study

The usual wave energy resource uses the wave gross power, defined as the wave
power that crosses through a circle of 1 metre (Folley and Whittaker (2009)). This
definition includes all possible wave directions of propagation. This is why this
measure of the wave resource is not useful for evaluating a site where devices capturing wave power in one direction only are to be installed. On the contrary, the
net wave power is defined as the wave power for one direction of propagation.
Since wave refraction is not a dissipative phenomenon, there is almost no net
wave power density loss from offshore location to nearshore site. However, the
percentage of wave power density loss from offshore to nearshore site increases
significantly due to directional dispersion (Figure 2.10).
In order to compare more accurately offshore and nearshore resource, Folley and
Whittaker (2009) defined a new representation for the wave resource called exploitable wave energy resource. Defining a threshold wave power for the
device/wave farm as four times the average power received (above this threshold there is no increase in power capture for more energetic sea-states), the exploitable wave energy resource is defined as the average wave power in a fixed
direction as limited by the aforementioned threshold. This makes it possible to
get a better comparison (regarding the wave resource, rather than the economics
of the devices) between offshore gross power captured by devices (regardless of
wave direction) and nearshore power.
As an example, Table 2.1 compares the values for the three different definitions of
wave power resource. The table shows that the reduction of wave power from an
offshore to a nearshore site using the gross wave power is reduced from 20% to 7%
using the exploitable wave energy resource definition.

Deep water
(200m)

Offshore
site

Nearshore
site

Average gross wave power kW/m


Average net wave power kW/m

30.8
26.8

27.2
25.0

21.8
21.4

Average exploitable wave power kW/m

22.6

21.2

19.7

Table 2.1: Comparison of wave power calculations for West Orkney site, after
Folley and Whittaker (2009).

24

Chapter 2 Wave energy glossary and relevance of the study

Another advantage of using nearshore power is that devices placed in this region
will not undergo the same hazards as offshore devices, such as storms or rogue
waves. This has a considerable impact upon the cost of the structure and hence
of the electricity generated.
In summary, it may be the time to reconsider the potential of nearshore regions
for harnessing wave power. Some concepts, rapidly introduced in section 2.3.3.4,
are aimed for that purpose and are presented in greater detail next.

2.4.3.3

Nearshore bottom-mounted wave energy converters

The OYSTER (Whittaker et al. (2007)), developed by Aquamarine Power (Scotland) is a hinged flap attached to the seabed that is driven by the back and forth
movement of the waves. The oscillation of the flap (shown in Figure 2.11) compresses and extends the two hydraulic cylinders, which in turn, pump water ashore
at high-pressure supplying the hydroelectric turbine. After tank testing at scale
1/40th and later 1/20th at Queens Univerity of Belfast, the first full-scale prototype, Oyster 1, was pinned to the seabed at EMEC in summer 2009. It was later
connected to the national grid in October of the same year.
The 18m wide flap was constructed of five cylinders with a height of 11m. It
was set in a mean water level of 10m, so that the top flap emerged. The analysis
carried by Folley et al. (2007a) has shown that the power capture depends primarily on the incident wave force. For that reason, the flap is surface piercing in
order to maximise the force acting on it. This 350kW rated prototype operated
for 6000 hours, giving valuable information for the next built generation of Oyster
2.
Oyster 2 consists of three Oyster 800 devices (rated at 800kW). The width of
the flap has been extended to 26m, as a wider flap was shown to be more efficient
(Folley et al. (2007b)). However, too wide a flap experiences greater structural
load (Cameron et al. (2010)) and this is why this width was not extended more.
The first Oyster 800 has been installed at EMEC and two others, one in 2012 and
one in 2013, will make the 2.4MW wave energy farm complete.

Chapter 2 Wave energy glossary and relevance of the study

25

Figure 2.11: The offshore Oyster 1 device, after Cameron et al. (2010).

The WaveRoller designed by AW-Energy uses the same concept as the Oyster but
is not surface piercing, using only the (less powerful) bottom-waves. It is developed by AW-Energy (Finland) and has continuously been developed since 1999
with small-scale tests in real sea and laboratory for validating the concept. This
led to the first version of WaveRoller, a 3.5m high and 4.5m plate anchored in water depths of 10-15m off the coast of Peniche, Portugal, in 2007 and rated at 10kW.
Successful results led to the development of the next generation of WaveRoller
prototype. It showed good capability in harnessing wave power for a wide range
of wave period.
The second step, helped by EU funding, is to develop a first full-scale demo consisting of three plates rated at 300kW and linked together. This full scale prototype
has arrived (January 2012) at Peniche and is awaiting installation at the same
location as their smaller prototype.

2.4.3.4

Conclusions

Among all the new extraction techniques, more and more are aimed at nearshore
locations for the reasons mentioned. These are mainly easy access and less rough
(yet interesting) wave climate and short underwater electrical cable (conversion
can be done onshore), both arguments leading to a cost reduction of the electricity
produced.

26

2.5
2.5.1

Chapter 2 Wave energy glossary and relevance of the study

Starting point of the project


Introduction

It has been shown that the move into using nearshore power has taken place in
the last years. This is mainly due to the fact that these locations are easier to
access and much less prone to severe storm occurring in offshore sites. In an effort
to use these locations, the available energy has been reevaluated and reconsidered. Nearshore sites have always been seen as less energetic than their offshore
counterpart because the offshore power spectrum considers contributions from all
wave directions. However, most devices have one or two predominant directions
for absorption and disregard the energy coming from other directions. In contrast, wave energy is concentrated in the nearshore in a narrow band of directions,
strengthening the potential use of directionality dependent WEC.
In the Anaconda, bulge waves travels back and forth in a long rubber tube placed
beneath the water surface, extracting the energy form the wave above. The theory
and the main advantages of using and developing such a concept have been presented, and recent developments show the potential of the Anaconda in harnessing
wave power.
These two aspects justify the investigation of the potential of a seabed-mounted
version. In order for the membrane to be under the influence of dynamic pressure
the device must be set in shallow waters and hence addresses harnessing nearshore
wave power.

2.5.2

Aim and objectives

This project will be driven by different aims to be achieved through either experimental work or theory. It is worth mentioning again here that the aim is not to
develop another WEC that would supposedly be more efficient than the others,
less noisy or have less impact on the surrounding environment. The main aim
is to find out whether the bulge wave concept developed through the
Anaconda project is applicable in a seabed version. In other words, be
able to explain why, or why not, it can harness wave power. This aim is achieved
with precise experimental and theoretical objectives in mind.

Chapter 2 Wave energy glossary and relevance of the study

27

Regarding the theoretical study of the present geometry, the objectives consist
of
determining the membrane deflection for different pressure heads in static
water.

determining the theoretical free bulge speed.


having a validated theoretical model for predictions of pressure amplifica-

tions (closed ends) and power capture (PTO) regarding parameters involved
(membrane tension, membrane submergence, water depth and PTO characteristics).

The intended experimental work can therefore be summarised as:


designing and building a long rectangular duct covered with an elastic rubber
membrane to be tested in the available facilities. Equip it appropriately for
measurements of static deflection, free bulge speed, membrane deflection,
and pressure oscillation beneath waves.
test a closed end configuration beneath waves with the purpose of investigating pressure oscillation at its stern.
equip it with a power take-off at its stern for measurement of extractable
power.

2.5.3

Novelty and advantages of the studied device

The design studied is a rectangular duct enclosing water and closed on its top face
with an elastic rubber membrane as shown in Figure 2.12. It rests on the seabed
and is placed in the direction of wave propagation. This configuration results in
the following novel features:
1. use of a distensible device for harnessing nearshore wave power.
2. non-surface piercing bottom-mounted device, like the WaveRoller but without any movable solid parts.

28

Chapter 2 Wave energy glossary and relevance of the study

If possible power capture is to be measured within this study and the potential
of such a device demonstrated, the latter would have the main advantages of
being sited nearshore (cost reduction), made of relatively cheap material and not
interfering with marine and leisure activities.

rubber
membrane
Direction of
wave propagation

bow
z
x

stern
Figure 2.12: 3D view of the tested duct.

2.6

Conclusions

In the effort to rethink the way electricity is produced in societies, wave power
is seen as a viable source. Resource is abundant, renewable and its predictability
ensured by the constant development of spectral models and data assimilation
techniques.
This resource is unevenly spread around the world and this is why some countries more than the others (UK, Ireland, Portugal, USA, Canada, Australia and
Japan) launched research and developments programmes in the 1970s, mainly triggered by the oil crisis of 1973. Those efforts have paid off with the development
of full-scale testing and wave farms projects.
The cost of the produced electricity is still unattractive compared to traditional
ways (oil and coal), or even offshore wind energy, for reasons deliberately not mention here. Techniques, on the other hands, are known and have been detailed in

Chapter 2 Wave energy glossary and relevance of the study

29

section 2.3.3. Unconventional devices are created each year but leading technologies are usually quite simple. The main effort is in reducing cost of operation and
maintenance.
The two main motivations of the current study have been given and a description of how a combination of them could possibly lead to the development of a
device that could contribute to this cost reduction task. The studied device, a long
rectangular duct enclosing water with its top face made of rubber was fashioned
and aimed to be tested following precise tasks.
The advantages of using rubber have been detailed. However, such use is not
new for marine applications. The next chapter gives a review of studies dealing
with interactions of water waves with submerged elastic membranes or with closely
related configurations.

Chapter 3
Literature review
3.1

Introduction

The general consideration of a fluid-filled membrane covered duct can relate to a


great number of similar studies, in various areas of science. The aim of the literature review is to detail previous investigations carried out and theories developed
that can be of use for the present study. This is also a way to synthesise the
information provided and to spot areas where the present work can contribute in
extending the current knowledge. A first section gives some generalities on the
use of rubber in the marine environment. The studied device presented in Figure
2.12 has various similarities with previous studies dealing with use of submerged
membranes for marine coastal applications, and in particular, flexible breakwaters
of various shapes, and this is detailed in another section.

3.2

Use of rubber for marine applications

Rubber use is not limited to the Anaconda (and present work). One of the main
uses of rubber in wave power technologies is for hose pumps and hydraulic pipes in
which water (or oil) is compressed to drive a turbine. An example of this concept
is given by the AquaBuOY (Finavera Renewables). This is a buoy composed of a
tube, enclosing a hose pump. This hose pump contains a piston that is driven by
the vertical movement of the buoy. It stretches the hose-pump, resulting in pressurising the seawater contained in it. This is then used to drive a turbine (the first
prototype sank off the coast off Oregon and the whole wave energy project was
31

32

Chapter 3 Literature review

abandoned for economic reasons. Finavera Renewables now focus on wind energy).
Rubber is also used on the Pelamis wave power device where it is used in spring
system technology. It provides an ultra-flexible and hard-wearing seal between the
steel cylinders and the hydraulic units.
Another device based on the bulge wave concept is the Fabriconda (Hann et al.
(2011)). In this case, the tube is made of several cells made of inelastic fabric and
joined together to form a larger central tube, as seen in Figure 3.1. Each cells
is filled with elastic tube in which water can be pumped at desired pressure for
tuning the device. Bulges propagate in the cells and in the central tube and power
is captured as in the Anaconda. The main advantage of this configuration is that
the elastic material is constrained in shape by the inelastic fabric, avoiding any
undesirable aneurysm effects.

Figure 3.1: Example cross-section of a Fabriconda with 10 cells, showing the


structure at its minimum cross-section (left) and at its medium point (right),
after Hann et al. (2011).

Rubber has also been used for WECs enclosing air. Some devices with this common approach can be seen in the AWS-III (AWS Ocean) technology, which is an
updated version of the SEA Clam, that is a ring-shaped array of twenty interconnected cells. Wave energy is transferred into pneumatic energy in the air pumped
in and out of the cells due to wave motion. This moving air spins a turbine which
produces electricity. A 1/9th scale prototype has been deployed in Loch Ness,
Scotland and a full-scale prototype is aimed to be deployed at sea within the year
2012.

Chapter 3 Literature review

33

An earlier similar structure is the Lancaster flexible bag, made of a long spine
aligned head to waves, equipped with air-filled bags on each of its sides, giving
the structure enough buoyancy to float. The up and down motion of the waves
compress the bags and the air inside, and the resulting flow is used to drive a
turbine.
Outside wave energy conversion, rubber is used for various marine applications.
For instance, inflatable flexible membrane dams (IFMD) have been deployed in
various regions. They consist of bottom-mounted (usually on a concrete block)
rubber bags. They can be inflated with either air or water or a combination of
both, although air is the most common type as it is faster to inflate/deflate the
membrane with air than water. They are commonly placed across rivers and in
coastal regions to prevent flood, control irrigation, or for water storage. General
theories for the static shape under load have been studied by various researchers
and are nicely summarised in Chu et al. (2011) for various bottom topography
and either air or water inflation. Practical uses have been studied by Chanson
(1996) for instance, who also detailed the design techniques used for controlling
the flow over and behind the membrane to avoid instabilities of the whole system. Many of these IFMDs exist today (more than 2000) at every scale, from a
small design application (rice cultivation at Baghamari, Odisha state, India) to
large scale, such as the ones in Ramspol, Netherlands (Jongeling and Rovekamp
(1997)) or at Linyi (Shandong Province, China). Some of these IFMDs burst as
the one in Temple Lake (Arizona, USA) last year that is to be replaced by a steel
dam. Those membranes are usually not made of rubber only, but together with
other variety of geotextiles.
More closely related to the present work are the previous investigations dealing
with the use of flexible membrane for breakwater purposes. They are described in
the next section.

3.3

Breakwaters

3.3.1

Introduction

In order to protect harbours and coasts from the negative influence of strong
waves, especially during storms, use of breakwaters has been studied as a possible solution for wave attenuation. Conventional breakwaters are structures that

34

Chapter 3 Literature review

are made to firmly stay on the sea-floor. When incoming waves hit them, wave
energy is partly reflected and partly dissipated. A calmer sea state is obtained
behind the breakwater. These breakwaters are efficient but require high costs of
installation/maintenance and are designed to stay in one specific location. Although submerged horizontal plates can be seen as a potential solution (McIver
(1985); Patarapanich and Cheong (1989); Neelamani et al. (1992)), studies have
also focused on the use of flexible structures to counter these drawbacks. They
could be used as single breakwaters or be placed seaward of the solid structures
to reduce the amount of energy reaching them by dissipating the energy carried
in long waves.

3.3.2

Submerged elastic membrane breakwaters

Relevant work has been done regarding the use of elastic membranes as possible
breakwaters. An important part of the literature deals with vertical membranes
as one of the configurations shown in Figure 3.2. The configurations correspond
to: (a) surface piercing and hinged-hinged or hinged-elastically supported (at the
top) membrane, (b) submerged buoy-membrane system and (c) floating buoymembrane system.

buoy
d
l

(a)

(b)

(c)

Figure 3.2: Vertical membrane configuration.

Chapter 3 Literature review

35

Figure 3.3: Reflection coefficient of hinged-hinged membrane breakwater as


function of dimensionless tension Tz /gh2 and wave number kh, after Kim and
Kee (1996).

Kim and Kee (1996) studied the first configuration theoretically. They developed
an analytical model, based on an eigenfunction expansion method for the fluids
in both regions and represented the membrane as a tensioned spring being driven
by the difference of potentials in each regions. They also developed a boundary
element model using a simple source distribution method coupled with the discrete
form of the membrane equation. These models were used to theoretically represent the behaviour of the membrane in waves, and a study of best efficiency of the
structure as a wave barrier was carried using the different parameters involved.
Kim and Kee (1996) used their model to study the influence of the membrane tension, boundary condition (hinged-hinged or hinged-elastically supported) at the
membrane ends and mass of the membrane. The general conclusion is that such
system can be efficient as a wave barrier. Complete reflection of incident waves
could be obtained. While the tension in the membrane plays an important role in
the efficiency of this system, the mass of the membrane was seen to be without
influence, because the membrane mass is negligible compared to its added mass.
An example of reflection coefficient has been reproduced in Figure 3.3 where the
overall efficiency of the system can be appreciated. An interesting observation is
that the reflection coefficient drops at a specific frequency and that this frequency
increases as the membrane tension increases. Those peaks were found to be rep
lated to the lowest natural frequency of the system (/h) Tz /mt , where Tz is the

tension in the membrane and mt the sum of the membrane mass and its added
mass due to the surrounding fluid (the latter was approximated as the added mass

of a solid plate of unit width, /4). The mesh plot of the membrane response as

36

Chapter 3 Literature review

a function of wavenumber and position along the membrane was presented, and
at the resonant frequencies, the amplitude of this displacement was maximum,
yielding large transmission coefficient, hence minimum reflection.
Such a configuration is unlikely to be used in practice. This is why Kee and
Kim (1997) extended the previous study to more feasible situations ((b) and (c)),
where the top part of the membrane is attached to a buoy, either floating or submerged. The buoyancy of this buoy provides the tension in the membrane. Their
study was theoretical and experimental. They compared the results from their
boundary element method (with different boundary due to the presence of the
buoy, the membrane equation being the same) with experiments carried out in a
2D tank. These experiments are more detailed in Kim et al. (1996). They studied
the influence of mooring type (two types are shown in Figure 3.2 and a third one
is with both at the same time), buoy radius and membrane parameters. The overall conclusion is that for both configuration (floating and submerged), optimum
conditions on buoy radius and membrane parameters exist for wave-blocking efficiency. This study was extended later by Cho et al. (1997) who used the same
model but using oblique incident incoming waves, yielding the same kind of study
for best efficiency.
It is interesting to note at this point that these previous configurations have been
studied modelling the wave barrier as a tensioned membrane. Other authors have
considered the same geometries, using a beam plate. Among them, Lee and Chen
(1990); Williams (1991); Abul-Azm (1996). As a matter of fact, the tensioned
membrane case is the same as that of a tensioned beam-plate with zero bending
stiffness. The thickness of the membrane in the present work is thin enough to
model it as a membrane. Therefore those studies are mentioned here, but not
developed.
As an extension of these studies, the use of dual vertical membranes was investigated. Cho et al. (1998) investigated analytically the interactions of oblique incident waves with dual hinged-hinged membranes (configuration (a) with notations
from Figure 3.2) and extended it to a more realistic case (configuration (b) and
(c)) by developing a boundary element model using a simple source distribution
coupled with the discrete form of the membrane equation. This computer program
was used to evaluate the performance of the two above-mentioned configurations
and study the influence of several parameters like the separation between each

Chapter 3 Literature review

37

membrane, the buoy physical properties, mooring stiffness and membranes tension. Results were compared with a single membrane system. The hinged-hinged
dual membrane system showed itself more effective, especially in long waves, and
this efficiency was increased by reducing the gap between the two membranes. Interestingly, asymmetric systems (membranes of different tensions, whatever their
sequence was) were seen to be more efficient for a wide range of wave frequencies
and headings. The dual buoy-membrane system was also evaluated. As for the
hinged-hinged dual membrane, the system performed better when the gap between
the two buoy-membranes is small (compared to the wavelength). When the gap is
large, using two membranes instead of one does not seem advantageous. The use
of asymmetric systems (different stiffness between front and rear moorings) was
also observed to widen the bandwidth of best efficiency.
The hinged-hinged dual membrane was also investigated by Lo (1998) using eigenfunction expansion series for the fluid potential in the three regions. Influence of
the gap between the two membranes, their tensions and their mass was studied.
As for the single membrane case, varying the masses of the membranes seemed
without influence on the transmission of the system as those masses were anyway
negligible compared to the added masses of the membranes. Lo (1998) plotted
the modal amplitudes of the two membranes in order to study the influence of the
membranes displacement on the transmission coefficient. The membrane response
was seen to be dominated by the two lowest modes and the largest modal amplitudes corresponded to maximum transmission coefficient (hence lower reflection
coefficient). This is similar to what happened for the single membrane of Kim
and Kee (1996) as mentioned earlier. Interest was given also to the viscoelastic
losses in the rubber, which were found to be negligible even for large membrane
displacement.
Lee and Lo (2002) studied the case of successive vertical surface piercing membranes with variable draft. An eigenfunction expansion method was used for expressing the the velocity potential in each regions separated by a membrane. Influence of the protrusion and draft of the membranes were evaluated, together with
different membrane tensions. This was compared with experimental values. Good
agreement between model and experiment were obtained when Lee and Lo (2002)
included in their model the energy loss due to their experimental configuration
(spacing between tank walls and membranes).

38

Chapter 3 Literature review

The main problem with any vertical configurations is the large wave loading on
the membrane and a possible blockage of currents, with everything that it can
include on the surrounding ecosystem.
That led Cho and Kim (1998) to study the possibility of using the membrane
horizontally rather than vertically, as shown in Figure 3.4, where the tension Tx is
applied in the direction of wave propagation. Cho and Kim (1998) developed an
analytical model (expansion series with matching of the potentials at the interface)
and numerical one (boundary element method). Both methods yielded the same
results and the ones from the analytical model were compared with the results obtained from experiments. As will be detailed in chapter 8, their analytical model
was coded and checked during the course of this study with the aim of using it
for the duct configuration. This is why all figures from Cho and Kim (1998) are
indeed not a reproduction, but another computation as described in chapter 8.

d
Tx
h
s

Figure 3.4: Horizontal membrane configuration.

While the model with infinite tension was successfully compared with McIver
(1985) results for the rigid plate, the influence of the different parameters were
studied in order to predict conditions for an efficient wave-blocking system. For
instance, the membrane tension was varied for fixed membrane length and submergence (s/h = 0.5 and d/h = 0.2 respectively). It was seen that these flexible
membranes perform better than rigid plates as the bandwidth of best efficiency for
the case of a membrane with infinite tension (corresponding to a rigid plate) was
extended by varying the membrane tension and allow its motion. The length of
the membrane was also varied and the reflection coefficients for various configurations are given in Figure 3.5. Regions of best efficiency are noticed for all of them

39

Chapter 3 Literature review

and interestingly, the largest bandwidth is obtained for s = 0.4h meaning that the
efficiency was not necessarily improved by the longer size of the membrane.

Figure 3.5: Reflection coefficients of a submerged impermeable membrane


breakwater as function of length of membrane s/h and wavenumber k1 h for
d/h=0.2, Tx /gh2 =0.1, after Cho and Kim (1998)

4
3
/ A 2
1
0
0

1
kh

0.5
4

0
0.5

x/L

Figure 3.6: Response of the membrane (/A) as a function of wavenumber


k1 h and horizontal coordinate x/L for d/h = 0.2, s/h = 0.5 and Tx /gh2 =0.1,
after Cho and Kim (1998).

40

Chapter 3 Literature review

Figure 3.6 shows the response of the membrane (normalised by the incident wave
amplitude A) for the case s/h = 0.5 with the other parameters fixed as in Figure
3.5. It is seen that the efficiency of this membrane as a wave barrier for that
configuration is linked to its displacement not being too small.
Because of the geometry, this efficiency does not lie in an actual wave blocking but, as will be seen later for the bottom-mounted cases, into the cancellation
of the transmitted waves by the radiated waves from the membrane displacement.
The first peak for the reflection coefficient for that configuration corresponds to a
shape characteristic of the membrane first mode: this is the contrary to what happened for the vertical membrane mentioned earlier, where the reflection coefficient
dropped down at the membrane natural frequency(ies). Figures 3.7 and 3.8 shows
that the membrane displacement at this first peak is mainly composed of the first
symmetric mode, while the contribution of the higher mode is more pronounced
for the higher k1 h for the total membrane displacement.
Later Cho and Kim (1999) studied the same configuration but with a circular,
flexible, horizontal and stretched (uniform tension Tu ) membrane. Using a similar
expansion as for the case of the rectangular membrane but in polar form, they expressed the velocity potentials ahead, above and beneath the membrane. Results
from this analytical model were used to evaluate the difference between circular
rigid and flexible membranes. As before, the case with infinite tension in the membrane was successfully compared with the results from Yu and Chwang (1993) for
the case of a rigid submerged circular disk. Cho and Kim (1999) computed the free
surface elevation for various tensions and submergence depths and noticed that
for small membrane tension (0.02 < Tu /gh2 < 0.05), pronounced peaks of wave
focusing (free surface amplitude three to four times the incident waves amplitude)
in front of the membrane occurred even for deeply submerged membranes. For
those frequencies, the membrane response exhibited interesting patterns (two or
four peaks, sombrero shape) that are characteristic of the different modal shapes.
For higher tension in the membrane (Tu /gh2 = 0.1) this wave focusing occurred
for low values of membrane submergence. The influence of this submerged disk
was made more noticeable by plotting (contour map) the free surface elevation
around the disk for various disk size, submergence and tension. For different yet
close conditions, free surface patterns were found to be totally different and again,
those patterns were linked to the membrane response and which mode was predominantly excited.

41

Chapter 3 Literature review

4
3
2
1
0
0

kh

0.5
4

0
x/L

0.5

0.04
0.02
0
0

kh

0.5
4

0
x/L

0.5

x 10

4
2
0
0

1
kh

0.5
4

0
0.5

x/L

Figure 3.7: Modal response amplitude as function of wavenumber k1 h and horizontal coordinate x/L for d/h=0.2, s/h=0.5, Tx /gh2 =0.1; upper plot: first
symmetric mode, middle plot: second symmetric mode; lower plot: third symmetric mode.

42

Chapter 3 Literature review

0.2
0.1
0
0

kh

0.5
4

0
x/L

0.5

x 10

10
5
0
0

kh

0.5
4

0
x/L

0.5

x 10
3
2
1
0
0

1
kh

0.5
4

0
0.5

x/L

Figure 3.8: Modal response amplitude as function of wavenumber k1 h and


horizontal coordinate x/L for d/h=0.2, s/h=0.5, Tx /gh2 =0.1; upper plot: first
asymmetric mode, middle plot: second asymmetric mode; lower plot: third
asymmetric mode.

Chapter 3 Literature review

3.3.3

43

Submerged fluid-filled membranes

An important amount of studies dealing with rubber use is linked to the concern
of storage or transport of oil or freshwater at sea, early introduced by Hawthorne
(1961). Such devices are usually floating fluid-filled membranes. Importance has
to be given on the dynamic tension in the flexible container (resistance) and on
how to identify the resonant response of such system. These concerns have been
studied for instance by Zhao (1995); Zhao and Aarsness (1998) (first concern) and
Phadke and Cheung (2001) (second concern) and many more researchers. In practice, those bags are made not only of rubber, but combined with other types of
textiles to their design.
Closer to the present submerged configuration are the submerged fluid-filled membranes used for breakwater purposes. Frederiksen (1971) proposed a flexible fluidfilled bag placed head to waves just beneath the water surface. The bag was made
of rubber, but slightly reinforced with nylon fabric. Bags of different thicknesses
and lengths were tested, filled with fluid with different densities. Values of wave
attenuation (ratio of the difference of the incident wave height and the transmitted wave height to the incident wave height) between 0.8 and 1 were obtained for
bags with length two or more times greater than the wavelength. Others values
lay in the range of [0.3:0.8], showing the effectiveness of this device in attenuating
incident wave amplitude.
Broderick and Jenkins (1993) carried out experiments on a submerged fluid-filled
cylinder, placed horizontally across a wave tank (with its axis perpendicular to the
direction of wave propagation). This system was tested for two submergence depth
and was observed to be efficient in reducing the transmitted wave energy. Broderick and Jenkins (1993) used string pots to measure the membrane displacement at
three points on the cylinder and noticed that the maximum displacement occurred
for wave frequency at which the transmitted wave energy is a minimum. They also
equipped the cylinder with a pressure transducer to measure the internal pressure,
but the obtained data did not show any extreme for the low transmission frequency.
Later, Broderick and Leonard (1995) developed a boundary element method, coupled with the finite element model for the membrane equation discretisation, to
study the behaviour of the previous cylinder in waves. The free water surface on
top and aft the cylinder was computed and compared with the case of a rigid cylinder, showing the positive effect of using a flexible cylinder as a wave attenuator.

44

Chapter 3 Literature review

Results from the numerical model were also compared to the experimental ones:
the free surface elevation was under predicted by the model while the displacement
of the membrane was over predicted, the general trend being still satisfactory. This
discrepancy might come from the fact that Broderick and Leonard (1995) assumed
the internal pressure as constant in their model.

3.3.4

Bottom-mounted fluid-filled membranes

Bottom-mounted membrane devices can either be fully or partially filled with


water. For the latter case, Frederiksen (1971) described the experimental work
carried on what he called a rubber mattress. It was basically made of two mattresses made of rubber-nylon, enclosing approximately 50% water and 50% air,
and connected by a duct. When incoming waves were propagating above it, air
pockets were formed and propagated forth and back, altering the wave above.
The study mentioned an attenuation of 40% of the incoming wave amplitude for
wavelength equal to twice the device length.

Incident wave

Transmitted wave

Reflected wave

Radiation wave

Radiation wave

Incident wave
Transmitted wave

Reflected wave

Figure 3.9: Wave interception by a pair of sea balloons, after Ijima and Uwatoko (1985).

45

Chapter 3 Literature review

Another partially fluid device is the so-called sea balloon, as proposed by Ijima and
Uwatoko (1985). This consisted of a vertical tube, made of rubber, with its axis
perpendicular to the direction of wave propagation. About 60% of its volume was
filled with air, giving the cylinder a bulb shape. Sea balloons could be used in pairs
by being connected to each other by an air pipe . Beneath the waves the volume
of air in one balloon changed, affecting the shape and response of the following
one. Numerical analysis (boundary integral method) and experimental work were
carried out to determine the possibility of attenuating incoming waves with this
configuration and the effects of size, number of balloons, gap between them and
the coefficient of resistance of the air-flow in the pipe on this damping. Those pairs
of balloon could be used on a artificial submerged plate or on the seabed. The
overall configuration was seen to be effective in long waves, where the transmitted
wave is cancelled by the radiated wave from the balloons displacements by the
process shown in Figure 3.9.

d
h
p=rgh+p0
L
Figure 3.10: Sketch of the flexible mound device.

Most of the flexible breakwaters were thought to be completely water-filled membrane. For instance, Ohyama et al. (1989) proposed a device called flexible mound
that consisted of a thin elastic membrane bag filled with water and mounted on
the seabed. A sketch of it has been reproduced in Figure 3.10 with notations
used in the present study. The motions of the internal and external fluids were
computed using two boundary elements models, which were coupled through a
force-equilibrium relationship for the membrane, the latter being modelled as a
series of spring. Experiments were carried as well in order to validate the mathematical model and study the influence of internal pressure, submergence of the
membrane or length of the device on the reflection and transmission coefficient.

46

Chapter 3 Literature review

Peaks of maximum reflection (hence minimum transmission) were obtained theoretically and experimentally at some specific frequencies. At those frequencies,
the membrane displacement was computed, and exhibited nodes and antinodes
specific of the natural mode shapes of the membrane. However, the analysis carried by Ohyama et al. (1989) could not predict those resonant frequencies. The
general conclusion regarding this geometry was that this concept was efficient under certain circumstances to attenuate the incoming waves and create a calmer
sea-state behind it. That is why later two of these were installed to protect a harbour (Tanaka et al. (1992a)). They were 9m wide, 40m long and 3m high (when
inflated). They were normally deflated so liners could pass over them, and inflated
in case of emergency, as during a storm, to protect the harbour.

Figure 3.11: Wave energy loss in the flexible mound, after Tanaka et al.
(1992a).

For the same device, Tanaka et al. (1992b) extended the information they obtained
from the experimental tests mentioned above and the power lost in the system.
This has been reproduced in Figure 3.11 where it is seen that up to 50% of the
incident wave power was lost. However, no explanation is given on the mechanisms
of this energy loss.
Phadke and Cheung (1999) used a second-order differential equation governing the
membrane deformation, derived from the membrane theory of cylindrical shells to
deal with the same geometry as the flexible mound. Because they extended the
study of Ohyama et al. (1989) to cases where the density ratio between the inner and outer fluid is different from one, they studied first the static shape of
the membrane which was a function of the hoop stress, which is the circumferential stress resulting from the difference between internal and external pressures.

Chapter 3 Literature review

47

The inner and outer flows were computed using a boundary element method and
coupled to the finite element model obtained from the membrane equation. This
study was theoretical and the computed values of reflection and transmission coefficients compared with the experimental values of Ohyama et al. (1989). Peaks
of zero transmission are theoretically observed for specific frequencies and over a
wide range of conditions. For the same conditions, the dimensionless membrane
response was given for the peak frequencies (Figure 3.12). Node and antinodes
are observed yielding the conclusions that those those frequencies were in fact the
natural frequencies of the fluid-membrane system and completing the study of
Ohyama et al. (1989). Finally, they performed analyses on the influence of the
density ratio between the inner and outer fluid and concluded that the response
amplitude of the membrane deformation decreased with increasing density ratio
due to an increase in both the fluid mass and the membrane hoop stress.
Later, Phadke and Cheung (2003) carried out the same study in the time domain to incorporate the geometric and material non-linearity of the membrane.
Frequencies of zero transmission were shifted towards higher frequencies due to
hysteresis. Normalised responses of the membrane exhibited the same patterns as
in Figure 3.12 while their magnitudes were smaller than the magnitudes of the
normalised response obtained with the linear model. Moreover, Phadke and Cheung (2003) showed that the magnitude of the normalised response decreased with
increasing wave amplitude due to a phenomena called strain stiffening, coming
from the nonlinear term in the strain-stress relationship of the membrane, making
the membrane stiffer when it was more and more deflected.

Figure 3.12: Amplitude of membrane normal deflection for p0 /gh = 0.07


and d/h=0.5: (a) fundamental mode; (b) second mode, (c) third mode ( ,
i / = 0.9; , i / = 1.0; , i / = 1.1), after Phadke and Cheung (1999);
i , density of the inner fluid.

48

Chapter 3 Literature review

Despite the inclusion of the membrane non-linearity in their model, the theory from
Phadke and Cheung (2003) could not fairly agree with the experimental results
from Ohyama et al. (1989). For that purpose, Das and Cheung (2009) extended
the initial work of Phadke and Cheung (1999) by extending it to three dimensions.
The problem was formulated as before for the membrane and outer and inner flow,
taking into account the supplementary dimension. An interesting comparison is
to plot, for the same configuration, the transmission coefficient obtained from the
two dimensional model (Phadke and Cheung (1999)), the two dimensional model
with the non-linearity of the membrane, the linear model from Das and Cheung
(2009) and to compare it with the experimental values from Ohyama et al. (1989).
This comparison has been reproduced in Figure 3.13. The main comment is that
peaks of complete zero transmission were non longer observed and that was more
consistent with the laboratory data. The other interesting fact is that the low
transmissions frequencies are obtained for higher frequencies than the two other
models. This is a result of Poissons effect on the membrane, yielding a lateral
stress affecting the natural frequencies of the system. The obtained model seems to
fairly agree with the data from Ohyama et al. (1989), yielding the conclusion that
non linear effect are less prone to explain the discrepancies between Phadke and
Cheung (1999) model and Ohyama et al. (1989) than the three dimensional effects.
By mapping the free surface wave ahead, above and behind the membrane, Das
and Cheung (2009) could characterise the modification of the incoming wave by
the membrane displacement and the reason of low transmission coefficients for the
resonant frequencies. The motion of the membrane at those resonant frequencies
is associated with the generation of radiated waves that travel opposite to the
incoming waves in front of the structure, generating standing waves, but travelling
in the the same direction behind a structure, but with a 180 lag, cancelling out
the transmitted waves.

Chapter 3 Literature review

49

Figure 3.13: Transmission coefficient as function of excitation dimensionless


frequency L/: (a) p0 /gh = 0.065, (b) p0 /gh = 0.208; data from

Ohyama et al. (1989); two dimensional linear model from Phadke and

Cheung (1999); two dimensional non linear model from Phadke and Cheung
(2003); three dimensional model from Das and Cheung (2009)

The same type of device was investigated theoretically by Liapis et al. (1996), who
considered waves over a submerged, bottom-mounted water-filled membrane that
was semicircular in shape, as shown in Figure 3.14. The membrane deformation
was computed by finding the hydrodynamic pressure acting on the membrane at
each time step and using this pressure to compute its deformation at the next
time step. The efficiency of the structure was, as before, tested by considering the
reflection and transmission coefficients obtained from varying the parameters of
the problem (length of the structure, submergence). These values were compared
with that obtained from the rigid structure of the same shape. The general conclusion was that the flexible one was more effective as a wave barrier than its rigid
configuration. The membrane displacement was not included in this analysis of
best efficiency.

50

Chapter 3 Literature review

d
h

z
Rm
x

Figure 3.14: Water-filled hemicircular membrane breakwater.

Dewi et al. (1999) extended the configuration of Liapis et al. (1996) to three
dimensions, considering the membrane as a shell. The tested model was a 150m
long shell of 4m radius placed in a water depth of 6m. A boundary integral
method was used for the inner and outer fluid and coupled with a finite element
method for the shell. Before dealing with the structure in waves, the different
modes of vibrations of the fluid-membrane system were computed for an internal
hydrostatic pressure 5% higher than the external hydrostatic pressure. Wet and
dry modes were computed by solving the membrane equation of motion without
external force. Wet modes corresponded to the natural modes of vibrations when
the system was in water. On the other hand, dry modes corresponded to the
natural modes of vibrations if the system was placed in air and the enclosed
volume was subject to the same pressure difference. When the system was subject
to incoming waves, the free surface elevation was computed as in Figure 3.15
where the incident water waves have an amplitude of unity. In this figure, the
free surface elevation obtained for the rigid structure of the same geometry ((a)
and (b)) was compared to the one obtained from the flexible version ((c) and (d)).
The shell expands from x=-4m to x=4m and from y=0m to y=150m. Yellowish
regions were reduced from upper plots to lower ones, especially for the higher
frequency, yielding the conclusion that the flexible structure was more effective
in reducing the transmitted wave. Red regions are visible ahead of the structure
and correspond to the combination of incident and diffracted (and reflected for
the flexible structure) to form partial standing waves in front of the structure.

51

Chapter 3 Literature review

There is however no clear explanation why the two frequencies used for this plot
( = /4 and = /6) were chosen rather than any of the wet modes for the
fluid system. A shadow-zone transmission coefficient was defined as the average
wave amplitude along the line y=75m (middle point of the membrane in the y
direction) from x =4m to x =50m. Results have been reproduced in Figure 3.16
for the flexible structure and three different incident angles ( = 0 representing
waves parallel to the x-axis). It can be seen that the optimum efficiency is obtained
when this structure is placed perpendicular to wave propagation direction.

(a)

(b)
1.4
1.3
1.2
1.1
1
0.9
0.8
0.7
0.6

y(m)

170
150
75
20
0
-20

1.4
1.3
1.2
1.1
1
0.9
0.8
0.7
0.6

(c)

(d)

y(m)

170
150
75
20
0
-20
-50

0
x(m)

50

1.5
1.4
1.3
1.2
1.1
1
0.9
0.8
0.7
0.6
0.5

1.3
1.2
1.1
1
0.9
0.8
0.7
0.6
-50

0
x(m)

50

Figure 3.15: Free surface wave amplitudes for normal waves over (a) rigid and
(c) flexible structure for = /6 and over (b) rigid and (d) flexible structure
for = /4, from Dewi et al. (1999).

52

Chapter 3 Literature review

Figure 3.16: Variation of shadow transmission coefficient KT with kR , after


Dewi et al. (1999).

Stamos (2000) compared experimentally the effectiveness of rectangular and hemicylindrical, flexible and rigid submerged and bottom-mounted breakwaters by
measuring reflection and transmission coefficients ahead and behind the model.
An example of his hemi-cylindrical model is given in Figure 3.17 where the axis of
this bottom-mounted structure was perpendicular to the direction of wave propagation. Internal added pressure (for the flexible models) and submergence ratio
(for the rigid one) were the variables of the problem and a parameterisation on
the influence of those parameters was conducted. An energy loss coefficient was
calculated and explained as wave breaking over the structure or turbulent flow inside the flexible breakwaters. Interesting values are shown in Table 3.1 and Table
3.2 where the flexible and rigid transmission and reflection coefficients are given,
respectively. It is worth noticing that an important amount of energy was lost in
the system and that the role of the reflection was not the principal reason of the
efficiency of the system in term of wave attenuation. Rectangular models (rigid or
flexible) were seen to be the most effective structure to dissipate energy (and not
reflect it).

53

Chapter 3 Literature review

Figure 3.17: Side view of the hemi-cylindrical flexible composite breakwater


model, after Stamos (2000).

h (cm)

d/h

p0 /gh

Kr

Kt

Kl

Hemi-cylindrical
Rectangular
Hemi-cylindrical

22.5
22.5
22.5

0.24
0.24
0.24

0.007
0.007
0.141

38%
32%
49%

54% 69%
35% 87%
59% 60%

Rectangular
Hemi-cylindrical

22.5
22.5

0.24
0.24

0.141
0.282

33%
46%

53% 74%
67% 53%

Rectangular
Hemi-cylindrical
Rectangular

22.5
27.5
27.5

0.24
0.38
0.38

0.282
0.007
0.007

36%
38%
29%

52% 74%
61% 66%
51% 79%

Hemi-cylindrical
Rectangular
Hemi-cylindrical

27.5
27.5
27.5

0.38
0.38
0.38

0.141
0.141
0.282

37%
29%
34%

74% 53%
60% 72%
78% 51%

Rectangular

27.5

0.38

0.282

33%

61% 70%

Table 3.1: Average wave coefficients of flexible models, after Stamos et al.
(2003).

54

Chapter 3 Literature review


h (cm)

d/h

Kr

Kt

Kl

Hemi-cylindrical
Rectangular
Hemi-cylindrical

22.5
22.5
27.5

0.24 40%
0.24 38%
0.38 24%

72% 55%
64% 63%
78% 55%

Rectangular

27.5

0.38 40%

69% 57%

Table 3.2: Average wave coefficients of rigid models, after Stamos et al. (2003).

3.3.5

Conclusions on breakwaters studies

Further developments for flexible breakwaters studies would be the use of porous
membranes, as studied by Cho and Kim (2000) for the horizontal case and Kee
et al. (2003) for two vertical membranes attached at the top of a floating pontoon.
This is not of interest here.
The advantages of using a flexible (light, easy to carry) breakwater rather than
a rigid one led to the ocean-scale deployment of flexible rapidly installed breakwater systems (RIBS). A practical example is given with the RIBS XM99 (Briggs
et al. (2002)) which is a combination of vertical membranes and steel frames. This
breakwater is a V-shaped structure, with each side like the panel shown in Figure
3.18.
It was tested offshore Cape Canaveral, Florida in May 1999 (Briggs (2001)), at
full-scale. Wave transmission in the lee side (inside the V) was evaluated, together with the forces on the structures (pressure transducers were mounted on
each side). A threshold of 0.5 for the transmission coefficient in the lee side was
used, above which the system was considered inefficient. Transmission coefficients
below or around this value were measured during several days and different wave
environment, giving full satisfaction in the use of the XM99 as a temporary breakwater. This efficiency was more pronounced in long waves.
The interactions of water waves with flexible membranes have been widely studied
for the purpose of breakwaters. The studied geometries were considering the membranes as either a tensioned vertical or horizontal sheet, or enclosing water. The
reviewed studies pointed out that such a structure can attenuate the transmitted
waves when properly tuned.

Chapter 3 Literature review

55

Figure 3.18: Rapidly installed breakwater system XM99 (from Briggs (2001))

A common point from the reviewed studies is that peaks of theoretical complete reflection (or zero transmission) coefficient were observed at the natural frequencies
of the system (for the vertical membrane, the inverse is true). This was confirmed
by computing the membrane displacement at these frequencies. However, no experimental data was given for the membrane displacement. Broderick and Jenkins
(1993) used string pots but only at three locations along a 3.66m long cylinder.
Kee and Kim (1997) mentioned measuring the membrane displacement using a
video set-up but did not present their results.
The present study does not have a breakwater objective. However, among the
objectives stated in section 2.5.2, the membrane behaviour beneath waves need
to be measured adequately and for that, an efficient technique using laser sensors
was developed. Using such a technique will provide more consistency in the link
between maximum reflection and resonance of the structure. This would also be
useful for measuring the losses in the rubber due to hysteresis, the membrane displacement being known, and what part of the power losses presented in Figure
3.11 or in Table 3.1 can be attributed to these viscoelastic losses.
In a same way, no experimental (and theoretical) consideration has been given
on the pressure variation inside the fluid-membrane system, in particular at resonance. Again, Broderick and Jenkins (1993) measured the pressure in their flexible

56

Chapter 3 Literature review

cylinder but the trend of the presented results did not show any extrema. Ohyama
et al. (1989) and Tanaka et al. (1992b) noticed an increase of the static component
of the internal pressure when waves passed over their flexible mound but gave no
further details. In the present work, the information regarding pressure oscillation
in the duct is important and will be analysed thoroughly by developing adequate
techniques that could be useful for future similar studies on membranes in waves.
The next chapter gives a description of the instrumentation used for that purpose
and the facilities in which experiments took place.

Chapter 4
Experimental equipment
4.1

Introduction

This chapter describes the experimental apparatus used during this work, together
with the instrumentation. The power take-off system, briefly described in section
2.4.2.3, is detailed, and power calculations are explained.

4.2
4.2.1

Narrow wave flume


Dimensions

The narrow flume at the University of Southampton, School of Civil Engineering


and the Environment, is 17m long, 0.44m wide and about 1m deep. The wavemaker detailed in section 4.2.2 operates with a water depth of 0.7m. The flumes
side walls are made of glass, as can be seen in Figure 4.1. The end of the flume is
equipped with high density foam to minimise wave reflections.

57

58

Chapter 4 Experimental equipment

Figure 4.1: Narrow flume at the University of Southampton.

Figure 4.2: Wavemaker for the narrow flume.

Chapter 4 Experimental equipment

4.2.2

59

Wave-maker

The wave-maker is of a flap-type and is manufactured by Edinburgh Designs. The


flap is dry back and driven by an electric servo-motor via a stainless steel belt
running over the curved top of the paddle (Figure 4.2). Hydrostatic forces acting
on the paddle are compensated by springs located behind it. While running, the
system measures the forces acting on the paddle from any reflected waves and
corrects the paddle motion accordingly.

4.2.3

Absorbing beach

In order to prevent any undesirable reflections, the end of the flume is equipped
with an absorbing beach made of foam blocks (in the same way as for the wave
basin shown in Figure 4.5). The efficiency of this beach is given by Blenkinsopp
(2007) with a reflection coefficient varying from 2% to 13% for the wave conditions
covered in chapter 6.

4.3

Wave basin

4.3.1

Dimensions

The wave basin (Figure 4.3) is 1.5m wide, 0.8m deep and 11m long. It is operated
with a water depth of 0.6m. Its side walls and bottom are made of glass, except
over a distance of 2 meters ahead and behind the wave paddle where the bottom
is made of Grey Plastic. One side is equipped with the wave-maker detailed in
section 4.3.2 and the other with an absorbing beach made of foam blocks.

4.3.2

Wave-maker

The wave-maker shown in Figure 4.4 is of piston type. The paddle is driven by
hydraulic piston rod and moves backwards and forwards horizontally. Four wave
gauges are placed on the front part of the paddle to give a feedback signal to the
wave monitor that includes a dynamic absorption servo loop.

60

Chapter 4 Experimental equipment

Figure 4.3: Wave basin at the University of Southampton.

Figure 4.4: Wave paddle for the wave basin.

61

Chapter 4 Experimental equipment

4.3.3

Absorbing beach

The absorbing beach was built during the course of this study with wedge shaped
foam blocks as seen in Figure 4.5. Its efficiency was tested by running some
waves in the empty basin and using three pairs of waves gauges for measuring the
reflection coefficient. The two wave gauges constituting a pair were separated by
0.5m (in the width direction of the basin). The first pair was located 2.7m from
the equilibrium position of the paddle, the second 0.4m from the first pair, and
the third pair 0.5m from the second.

Figure 4.5: Absorbing beach for the wave basin.

A Fast Fourier Transform (FFT) was used to obtain the amplitude A and phase
of the signals at the three positions (each pair being averaged) and the reflection
coefficient
Kr =

Ar
Ai

(4.1)

62

Chapter 4 Experimental equipment

was calculated. The reflected and incident wave amplitudes, Ar and Ai respectively, were calculated following den Boer (1981) as
A2 cos 2 A1 cos kl
),
A2 sin 2 A1 sin kl
A1 cos kl A2 cos 2
,
=
2 sin kl sin r

r = arctan (

(4.2)

Ar

(4.3)

and
A1 cos kl A2 cos 2
),
A1 sin kl + A2 sin 2
A1 cos kl A2 cos 2
,
=
2 sin kl sin i

i = arctan (

(4.4)

Ai

(4.5)

where the subscripts stand for the pair of wave gauges used and l is the distance
between the two pairs of gauges used. It can be seen that the calculation of Ar and
Ai using this method requires that l 6= n/2 (n=1,2,3...) for the denominator of
(4.3) and (4.5) not to be zero.
The values of Kr were calculated using the first two pairs for wave height used in
this basin (in the same layout used as in chapter 7) and are shown in Figure 4.6.
The second and third pairs were also used to calculate Kr and values were in good
agreement with the ones presented in Figure 4.6.
The primary comment from Figure 4.6 is that the beach was seen to be effective with a reflection coefficient varying from 4% to 15% for wave frequencies
higher than 0.7Hz and equal to 20% for longer waves. This will be discussed in
chapter 7.

4.4
4.4.1

Rubber
Introduction

Rubber being the key material of this work, its different properties need to be accurately defined and determined. The complete procedure for determining those
characteristics is given in Appendix D, leading here to a summary of the methodology used and the obtained values.

Chapter 4 Experimental equipment

63

Figure 4.6: Measured absorbed reflection coefficients for the absorbing foam
beach in the wave basin.

4.4.2

Youngs modulus

To get the Youngs modulus, sample strips were clamped at both ends. One end
was fixed, the other one designed with a hook to place successive loads. First
load was applied and the change of length was recorded, as well as the change of
width and thickness (averaged over three positions along the length of the strip).
The process was repeated with ten successive increasing loads. A linear fit, passing trough the origin was then used to determine the ratio stress/elongation, and
hence the Youngs modulus.
It is worth mentioning that the successive masses were placed with two minute
intervals. A quicker test, with the same masses placed at less than thirty seconds intervals was also carried out and results detailed in Appendix D show few
differences in the obtained Youngs moduli.

64

Chapter 4 Experimental equipment

4.4.3

Structural damping

While successively unloading the strips from their successive masses, the unloading
path was observed not to be the same as the loading path due to the hysteresis
effect. However, the hysteresis loop was not of paramount importance for the
quasi-static Youngs modulus determination as this test required only one cycle of
deformation. This is visible in Figure 4.7 where the stress in the strip is plotted as a
function of its elongation. This is also why the hysteresis was not of relevant effect
for the free bulge speed as the local deformation due to a bulge was happening
fast enough.

Figure 4.7: Stress in the strip of membrane as a function of the elongation


for the quasi-static Youngs modulus measurement. Each symbols represent
different loading.

However, when rubber is repeatedly deformed and restored to its original state
over a certain number of cycles, the role of the hysteresis is to dissipate the energy
involved in this deformation. Natural oscillations of a strip of rubber were damped
in the same way as oscillations are damped in a mass-spring-dashpot system.
This hysteresis damping was obtained by measuring the decaying oscillation of
a mass attached to a strip of rubber released from an initial position corresponding to the strip being stretched. The hysteresis parameter was obtained from the

65

Chapter 4 Experimental equipment

formula detailed in Appendix D. In order to characterise the effect of the hysteresis damping regardless of the frequency of oscillation, it is common to introduce
the loss angle defined as
loss angle = arctan ,

(4.6)

with being the angular frequency of oscillation.

4.4.4

Summary

The complete characteristics of the tested rubber, obtained from the above-mentioned
procedures are given in Table 4.1.
Thickness (mm)
0.2
1

Density r (kg/m3 )
1052
1143

Youngs modulus (MPa)


0.92
1.87

Loss angle ( )
1.8
2.5

Table 4.1: Measured characteristics of test rubber

4.5
4.5.1

Power take-off system


Introduction

The primary concern regarding the PTO design is how to use the horizontal internal oscillatory flow due to the propagation of bulges. One option was to have a
set of high and low pressures hydraulic accumulators at the stern. One-way valves
would allow water to flow into the high pressure accumulator and flow out from
the low pressure accumulator back to the tube, hence maintaining the enclosed
volume of water in the tube constant. A turbine could be placed between the two
accumulators and would be driven by the flow from the high pressure accumulator
to the low pressure accumulator.
The idea developed by Chaplin et al. (2012) and used hereinafter was to transfer
the horizontal internal oscillatory flow into the motion of a slug of water. The
heave motion of this slug of water was used to push air trough a turbine. Experimentally, this turbine was modelled as a linear dashpot, in the form of 17 copper
pipes, each of them containing 140 stainless steel pipes of 1.6mm internal diameter

66

Chapter 4 Experimental equipment

140 pipes, ID=1.6mm

copper pipe, ID=26.2mm


Figure 4.8: Cross section power take-off.

and 0.8m long as seen in Figure 4.8. The air flow through those pipes lost energy
by friction at a rate equal to power. This way, the power captured by a turbine
was modelled as the rate of energy lost by this flow in the pipes.

Capillary
pipes

Air
Vertical
motion
rigid
bent tube

rubber

slug of
water

horizontal
motion
Figure 4.9: Schematic of the power take-off system.

67

Chapter 4 Experimental equipment

Connector
Copper
pipes

Pressure
transducers

Rectangular
bent tube

Rubber

Figure 4.10: Power take-off system set-up

4.5.2

Components

The slug of water was enclosed in a rigid and bent aluminum tube, one side being
connected to the ducts stern, the other to the copper pipes detailed later. This
bent tube was rectangular, having a width of 0.276m and a height of 0.153m, covering a cross section of St =0.042m2 . A two-dimensional representation is given
in Figure 4.9.
The top part of the aluminium tube is connected to the dashpot, via a distortable
aluminium tube as can be seen in Figure 4.10. The vertical displacement of the
slug of water induces an excess of pressure in the chamber above the water-air
interface. This difference in pressure under and above the pipes induce a flow
in the pipes. This flow is assumed to be linear and the flow rate q and pressure
difference p = pa + p pa = p (notations from Figure 4.11) are related by the

68

Chapter 4 Experimental equipment

St

pa

pa+p

Sp

V
Flow rate q=p/ZPTO

e
Figure 4.11: Model for the power take-off.

Hagen-Poiseuille relationship

8air Lp
q,
(4.7)
r 4
with r the radius of one stainless steel pipe, Lp the length of a stainless steel pipe
and air the dynamic viscosity of air. Using the value of air = 1.85 105 Pa.s
at room temperature, the ratio p/q for one copper pipe (140 stainless steel pipes)
p=

is Z1 = 657kPa/m3 s1 . Those copper pipes are in parallel, so the equivalent


impedance for a number N of them open is
1

N
,
Z1

(4.8)

ZPTO =

Z1
.
N

(4.9)

ZPTO
or in other words,

4.5.3

Power calculation

The power take-off can be modelled as in Figure 4.11, where the water-air interface
displacement e acts as a piston, increases the pressure in the chamber and pushes
the air through an exhaust pipe of impedance ZPTO .
Assuming air as an ideal gas leads to
p = RT

(4.10)

where T is the temperature of the air inside, assumed to be equal to room tem-

69

Chapter 4 Experimental equipment


perature and R the gas constant. The rate of change of the mass density is
m
m

2 V
V
V
m
V
=
.
V

(4.11)
(4.12)

The rate of change of mass due to the outflow is


m
= q
p
,
=
ZPTO

(4.13)
(4.14)

while the rate of change due to piston motion is


V = St e

(4.15)

So the rate of change of pressure in the chamber is


F
St
m
V
=
RT
V
p
RT
= (
+ St e)

ZPTO
V
pa
F
+ St e)

,
= (
St ZPTO
V

p =

(4.16)
(4.17)
(4.18)
(4.19)

so from (4.16), multiplying by St


F
pa
F = (
+ St2 e)

,
ZPTO
V

(4.20)

which leads to the equation for the flow rate


q + q = St e,

with
=

pa
ZPTO V

(4.21)

(4.22)

This is the equation that relates the water-air interface displacement to the flow
rate in the pipes. Substituting
e = e0 eit
(4.23)

70

Chapter 4 Experimental equipment

and
q = q0 ei(t+eq )

(4.24)

into (4.21) yields the phase shift between water-air interface displacement and flow
rate
eq = arctan

pa
ZPTO V

(4.25)

and the amplitude of the flow rate as a function of the water-air interface displacement amplitude
St
e0
(4.26)
q0 =
cos eq + sin eq
Energy is dissipated in the pipes as air flows through it. The rate at which this
dissipation takes place (hence power) is
1
p0 q0
2
1
ZPTO q02 ,
=
2

Pf =

(4.27)
(4.28)

where p0 is the amplitude of p. In the case eq = /2, that corresponds to low


values of impedances, (4.26) yields
q0 = St e0 ,

(4.29)

meaning that the the flow rate that goes through the pipes is exactly the same to
that of the air above the water-air interface. However, as the impedance increases,
air is more and more compressed and the phase lag due to this compression is
taken into account by using (4.28).

4.6
4.6.1

Instrumentation
Data acquisition and equipment driving system

A PCI data acquisition card was installed on a Personal Computer and could
accommodate up to 32 input channels and 2 outputs. Purpose made TestPoint
programs were used to either acquire signals from the wave monitors, amplifiers
or laser sensors, or to drive the two wave-makers.

Chapter 4 Experimental equipment

4.6.2

71

Wave probes

The wave gauges used in the experimental work were resistance-type gauges. They
consisted of two wires of 3.2mm in diameter, 620mm in length and separated by
20mm. Wave monitors energised the two wires and output a signal proportional to
the submergence depth. This signal was simultaneously sampled by the indicated
data acquisition card.
Wave probes were used for recording free surface elevation (incident, reflected
and transmitted wave calculation) or the water-air interface displacement of the
slug of water in the aluminium tube shown in Figure 4.10. They were calibrated
daily when in use.

4.6.3

Water pressure transducers

Pressure transducers used in this study are manufactured by Omegadyne (model


PX42G7 as shown in Figure 4.12) and were strain-gauge-based transducers. They
had four strain gauges bonded into the diaphragm (Figure 4.12) forming a Wheatstone bridge circuit as seen in Figure 4.13.
The bridge was excited by an amplifier (Excitation and Excitation ). When
no pressure was applied on the diaphragm, the bridge was balanced and VA = VB .
Pressure acting on the diaphragm resulted in physical deformation of the gauges
and a change in their electrical resistance (R1 to R4 ). The bridge was then unbalanced (VA 6= VB ) and the analog electrical signal produced was amplified and

recorded.

The water pressure transducers were found not to need recalibration often. Starting from a arbitrary zero condition with water level above the transducers, the
water level was increased in the flume (or basin) and the signal delivered by the
above-mentioned bridge was recorded for three more water heights.

72

Chapter 4 Experimental equipment

Figure 4.12: Pressure transducers.

Excitation

R2

R1
Signal

Signal

R4

R3

Excitation

Figure 4.13: Wheastone bridge for the pressure transducers.

4.6.4

Air pressure transducers

For the configuration with the PTO system detailed in 4.5, two air pressure transducers were used to record the pressure variations in the chamber above the waterair interface. They were mounted on a wooden box and linked to the aluminium
tube as shown in Figure 4.10.

Chapter 4 Experimental equipment

73

They were calibrated using a purpose made manometer. Pressure was measured
via the head difference between the two air-water interfaces in the manometer
pipes, with the head difference being increased by blowing more air into the system, and vice versa. This process was repeated every two days when in use.

4.6.5

Digital camera

Pictures presented within this thesis were taken with a 12 Megapixel Pentax K-x
digital camera.

4.6.6

Laser sensors

Five laser sensors from Leuze Electronic (model ODSL 9/V6-450-S12) were used
for measuring membrane displacement when subject to incoming waves, or centreline deflection when inflated and free bulge speed in static water. One of them is
shown in Figure 4.14. These laser sensors had a operating range of measurement
of 50-450mm, a measurement time of 2ms and a resolution of 0.01mm.
For each of them, a 3-Pin DIN plug was used: two wires for the power (25 Volts
DC) and the third one for the signal. This signal was then connected to a BNC
in order to plug it on an oscilloscope or to connect to the data acquisition board.

Figure 4.14: Laser sensor.

74

Chapter 4 Experimental equipment

These laser sensors were not meant to be used under water. Hence, they were
placed in transparent boxes submerged by a depth slightly higher than the wave
amplitude, to be sure the free water surface did not cause undesirable diffraction
of the laser beam. The influence of these boxes on the surrounding flow was seen
to be negligible as the reflection coefficient from the duct system was not changed
by the presence of these boxes, as seen in Figure 4.16.

Figure 4.15: Laser sensors set-up.

These sensors were calibrated for every set of the five positions they were used at.
The first recorded voltage was for the flat membrane. Then successively, wooden
sticks of known thickness (20 mm each) were used to lift up the whole frame (or
just each frame for the set-up in the narrow flume). Output voltage was recorded
and the process repeated. If they were moved, they were calibrated again.

4.7

Tested model

A 5m long rectangular duct was initially built with the corresponding dimensions
shown in Figure 4.17. From those dimensions, the cross-section of the rubber duct
was Sd =0.055m2 . Its length was reduced to 2m for the tests described in chapter
7. The bottom plate was made of Grey Plastic while its side walls were made of
transparent Plexiglas. The side walls were incorporated in the bottom plate in

75

Chapter 4 Experimental equipment

Figure 4.16: Influence of the plastic boxes used for the laser sensors on the
reflection coefficient for the experimental set-up as in chapter 7 for two sets of
laser sensors positions.

v0
300mm
8mm

200mm

15 mm
Bottom plate
(Grey plastics)

Rubber layer

Side walls
(Plexiglas)

Angle brackets
(aluminum)

L-shaped metal
bracket
Figure 4.17: Cross section sketch of the rubber duct.

76

Chapter 4 Experimental equipment

a depth of approximately 5mm and glued inside. All the joints were sealed with
silicone to prevent any leaks. Some L-shaped brackets were also screwed in the
bottom plate to avoid lateral motions of the side walls. The rubber membrane of
thickness v0 (at rest) was clipped using office-style binder clips that proved to be
strong enough to secure the membrane. Photos of the duct with the membrane
are provided in Appendix C.

4.8

Conclusions

Experimental set-ups using the equipment described in this chapter will be detailed
in chapters 6 and 7. Along with this experimental work, a theory for free bulge
speed and rubber duct behaviour beneath waves needs to be detailed: this is the
purpose of the next chapter.

Chapter 5
One dimensional model for bulges
propagation in the duct
5.1

Introduction

The basic theory for bulge propagation in elastic tubes has been introduced when
introducing the Anaconda in section 2.4.2. This chapter details the use of this
theory for the present geometry, in terms of free bulge speed and bulge propagation
in the closed ends when the membrane is beneath waves. The inclusion of a linear
power take-off at the duct stern is then considered, and the method for calculating
the power absorbed is presented.

5.2

Load-deflection of rectangular membrane and


free bulge speed

Consider a long submerged and liquid-filled rectangular duct, covered by an elastic


membrane. Figure 5.1 defines key parameters and shows the membrane bulging to
a height w in response to a positive internal pressure pb , measured relative to the
external pressure pw . When pb is zero, the deflection w is zero, the stress in the
membrane is y,0 , and the membrane thickness is v0 . The profile of the membrane
is the circular arc
p
2 y2 + w R ,
z = Rm
(5.1)
m
77

78

Chapter 5 One dimensional model for bulges propagation in the duct

with

a2 + w 2
.
(5.2)
2w
To calculate the speed at which bulges in the membrane would travel along the
Rm =

length of the duct it is first necessary to determine the relationship between pressure and cross-sectional area, or in other words the distensibility
0

1 dS
1 dS dw
D=
=
,
Sd dpb
Sd dw dpb
where
0

S =

(5.3)

p
2 y 2 + w R )dy.
( Rm
m

(5.4)

Sd
b

pb=0

Sd
pb>0

2a

Rm
fm

Figure 5.1: Cross section sketch

Resolving forces in the vertical direction, the circumferential stress in the membrane is found to be
pb Rm
,
(5.5)
y =
v
where v, the thickness of the stretched membrane, differs from the initial thickness
v0 because of the effect of the Poissons ratio .
Assuming that the longitudinal strain is zero, and that the material of the membrane (like rubber) is for practical purposes incompressible, then
v=

av0
,
Rm m

(5.6)

where sin m =a/Rm (see Figure 5.1). The circumferential strain in the membrane
is
Rm m
1.
(5.7)
=
a

Chapter 5 One dimensional model for bulges propagation in the duct

79

Using the plane strain condition = E/(1 2 ), it follows that


pb =

v
E
Rm m
[
(
1) + y,0 ],
Rm 1 2
a

(5.8)

including the effect of the initial stress in the membrane 0 . Using a Taylor series
expansion for m about w/a = 0 in (5.8) (with y,0 << E) yields
pb =

2v0
w
2v0 2E
w
y,0 ( ) +
(
)( )3 + ...,
2
a
a
3a 1
a

(5.9)

as obtained for long rectangular membranes by Xiang et al. (2005); Vlassak (1994);
Huang et al. (2007); Larson et al. (2007). The linear form in w/a of (5.9) will be
used in subsequent analysis.
The distensibility follows from (5.1)-(5.4) and (5.9) and the free bulge wave speed
following Lighthill (1978)
1
c=
(5.10)
D
by neglecting the effects of longitudinal curvature.

5.3

Bulge equation

The internal (or bulge) pressure pb , the external pressure pw , and the displacement
of the membrane are now considered to be functions of distance along the duct x
and time t. The wave equation for a long duct with a tensioned membrane cover
is essentially the same as that for the Anacondas rubber tube (Chaplin et al.
(2012)):
3 pb
1 2
2 pb

=
(pb + pw ).
t2
t3
D x2

(5.11)

The motion of the membrane is assumed harmonic at frequency , hence representing the bulge pressure pb (x, t) as the real part of Pb (x)eit and the external pressure pw (x, t), generated by the incident waves as the real part of Pw (x)eit . The
internal velocity is also written with the harmonic notation as u(x, t) = U(x)eit
following (2.18)
U(x) =

i d
(Pb (x) + Pw (x))
dx

(5.12)

With this notation, (5.11) can be written as


d2 Pb
d2 Pw
2
+ K Pb =
,
dx2
dx2

(5.13)

80

Chapter 5 One dimensional model for bulges propagation in the duct

with K = k2 1 i and k2 = D.
Assuming the wave pressure to be of the form
Pw (x) = ai eik1 x ,
with
ai =

(5.14)

A
,
cosh(k1 d)

(5.15)

where d is the submergence of the membrane, the general solution of (5.13) is


Pb (x) = C1 cos(Kx) + C2 sin(Kx) +

gai k12 ik1 x


e
,
(K 2 k12 )

(5.16)

where C1 and C2 depend on the boundary conditions.

5.4

Closed ends configuration

For the case of a duct of finite length L closed at both ends, the velocity at the
bow and stern are zero. Following from (5.12), this yields
d
(Pb + Pw ) = 0
dx
d
(Pb + Pw ) = 0
dx

x = 0,

(5.17)

x = L.

(5.18)

The solution of (5.13) with boundary conditions (5.17) and (5.18) can be written
as
Pb (x) = Pb+ eiKx + Pb eiKx + PbA eik1 x ,
(5.19)
with the three components obtained using Maple software as
gai k12
,
(K 2 k12 )
gKai k1 (eik1 L eiKL )
=
,
2 sinh(iKL)(K 2 k12 )
gKai k1 (eik1 L eiKL )
,
=
2 sinh(iKL)(K 2 k12 )

PbA =

(5.20)

Pb+

(5.21)

Pb

(5.22)

81

Chapter 5 One dimensional model for bulges propagation in the duct

The bulge wave pressure is seen to be made of three distinct wave components:
one travelling from stern to bow with wavenumber K (Pb+ ), one from bow to stern
with wavenumber K (Pb ) and one from bow to stern with wavenumber k1 (PbA ).

Figure 5.2: Theoretical components of the wave pressure in the rubber duct
for a static internal pressure head of 7cm corresponding to a free bulge speed of
3.06m/s in the configuration of chapter 6.

The bulge pressure at the stern can also be expressed using Maple as
Pb (x)x=L =

gai k1 (iK iCL Keik1 L + k1 SL eik1 L )


,
SL (K 2 k12 )

(5.23)

with CL = cos(KL) and SL = sin(KL).


Providing the loss angle is small, resonant peaks are expected in the following
cases:
. k1 = k2 or in other words, the water wave speed matches the free bulge
speed, or
. k2 = n/L (n = 1, 2, ...) corresponding to the different harmonics of the
membrane.

82

Chapter 5 One dimensional model for bulges propagation in the duct

The magnitudes of the three components (5.20)-(5.22) are plotted in Figure 5.2
for the same range of experimental frequencies used in chapter 6. Resonant peaks
are observed for frequencies corresponding to the above second case. A better
understanding of this will be given in 6.4.2.

Figure 5.3: Theoretical internal velocities for the first four harmonics at 30
instants over one wave period for the duct with an internal static pressure head
of 7cm; (a) f =0.306Hz; (b) f =0.612Hz; (c) f =0.918Hz (d) f =1.224Hz.

Chapter 5 One dimensional model for bulges propagation in the duct

83

Figure 5.4: Theoretical internal pressure for the first four harmonics at 30
instants over one wave period for the duct with an internal static pressure head
of 7cm; (a) f =0.306Hz; (b) f =0.612Hz; (c) f =0.918Hz (d) f =1.224Hz.

For these resonant peaks, the magnitudes of Pb and Pb+ , two waves travelling in
opposite directions at the same speed, are the same, leading to the presence of
a standing wave. For those frequencies, the total pressure is then seen to be the
sum of a standing wave and a travelling wave PbA , making it a partial standing
wave. The characteristic patterns of these different harmonics can be seen in

84

Chapter 5 One dimensional model for bulges propagation in the duct

Figures 5.3 and 5.4, where the theoretical internal velocity and pressure in the
duct, respectively, are plotted at 30 different instants over one wave period, for
the configuration of the rubber duct with an internal static pressure head of 7cm
corresponding to a free bulge speed of 3.06m/s.

5.5
5.5.1

Duct with a power take-off system


Bulge pressure

When including a dashpot Power Take-Off (PTO) at the stern, the boundary
condition (5.17) remains the same whereas (5.18) changes to
Pb + Pw = iC

d
(Pb + Pw ),
dx

(5.24)

where the dashpot rate is CSd .


The idea used in this study is that of a bent tube fixed at the ducts stern (Figure
5.5), containing a slug of water as Chaplin et al. (2012). The internal horizontal
flow at the stern is transferred into the vertical motion of the slug of water. Similar
to what happens in an oscillating column, the vertical motion of this water surface
is used to push the air above it through a PTO. This PTO has been detailed in
section 4.5.
Theoretically, consider the slug of water of length l connected at the ducts stern,
where the turbine is modelled as the PTO described above and shown in Figure
5.5. From the momentum equation,
p1 = p2 + gh2 + l

du2
.
dt

(5.25)

The conservation of the flow rate yields


u2 = u1,

(5.26)

where = Sd /St , the ratio of the duct cross-section to that of the bent tube.

Chapter 5 One dimensional model for bulges propagation in the duct

85

2)

h2

l
1)

stern x=L
Figure 5.5: Definition sketch of the stern.

Differentiating (5.25) with respect to time and using the harmonic motion notation yields
P1 = P2 +

g dh2
+ilU1
i |{z}
dt

(5.27)

u2

yielding

P1 = P2 + (

dP1
g
l)
.
2

dx |x=L

(5.28)

Recalling that the impedance of the PTO ZPTO is the ratio (pressure/flow rate)
and that the duct impedance is
c
Zd = ,
(5.29)
Sd
the pressure at 2) can be written as
p2 = St u2 ZPTO = St u1ZPTO
ZPTO
Zd
= Sd u 1
Z
| {zd }
Z

c
= Sd u 1 Z
Sd
= Zcu1 .

(5.30)

86

Chapter 5 One dimensional model for bulges propagation in the duct

Using (5.12), this yields


i dP1
dx |x=L
i dP1
Z
.
=
k2 dx |x=L

P2 = ZcU(x = L) = Zc

(5.31)

Substituting (5.31) into (5.28) and substituting g/ 2 = 1/k1 tanh k1 h leads the
boundary condition at the stern x = L



 d
iZ
1
Pb + Pw =
+
l
(Pb + Pw ).
k2
k1 tanh k1 h
dx

(5.32)

In a similar way to the the closed end situations, it is possible to solve (5.11) with
the boundary conditions of zero velocity at the bow (5.17) and with the dashpot
condition at the stern (5.32). The resulting bulge pressure Pb (x) can be written
as the sum of three wave components as in (5.19), with the three components
obtained with Maple as
gai k12
,
(K 2 k12 )

gai K Kk1 C(eik1 L eiKL ) Keik1 L k1 eiKL
,
=
2(K 2 k12 )(iCSL K + CL )

gai K Kk1 C(eik1 L eiKL ) Keik1 L + k1 eiKL
,
=
2(K 2 k12 )(iCSL K + CL )

PbA =

(5.33)

Pb+

(5.34)

Pb

(5.35)

with C from (5.24) being






Z
1
C=
i
l .
k2
k1 tanh k1 h

(5.36)

For practical reasons, the impedance of the PTO cannot be completely equal to
the ducts impedance. The same remark applies for the ducts cross-section being
different to that of the rigid tube containing the slug of water. This formula takes
those mismatches into account, in the Z and coefficients for the impedance and
cross-section, respectively.
It is interesting to notice that having C in (5.24) corresponds to the
closed ends configuration. Making C in (5.33)-(5.35) and recalling that
sinh(iKL) = i sin(KL) yields the obtained respective components for the closed
ends as in (5.20)-(5.22).

87

Chapter 5 One dimensional model for bulges propagation in the duct

5.5.2

Power calculation

As for the closed ends case, the bulge pressure at stern can also be expressed with
Maple as
gai
Pb (x)x=L =



K 2 (k1 C 1) + k12 CL eik1 L Kk1 C K ik1 SL eik1 L
(K 2 k12 )(iCSL K + CL )

(5.37)

while the internal velocity at the same position is

U(x)x=L =




gai K 2 eik1 L iKSL + k1 CL k1
(K 2 k12 )(iCSL K + CL )

(5.38)

The flow rate (4.26) of the air flowing in the pipes is then expressed as
q0 = (St )e0
where
=

(5.39)

,
cos eq + sin eq

(5.40)

and e0 the magnitude of the complex surface displacement e obtained from


e=

i
u2 .

(5.41)

The velocity u2 is calculated from (5.26) and (5.38). This leads the expression of
the (complex) flow rate as

q=




St igai K 2 eik1 L iKSL + k1 CL k1
(K 2 k12 )(iCSL K + CL )

(5.42)

The power absorbed by the linear dashpot is then calculated as


1
Pf = ZPTO |q|2 .
2

5.6

(5.43)

Conclusions

A model for the free bulge speed has been developed, based on a load-deflection
analysis of the rectangular membrane. The free bulge speed is seen to be a function of the membrane properties (Youngs modulus and dimensions). The change

88

Chapter 5 One dimensional model for bulges propagation in the duct

of membrane thickness on the free bulge speed due to its deflection was introduced
and will be evaluated with measurements in chapter 6.
Bulge propagation was explained with a one dimensional model. The closed ends
rubber duct is seen to be subject to resonant peaks, for which the bulge pressure
inside the duct take the form of a partial standing wave. This will be compared
and analysed with experimental values from pressure transducers measurement
and laser sensor analysis of the membrane displacement in chapter 6.
The power absorbed by the PTO, modelled as a linear dashpot in which air flows
loses its energy, can be calculated by taking into account the physical constraints
of the problem, namely mismatches of impedances between PTO and duct, and
different cross-sectional areas between the rigid tube and the duct. Moreover and
unlike previous work using this PTO, the air compressibility is now part of the
problem. Theoretical capture widths presented in chapter 7 are calculated with
(5.43) and compared with experimental data inferred from wave gauges and air
pressure transducers measurement detailed in section 4.5.

Chapter 6
Narrow flume experiments: static
deflection, free bulge speed and
pressure amplification
6.1

Introduction

This chapter describes the tests carried on the models with closed ends in the
narrow flume. For these tests, the total length of the duct was 5m. The static
membrane deflection, free bulge speed and pressure amplification beneath waves
were monitored with this configuration. The membrane displacement was also
measured with the laser sensors, and the pressure in the duct was deduced from
these measurements. Measured values are compared with the calculated ones from
the theory described in 5.2 and 5.4.

6.2

Experimental layout

In order for sufficient pressure variations to act on the membrane, the duct was
raised from the flume floor. For this reason a false seabed and a ramp were installed, as shown in Figure 6.1. The ramp and the artificial seabed were held in
place using suction lifters attached to the glass wall/floor of the flume. Pictures
of the complete set-up are given in Appendix C. The bottom part of the duct was
also drilled so it could be attached to the aluminum profiles at several location
along its length. This way, no undesirable motion of the ramp, aluminum profiles
89

Absorbing
beach
Wg3

6.01m

Pt2

0.9m

Wg1 Wg2
Wavemaker

0.5m

2.0m

1.0m

Pt1

5.0m
d[2] 0.16m
d[1]

90

Chapter 6 Narrow flume experiments: static deflection, free bulge speed and
pressure amplification

Figure 6.1: Experimental layout for the closed ends configuration. Dashed
lines correspond to different internal static pressure heads.

Chapter 6 Narrow flume experiments: static deflection, free bulge speed and
pressure amplification

91

or duct was observed when the system was beneath waves.


Two water pressure transducers, Pt1 and Pt2, were mounted inside the duct at its
bow and stern respectively. Two wave gauges Wg1 and Wg2 were installed ahead
of the model for the purpose of identifying incident and reflected waves by means
of the theory described for example by den Boer (1981). A third was installed
behind the model to measure transmitted waves. The distance d[2] between Wg1
and Wg2 was changed with frequency in order to avoid it being an integer multiple
of half the water wave length (den Boer (1981)). A hydraulic pump pressurised
the duct with water, maintaining the desired internal static pressure head and
making up for any leaks.
When the membrane displacement was to be measured, the laser sensors were
placed in rigid transparent boxes that were submerged enough to avoid any undesired diffraction of the laser beam by the free water surface. When they were
not used, a fourth wave gauge (not drawn in Figure 6.1) was placed in the middle
length of the duct for the purpose of evaluating the evolution of the wave acting
on the membrane.

6.3
6.3.1

Static tests: 1mm thick membrane


Membrane deflection

The static centreline deflection was measured for several internal pressure heads.
Measurements were carried out either manually, with a ruler measuring the membrane submergence (hence its deflection) at ten points along the length of the
membrane, or electronically with the laser sensors at six points. These points
were chosen not to be to close to the duct edges to be sure to measure the deflection on a flat portion of the membrane.
Averaged values for the deflection shown in Figure 6.2 reflect fair agreement with
the linearised form of (5.9).
As stated earlier, the point was to determine the membrane submergence in order to determine the dynamic wave pressure acting on it. The submergence of the
membrane was not the same away from its centreline so an equivalent submergence

92

Chapter 6 Narrow flume experiments: static deflection, free bulge speed and
pressure amplification

Figure 6.2: Experimental values of membrane centreline deflection (+ + +),


load-deflection relation 5.9 ()

d1

w1

b
a
Figure 6.3: Equivalent membrane submergence depth

depth d1 was defined. It was chosen so that the cross-section of the duct stayed
the same, meaning
2ab + S 0 = 2ab + 2aw1
S0
w1 =
,
2a
0

with the analogy shown in Figure 6.3 and S obtained from (5.4).

(6.1)
(6.2)

Chapter 6 Narrow flume experiments: static deflection, free bulge speed and
pressure amplification

6.3.2

93

Free bulge speed

Bulges were manually triggered by suddenly displacing the membrane at the bow.
They propagated in the duct and were reflected at the stern. Time series of pressure at the bow and at the stern, together with the displacement time series of the
laser sensors were used to measure the free bulge speed. An example of time series
from the two pressure transducers is given in Figure 6.4, where the bulge is created
at the bow, reflected at the stern and back to the bow. The free bulge speed was
then calculated knowing the time the bulge took to travel a known distance.
Data obtained are compared in Figure 6.5 with predicted values from the theory developed in section 5.2 for several internal static pressure heads in the duct,
corresponding to different circumferential stresses. Calculated and measured values follow the same trend and are in good agreement, although the latter ones
seem to be slightly higher (by up to 12%). This can be the result of a slight error
in the Youngs modulus measurement.

Figure 6.4: Time series of pressure transducers from free bulge speed measurement

94

Chapter 6 Narrow flume experiments: static deflection, free bulge speed and
pressure amplification

Figure 6.5: Free bulge speed for the 1mm thick membrane, each symbols
representing one set of measurement.

6.4

Membrane beneath waves: 1mm thick membrane

6.4.1

Experimental conditions

Two configurations of the duct with closed ends were studied with the test conditions being summarised in Table 6.1. Waves were generated in the range of
[0.7Hz-1.5Hz] with increment of 0.025Hz and wave amplitudes varied from [0.1m0.04m] with increment of 0.0025m. Each configuration was tested a total of three
times, once without the laser sensors and twice with the laser sensors in place (at
different locations between the two runs).
Internal static
pressure head

Wave
frequencies

Number of
wave

Number
of

(m)

(Hz)

amplitudes

runs

0.07m

0.7-1.5

13

0.1m

0.7-1.5

13

Test no

Table 6.1: Closed ends experiment: test conditions

Chapter 6 Narrow flume experiments: static deflection, free bulge speed and
pressure amplification

Figure 6.6: Time series of laser sensors when rubber membrane is beneath
waves.

95

96

Chapter 6 Narrow flume experiments: static deflection, free bulge speed and
pressure amplification

Figure 6.7: Time series of pressure transducers (at bow and stern) when
rubber membrane is beneath waves.

Example of time series from the laser sensors and the pressure transducers are
given in Figures 6.6 and 6.7, respectively. The signals are predominantly simple
harmonic, hence the data analysis of other frequencies were initially neglected.
This was justified also by measuring reflected and transmitted wave power with
up to three frequency components (see section 6.4.5) where it was seen that the
power carried by higher harmonics was negligible. The noise on the signals from
the laser sensors in Figure 6.6 are characteristic of ripples propagating on the
membrane. The effect of these ripples will be discussed in section 6.5.

6.4.2

Pressure amplification

Pressure recordings at the stern made it possible to express the total internal pressure amplitude at the stern as a linear function of the pressure due to the undisturbed incident wave at the mean elevation of the membrane, |P (x)|x=L /gai ,
where the submergence d for ai was taken as d1 .
Figures 6.8 and 6.9 show this ratio for internal static pressure heads of 0.07m and
0.1m, corresponding to a natural free bulge speeds in the membrane of 3.06m/s
and 3.28m/s, respectively. Experimental values are compared with the theoretical pressures at the stern calculated from (5.23) (on which the wave pressure was
added).

Chapter 6 Narrow flume experiments: static deflection, free bulge speed and
pressure amplification

Figure 6.8: Magnitude of the pressure at the stern relative to the wave pressure
acting on the membrane with an internal static pressure head of 7cm. symbols:
experimental values; : theory.

Figure 6.9: Magnitude of the pressure at the stern relative to the wave pressure
acting on the membrane with an internal static pressure head of 10cm. symbols:
experimental values; : theory.

97

98

Chapter 6 Narrow flume experiments: static deflection, free bulge speed and
pressure amplification

Resonance peaks are observed for the frequencies set out in Table 6.2. These
resonant frequencies are compared with the theoretical resonant frequencies from
the bulge pressure maxima (5.23) for the case k2 = n/L (n = 1, 2, ...).

Internal static
pressure head
of 7cm (c=3.06m/s)

fn

n=3
n=4

0.918Hz 0.95Hz
1.224Hz 1.2Hz

3.4%
1.9%

n=5

1.53Hz

1.425Hz

6.8%

0.984Hz 1.075Hz
1.312Hz 1.325Hz

9.2%
0.9%

fres

k(fn fres )/fn k

Internal static
pressure head
of 10cm (c=3.28m/s)
n=3
n=4

Table 6.2: Comparison of experimental resonant frequencies fres with the


theoretical ones calculated as fn = nc/2L where c is the theoretical free bulge
speed for the corresponding internal static pressure heads.

Experimental and theoretical resonant frequencies seem to be in fair agreement,


and discrepancies of less than 10% are noticed. Those are possibly due to the
practical difficulty of maintaining a constant pressure (hence a constant free bulge
speed c) in the duct, yielding resonant frequencies that were not entirely equally
spaced. This is also consistent with the results from the free bulge speed measurements in section 6.3.2 where the measured free bulge speeds were higher than
the predictions. Despite this, the repetitive occurrence of resonance in the duct is
fairly well predicted, with respect to the frequency.
The theoretical amplitudes of the total internal pressure were calculated using
(5.23) and the wave pressure acting on the membrane. Discrepancies are noticed
between those calculated values and the measured ones at the stern. This can be
the result of the omission in the model of the interaction of radiated waves by the
membrane with the incoming waves.
Despite those two main discrepancies, the general trend of the pressure variation at the stern is seen to be fairly well predicted with the model developed in

Chapter 6 Narrow flume experiments: static deflection, free bulge speed and
pressure amplification

99

chapter 5. The resonant frequencies can be calculated in the same way as the
case of a stretched rope, on which an impulse will travel along the rope at a speed
depending on the tension in it. For those frequencies, the rope exhibits node and
antinodes characteristic of standing waves (or partial standing waves). Studying
the membrane displacement, related to the internal pressure variation, helps to
explain these observations. This was not done here.

6.4.3

Membrane displacement and pressure in the duct

Laser sensors recorded membrane displacements beneath waves at 11 locations


along the membrane, w being now a function of time. Using again the linearised
form of (5.9), the total internal pressure was inferred at those points from this
measurement as

w(x, t)
2v0
y,0 (
),
(6.3)
a
a
the membrane deflection being due in that case to pb (x, t) + pw (x, t) (and not just
p(x, t)

pb (x, t) as in the static case). The stress y,0 was due to the static deflection from
the internal static pressure. A spline interpolation passing through the measurement points was used to smooth out the line.
The total pressure in the rubber duct (bulge pressure and wave pressure) is plotted in Figures 6.10 for an internal static pressure head of 7cm corresponding to a
theoretical free bulge speed of c =3.06m/s. The wave frequency was f = 0.95Hz,
corresponding to the first peak observed in Figure 6.8 for this configuration.
Theoretical and experimental values are in fair agreement. Repetitive patterns
are observed, and measurements are seen to agree in shape and amplitude with
the predictions. Those patterns are characteristics of a standing wave - or as will
be justified by separating the three bulge components later, a partial standing
wave. The number of repetitive patterns increases with frequency as seen in Figure 6.11 where the total internal pressure is plotted for the second peak observed
in Figure 6.8 (f = 1.2Hz) for the same internal static pressure head and free bulge
speed.

Chapter 6 Narrow flume experiments: static deflection, free bulge speed and
100
pressure amplification

Figure 6.10: Pressure in the duct at thirty instants over one wave period with
an internal static pressure head of 7cm; wave frequency f =0.95Hz; (a) theory;
(b) inferred from laser sensors measurement of membrane displacement.

Figure 6.11: Pressure in the duct at thirty instants over one wave period with
an internal static pressure head of 7cm; wave frequency f =1.2Hz; (a) theory;
(b) inferred from laser sensors measurement of membrane displacement.

Chapter 6 Narrow flume experiments: static deflection, free bulge speed and
pressure amplification
101
For those frequencies close to the resonant frequencies fn , the wavelength of the
bulge wave is seen to be an integer multiple of the length of the duct, agreeing
with all the above regarding standing/partial standing waves. When the wave
frequency is between two resonant frequencies, this is not the case as can be seen
in Figure 6.12 where the total internal pressure is plotted for f = 1.1Hz for the
same conditions as above and in this case, measurements and predictions do not
agree as well as before.

Figure 6.12: Pressure in the duct at thirty different instants over one wave period with an internal static pressure head of 7cm; wave frequency f =1.1Hz; (a)
theory; (b) inferred from laser sensors measurement of membrane displacement.

The same total pressure mapping is given in Figure 6.13 for the rubber duct with
an internal static pressure head of 10cm corresponding to a theoretical free bulge
speed of c =3.28m/s and a wave frequency of f =1.075Hz. This corresponds to
the first peak observed for this configuration as seen in Figure 6.9. This peak was
earlier seen to be shifted up towards higher frequencies due to experimental free
bulge speeds being higher than the predicted ones. This is why the theoretical
total internal pressure was calculated with a free bulge speed of c =3.48m/s. In
the same way as for the other studied configuration, the total internal pressure is
seen to exhibit similar repetitive patterns at those frequencies and the measured
values are in good agreement, in shape and amplitude, with the calculated ones
using experimental values of free bulge speed.

Chapter 6 Narrow flume experiments: static deflection, free bulge speed and
102
pressure amplification

Figure 6.13: Pressure in the duct at thirty instants over one wave period with
an internal static pressure head of 10cm; wave frequency f =1.075Hz; (a) theory;
(b) inferred from laser sensors measurement of membrane displacement.

The total internal pressure inferred from the laser sensor measurements of the
membrane displacement is seen to be in fair agreement with the calculated values
from the theory. To go further in the verification of this theory, it is necessary
to verify that this computed total internal pressure is actually made up of the
three components of the bulge wave introduced in section 5.4 (on which the wave
pressure has to be added to). The following task was to separate those components
and reconstruct the total internal pressure in the duct as in Figures 6.10-6.12 by
summing up these three waves.

6.4.4

Bulge pressure components

6.4.4.1

Separation of components

Laser sensors recorded membrane displacements at 11 points. The displacement


(hence the pressure in the duct) was found to be predominantly harmonic, at the
wave frequency. Using an FFT on the obtained signal for the three points shown

Chapter 6 Narrow flume experiments: static deflection, free bulge speed and
pressure amplification
103
LS[1]

LS[2]

LS[3]

Rubber
membrane

x1
x2
Figure 6.14: Separation of bulge wave components from laser measurements.

in Figure 6.14, the total internal pressure p of those points could be written as
pLS[j] = ALS[j] cos t + BLS[j] sin t j = 1, 2, 3...11.

(6.4)

As introduced in section 5.4, the total pressure in the duct is expected to be


p
made of one wave of amplitude Pb+ = a21 + b21 travelling from stern to bow with
p
complex wavenumber K, one wave of amplitude Pb = a22 + b22 travelling from
p
bow to stern with complex wavenumber K, one wave of amplitude PbA = a23 + b23
travelling from bow to stern with k1 . Using these notations
pj =

a1 cos(Kxj1 t) + b1 sin(Kxj1 t) + a2 cos(Kxj1 + t) +


b2 sin(Kxj1 + t) + a3 cos(k1 xj1 t) + b3 sin(k1 xj1 t). (6.5)

Setting equal those pressures at each point and separating cosine and sine terms
leads to a linear system of 6 equations and 6 unknowns (a1 , b1 , a2 , b2 , a3 , b3 ). The
signal from the laser sensors was pre-processed in order to subtract the wave pressure from the total internal pressure, hence dealing with only the bulge pressure.
The first measurement point was taken as the origin for the phase calculation.
Several combinations of (LS[1],LS[2],LS[3]) triplets could be used to separate the
three absolute amplitudes of the bulge pressure component.
This technique can be successfully applied and amplitudes of each components
can be obtained for specific triplets of points. However, those solutions are sought
with constant amplitude all over the length of the membrane as seen in (5.20)(5.22). This means that the obtained values should not depend on the chosen
triplet.

Chapter 6 Narrow flume experiments: static deflection, free bulge speed and
104
pressure amplification

The idea is then to write the system obtained for all the points j. That yields an
overdetermined system (6 unknowns for 22 equations if 11 points are taken). Such
systems have usually no exact solution and a best fit solution can be obtained
using a least squares method. In practice, this means that the obtained solution
is that which minimises the sum Sq of the squares of the errors at each point of
measurement j as
Sq =

11
X
j=1

(pLS[j] p(j))2 .

(6.6)

This was solved with an iterative method using Maple software.

6.4.4.2

Components at resonance

Using this technique, the three components were obtained for the rubber membrane with an internal static pressure head of 7cm for the wave frequency f =
0.95Hz corresponding to the first experimental resonant peak observed for this
configuration.
Figure 6.15 shows the measured amplitude and phase of the total pressure in
the duct and compares them with the values of amplitude and phases from the
total internal pressure along the duct calculated with (5.19), the different pressure
components being those computed from the least squares method. The magnitudes of pressure in the duct seem to be in fair agreement. The repetitive patterns
observed in Figure 6.10 are confirmed here. Measured phases do not seem in
perfect agreement with the computed ones. However, the kind of nodes visible in
Figure 6.10 can also be characterised here as each of them corresponds to the phase
crossing = . This change of phase is similar between measured and computed
values, the difference in magnitude of the phase affecting only the amplitude of
the instantaneous pressure.

Chapter 6 Narrow flume experiments: static deflection, free bulge speed and
pressure amplification
105

Figure 6.15: Measured total internal pressure, normalised by the pressure


acting on the membrane (a) and phase (b) for the duct with an internal static
pressure head of 7cm, plotted as points. Lines represent amplitudes and phases
computed as the sum of the components obtained from the least squares method.
Wave frequency f =0.95Hz.

Figure 6.16: Pressure in the duct at thirty different instants over one wave period with an internal static pressure head of 7cm; wave frequency f =0.95Hz; (a)
inferred from laser sensor measurement of membrane displacement amplitude;
(b) sum of the three bulge wave components (and the wave pressure) obtained
from a least squares method.

Chapter 6 Narrow flume experiments: static deflection, free bulge speed and
106
pressure amplification
Figure 6.16 shows those two ways of mapping the total internal pressure over
several instants of one wave period: the first one obtained from the laser sensor
measurement as before, and the second one by reconstructing this total internal pressure by summing up the three bulge wave components (amplitudes and
phases) obtained at any point on the membrane. The two obtained mappings
seem to agree with this approach of decomposing the bulge pressure as the sum
of three components and the slight mismatch of measured and computed phases
seen in Figure 6.15 is consistent with Figure 6.16.
Finally, the pressure amplitudes are compared with those predicted from (5.20)(5.22) for the same configuration and results are shown in Figures 6.17, 6.18 and
6.19. Measured and calculated results similarities around the first resonant peak
f =0.95Hz are in quite fair agreement with respect to their amplitudes. Keeping
in mind that those amplitudes were obtained from a least squares method, it is
quite satisfactory.

Figure 6.17: Magnitude of Pb+ relative to the wave pressure acting on the
membrane with an internal static pressure head of 7cm. symbols: experimental
values inferred from a least squares method; : calculated from (5.21).

Chapter 6 Narrow flume experiments: static deflection, free bulge speed and
pressure amplification
107

Figure 6.18: Magnitude of Pb relative to the wave pressure acting on the


membrane with an internal static pressure head of 7cm. symbols: experimental
values inferred from a least squares method; : calculated from (5.22).

Figure 6.19: Magnitude of PbA relative to the wave pressure acting on the
membrane with an internal static pressure head of 7cm. symbols: experimental
values inferred from a least squares method; : calculated from (5.20).

Chapter 6 Narrow flume experiments: static deflection, free bulge speed and
108
pressure amplification
The same idea has also been verified with the duct with an internal static pressure
head of 10cm and the first observable resonant frequency f =1.075Hz from Figure
6.9. Figure 6.20 shows the measured amplitude and phase of the total pressure in
the duct and compares it with the values of amplitude and phases computed as
before from the least squares method while Figure 6.21 shows the total internal
pressure inferred from the laser measurements and the total internal pressure reconstructed by summing the different components obtained from the least squares
method. Finally Figures 6.22, 6.23 and 6.24 compare the computed amplitudes of
the three components of the bulge wave with those computed from (5.20)-(5.22)
for this condition.

Figure 6.20: Measured total internal pressure, normalised by the pressure


acting on the membrane (a) and phase (b) for the duct with an internal static
pressure head of 10cm, plotted as points.

Lines represent amplitudes and

phases computed as the sum of the components obtained from the linear squares
method. Wave frequency f =1.075Hz.

Chapter 6 Narrow flume experiments: static deflection, free bulge speed and
pressure amplification
109

Figure 6.21: Pressure in the duct at thirty instants over one wave period
with an internal static pressure head of 10cm; wave frequency f =1.075Hz; (a)
inferred from laser sensor measurement of membrane displacement amplitude;
(b) sum of the three bulge wave components (and the wave pressure) obtained
for a least squares method.

Figure 6.22: Magnitude of Pb+ relative to the wave pressure acting on the
membrane with an internal static pressure head of 10cm. symbols: experimental
values inferred from a least squares method; : calculated from (5.21).

Chapter 6 Narrow flume experiments: static deflection, free bulge speed and
110
pressure amplification

Figure 6.23: Magnitude of Pb relative to the wave pressure acting on the


membrane with an internal static pressure head of 10cm. symbols: experimental
values inferred from a least squares method; : calculated from (5.22).

Figure 6.24: Magnitude of PbA relative to the wave pressure acting on the
membrane with an internal static pressure head of 10cm. symbols: experimental
values inferred from a least squares method; : calculated from (5.20).

Chapter 6 Narrow flume experiments: static deflection, free bulge speed and
pressure amplification
111

6.4.5

Wave analysis and loss in the rubber

Signals of the three wave gauges were used to measure the incident (Pi ), reflected
(Pr ) and transmitted (Pt ) wave powers. Figures 6.25 and 6.26 show the ratio of
the transmitted and reflected wave power to the incident wave power.
The transmitted wave coefficient increased with frequency. The higher the wave
frequency, the less the membrane is affected by the incident wave, hence most of
the wave power is transmitted behind the rubber duct. Peaks of minimum transmission occurs for f = 0.95Hz and f = 1.2Hz that were seen earlier to be resonant
frequencies of the system.

Figure 6.25: Ratio of the transmitted wave power, Pt , to the incident wave
power Pi for the duct with an internal static pressure head of 7cm. Black line
and symbols: simple harmonic component of wave gauge signals; Grey line and
symbols: first three harmonic components of wave gauge signals.

Chapter 6 Narrow flume experiments: static deflection, free bulge speed and
112
pressure amplification

Figure 6.26: Ratio of the reflected wave power, Pr , to the incident wave
power Pi for the duct with an internal static pressure head of 7cm. Black line
and symbols: simple harmonic component of wave gauge signals; Grey line and
symbols: first three harmonic components of wave gauge signals.

In contrast, the reflected ratio seems to exhibit the peaks at the resonant frequencies. This is due to maximum membrane deflection at those frequencies, triggering
radiated waves. However, this reflected ratio is seen to be very low. Those peaks
of maximum reflection and minimum transmission at the resonant frequencies are
characteristic of such system, as developed in section 3.3.4.
The energy in the system should be conserved. In theory, this means
Pt + Pr = Pi .

(6.7)

However, losses can occur for various reasons as described later and a power loss
Ploss is defined as
Ploss = Pi Pr Pt
and is plotted in Figure 6.27.

(6.8)

Chapter 6 Narrow flume experiments: static deflection, free bulge speed and
pressure amplification
113

Figure 6.27: Ratio of the wave power loss, Ploss , to incident wave power Pi
for the duct with an internal static pressure head of 7cm. Line and symbols:
simple harmonic analysis of wave gauge signals.

It can be seen that about 50% of the incident wave power is dissipated in the
system over one wave period. Here is a explanation of the possible causes of this
loss:
Power in higher wave harmonics- The power loss shown in Figure 6.27 was calculated using only the first harmonic of the signals from the wave gauges. Some
power is often expected to be found in higher harmonics, especially in the transmitted power with such a step configuration. A description of this phenomenon is
given by Massel (1983), for instance. This was not the case here, as can be seen
in Figure 6.25 where the transmitted wave power was calculated using two more
harmonics. Some power was found in the higher harmonics of the reflected wave
but not enough to explain the magnitude of lost power.
Energy dissipated in the flow - A portion of the incident wave energy can be lost
through the following processes: wave breaking over the structure, turbulence in
the inner flow or around the corners which were present in this configuration.
Typical values of 50% of incident energy dissipated are quite common for those
type of configuration as seen in Table 3.1.

Chapter 6 Narrow flume experiments: static deflection, free bulge speed and
114
pressure amplification
Loss in the rubber - The membrane displacement beneath waves stretches the rubber. This oscillatory stretching is characterised by some loss of the energy required
for the deformation into heat (hysteresis) in the rubber.
Consider the strip of membrane (length L, thickness v0 and width 2a) undergoing oscillatory loading in its length direction, so that the longitudinal strain is
 = 0 sin(t). Following the stress-strain relationship (2.24), the stress in the
rubber is
= E0 (sin(t) + cos(t)),
(6.9)
yielding the force in the rubber
F = 2av0
= 2av0 E0 (sin(t) + cos(t)).

(6.10)
(6.11)

The mean power loss due to hysteresis is then calculated as the mean product of
the force and the velocity of the point of action of the force as
Phys

Z
d
1 T
(2av0 E0 (sin(t) + cos(t)))( L)dt
=
T 0
dt
1
Vr E20 2,
=
2

(6.12)
(6.13)

with Vr = v0 2aL being the total volume of rubber.


An intuitive approach would consist in mapping the membrane deflection at several
instants over one wave period, as previously done for the pressure, and measure
the arc-length of the envelope of the obtained plot. However, this would be justified if the pressure wave in the duct was a pure standing wave. It was seen in
section 6.4.4.2 that this is not the case, and that at best, the pressure wave is a
partial standing wave at resonance only.
Instead, the instantaneous elongation was plotted at twenty instants over one wave
period, as seen in Figure 6.28. It allowed the measurement of 0 , the amplitude
of the elongation with respect to the incident wave amplitude. It is interesting to
note that the frequency of membrane stretching over one wave period is twice the
wave frequency, and this should be considered accordingly in (6.13), where is
the frequency of oscillation of the strip of rubber and not the wave.

Chapter 6 Narrow flume experiments: static deflection, free bulge speed and
pressure amplification
115

Figure 6.28: Instantaneous elongation in the rubber at 20 instants (plotted as


points) as a function of time normalised by the wave period T . Wave frequency
is f = 0.775Hz.

Figure 6.29: Ratio of the power lost in the rubber, Ploss , to the incident wave
power Pi for the rubber duct with an internal static pressure head of 7cm.

Figure 6.29 shows the ratio of the mean power loss due to hysteresis in the rubber to
the incident wave power. Two distinctive peaks are observed and correspond once
again to the resonant frequencies. At those resonant frequencies, the amplitude

Chapter 6 Narrow flume experiments: static deflection, free bulge speed and
116
pressure amplification
of membrane deformation is maximum, yielding a maximum elongation of the
rubber. However, the magnitude of this power loss is seen to be negligible in the
concern of power conservation in this system.

6.4.6

Transverse motion of the membrane

The membrane motion in its length is characterised by characteristics patterns, in


a similar way as for the velocity profile in Figure 5.3 for instance. At resonance,
an integer multiple number of complete patterns are visible, while this number of
patterns goes from one integer to the successive one between resonance.
A concern can arise with regard to the motion of the membrane in its width
direction, namely the general form of deformation beneath waves. Consider the
cross-sectional view of the duct in the narrow flume as seen in Figure 6.30. A
close configuration was studied by Bauer (1981) who investigated the hydroelastic
vibrations of a liquid in a rectangular container, with an elastic bottom and a
free liquid surface. He also studied the hydroelastic vibrations of a liquid in a
rectangular container covered by a flexible membrane. The idea here is to use a
similar method by adapting the three dimensional potentials he used for the two
dimensional configuration shown in Figure 6.30. The potential in the upper region
is sought to satisfy the Laplace equation, the zero velocity on the flumes wall and
the free surface boundary condition. It is then expressed as
2 =

X
n

cosh[n (z d)] +


ny it
2
sinh[n (z d)] An cos
e
, (6.14)
gn
2a
(6.15)

where n = n/2a. The potential in the lower region is sought to satisfy the
Laplace equation, together with the zero velocity on the ducts wall and bottom,
yielding
1 =

X cosh[n (z + b)]
n

cosh(n b)

Bn cos

ny it
e
.
2a

(6.16)

Consider the displacement of the membrane as


w(y, t) =

X
n

Wn (t) sin(

ny
).
2a

(6.17)

Chapter 6 Narrow flume experiments: static deflection, free bulge speed and
pressure amplification
117

flume
walls

(2)
h

(0,0)

duct
(1)

2a
Figure 6.30: Cross-section view of the duct and transverse motion of the
membrane.

The kinetic energy of the membrane associated with this displacement is


Ek

Z
m 2a  w 2
=
dy
2 0
t
m a X 2
Wn ,
=
2 n

(6.18)
(6.19)

with m the mass per unit area of the membrane. The potential energy associated
with this displacement is
Ep

Z
1 2a  w 2
Ty
dy
=
2 0
y
a X n2 Ty 2 2
Wn ,
=
8 n
a2

(6.20)
(6.21)

from which we can deduce the Lagrangian of the system L = Ek Ep . The loading
on the membrane due to the pressure p(y, t) makes it possible to determine the

Chapter 6 Narrow flume experiments: static deflection, free bulge speed and
118
pressure amplification
generalised force Qn with the help of the virtual work concept,
W =

XZ
n

2a

ny
p(y, t) sin(
)dy Wn .
2a
{z
}

(6.22)

Qn

The equation of motion of the membrane is then

d L 
L
= Qn .

n
dt W
Wn

(6.23)

The pressure acting on the membrane is


1 2
p(y, t) = (

) .
t
t z=0

(6.24)

Using the continuity of the velocities /y at the interface z = 0 to express An


as a function of Bn and using
cos(

X
jy
ky
)=
jk sin(
)
2a
2a
j

(6.25)

from Bauer (1981) with


jk =
(6.23) yields

Substituting

2 j[1 (1)j+k ]
,

j 2 k2

(6.26)

2i X
n[1 (1)n+k ]
2

Wn +
n Wn =
Bk k
.
k
n2 k 2
Wn = Cn eit

(6.27)

(6.28)

in (6.27) and the kinematic free surface condition


w
1
=

t
z z=0

(6.29)

and setting equal the two obtained Cn leads to


X n[1 (1)n+k ]  2(S)
k

n2 k 2

(
n2 2 ) +


k Bk = 0,

(6.30)

with
n2 corresponding to the natural frequencies of the stretched membrane in
air, that is

n2 =

2 Ty n2
.
4m a2

(6.31)

Chapter 6 Narrow flume experiments: static deflection, free bulge speed and
pressure amplification
119
Furthermore
2(S)

= gk tanh(k h1 ),




k g sinh k (b + d) 2 cosh k (b + d)

 .
=
cosh(k b) 2 cosh(k d) sinh(k d)

(6.32)
(6.33)

Solving the determinant of (6.30) yields the natural frequencies of the fluid-membrane
system, with the effect of the surrounding water. The formulation of the membrane
displacement does not verify the conservation of volume, in term of
Z

2a

wdy 6= 0.

(6.34)

However, this is not a requested condition as here the concern is about a crosssection of the duct. If at this cross-section the membrane deflects in a way, further
down (or up) the duct, it deflects in a way so that the volume is conserved.
The effect of the tension in the length direction (xdirection) is also neglected considering the dimensions of the problem, the length of the duct being way greater
that its width. Considering the length of the membrane L (hence modes in the
length direction), this all comes down to saying that
2  n2
2
m2 
n2
Ty 2 + Tx 2
Ty 2 ,
m
4a
L
m 4a

(6.35)

meaning that the obtained frequencies would be slightly modified if considering


the 3D problem, but not in a significant way.
With the parameters from the experiment (a=0.15m, Ty = 170.9N/m, b = 0.185m,
and taking here d = d1 =0.171m, m = 1.14kg/m2 ) and the physical values
(g=9.81m/s2 , =1000kg/m3 ), the first three resonant frequencies of the system
are found to be
f1 = 1.615Hz,

(6.36)

f2 = 2.279Hz,

(6.37)

f3 = 2.793Hz.

(6.38)

Given the range of wave frequencies used ([0.7Hz:1.5Hz]), it seems reasonable to


say that higher modes than the fundamental ones are unlikely to be excited. The
).
membrane shape in the y-direction is predominantly made of the form sin( y
2a

Chapter 6 Narrow flume experiments: static deflection, free bulge speed and
120
pressure amplification
This is not a circular shape as introduced in chapter 5, although it was checked
that using a sine shape for the membrane deflection does not make much difference
in the distensibility and free bulge speed calculation.

6.4.7

Conclusions

The duct was placed beneath waves and pressures at its bow and its stern were
recorded. It exhibited resonant peaks for frequencies at which k2 = n/L as predicted by the theoretical model. Pressure amplifications were found to be up to
2.7 times the wave pressure acting on the membrane (or to be precise, the wave
pressure acting at the equivalent submergence depth defined from the static tests)
for those frequencies.
The total internal pressure in the duct could be inferred from the laser sensors
measurement of the membrane displacement. It was plotted over the length of the
membrane for 30 instants over one wave period. For the frequencies mentioned
above, these plots exhibit nodes and anti-nodes. By expressing the bulge wave
pressure in the duct as the sum of three components, and separating them using
a least squares method, the wave inside the duct at those frequencies was seen
to be made of two components of almost similar amplitude travelling in opposite
directions and one component travelling with the wavenumber of the water wave.
The characteristic patterns observed from the total internal pressure mapping were
then proved to be characteristics of a partial standing wave.
Loss in rubber is one main concern when dealing with this material. The mean
power loss due to hysteresis was found to be noticeable only at these resonant
frequencies at which the displacement (hence elongation) of the membrane is maximum.

6.5
6.5.1

Thin membrane results


Introduction

The work described above concerned the experimental work undertaken with the
1mm thick membrane (hereinafter referred to as thick membrane). The 0.2mm
thick membrane (hereinafter referred to as thin membrane) was also tested using

Chapter 6 Narrow flume experiments: static deflection, free bulge speed and
pressure amplification
121
the same experimental set-up. Unlike the thick membrane, there was no agreement
between the one dimensional theory for bulge propagation described in section 5.4.
This section present the results obtained for the static deflections, free bulge speed
and pressure amplification and tries to give an explanation of the discrepancies
between the calculation and measurements, and then discuss the applicability of
the one dimensional model used so far.

6.5.2

Experimental conditions

The tests conditions are the same as before. Due to its thickness, this membrane
was initially stretched in both directions. This is summarised in Table 6.3.

Test no

Strain in
Strain in
y-direction x-direction
y,0
x,0

Internal static
pressure head
(m)

Number
of
runs

0.14

0.02

0.14

0.05

0.28

0.02

0.28

0.05

0.14

0.14

0.02

0.14

0.14

0.05

Table 6.3: Experimental conditions for the closed ends experiment with the
0.2mm membrane.

6.5.3

Static deflection and free bulge speed

Before running waves, static tests were carried out as before to measure the centreline deflection w and free bulge speeds, and to compare the obtained results
with the calculated ones. The centreline deflection is shown in Figures 6.31, 6.32
and 6.33 for the three configurations tested.
Deflection values are seen to be of a higher magnitude for the 0.2mm membrane, as
it is less resistant to stretching compared to the 1mm rubber sheet. Experimental
and theoretical values follow the same trend with fair agreement, and the influence
of x,0 (hence x,0 ) is negligible, as can be seen in (5.9), where the influential stress
is in the y-direction.

Chapter 6 Narrow flume experiments: static deflection, free bulge speed and
122
pressure amplification

Figure 6.31: Experimental values of membrane centreline deflection (+ + +),


load-deflection relation 5.9 () for y,0 = 0.14 and x,0 =0

Figure 6.32: Experimental values of membrane centreline deflection (+ + +),


load-deflection relation 5.9 () for y,0 = 0.28 and x,0 =0.

Chapter 6 Narrow flume experiments: static deflection, free bulge speed and
pressure amplification
123

Figure 6.33: Experimental values of membrane centreline deflection (+ + +),


load-deflection relation 5.9 () for y,0 = 0.14 and x,0 =0.14.

The influence of the initial longitudinal stress x,0 can be further studied with
the free bulge speed measurements. Figures 6.34, 6.35 and 6.36 show the average
measured free bulge speeds for tests 1, 2 and 3 respectively, for various internal
static pressure heads p0 . No major differences between the three figures are observed, except near p0 = 0.0m where the influence of the lateral initial strain y,0
is more pronounced. This can be explained by the fact that when p0 increases, the
stress due to the deflection y , such as the one in (5.5) takes over the initial stress
y,0 . This yields close values of free bulge speeds on each figures for the different
internal static pressure heads tested.
The free bulge speed is expected to be independent of the longitudinal stress x ,
as detailed in section 5.2. This is verified when comparing the measured free bulge
speeds from tests 1 and 2, that do not seem to be influenced by this longitudinal
stress.

Chapter 6 Narrow flume experiments: static deflection, free bulge speed and
124
pressure amplification

Figure 6.34: Averaged values of free bulge speed measurement, for y,0 = 0.14
and x,0 =0 for the 0.2mm membrane; () calculated values; (+ + +) averaged
values from pressure and laser measurements.

Figure 6.35: Averaged values of free bulge speed measurement, for y,0 = 0.28
and x,0 =0 for the 0.2mm membrane; () calculated values; (+ + +) averaged
values from pressure and laser measurements.

Chapter 6 Narrow flume experiments: static deflection, free bulge speed and
pressure amplification
125

Figure 6.36: Averaged values of free bulge speed measurement, for y,0 = 0.14
and x,0 =0.14 for the 0.2mm membrane; () calculated values; (+++) averaged
values from pressure and laser measurements.

6.5.4

Pressure variation at the stern

As before, pressure recordings at the stern made it possible to express the total
internal pressure amplitude at the stern as a linear function of the pressure due
to the undisturbed incident wave and results are shown in Figure 6.37 to Figure
6.42 for tests 1 to 6.
The measured pressures at the stern do not agree with the calculated ones. The
different peaks in the calculated values (right column) correspond, as before, to
the different harmonics of the membrane fn = nc/2L. Measured values are of
much smaller magnitude and an explanation of cause is given next.

6.5.5

Discussion

The main assumption made for the one dimensional model for bulge propagation
in the duct was that the external wave is undisturbed during its travel over the
membrane. This argument can be challenged with the data obtained from the
wave gauge located above the middle point of the duct.

Chapter 6 Narrow flume experiments: static deflection, free bulge speed and
126
pressure amplification

Figure 6.37: Magnitude of the pressure at the stern relative to the wave
pressure at the free surface elevation with conditions of test 1; left: measured
for the three runs (each symbols for one run); right: one dimensional theory
with c = 1.11m/s.

Figure 6.38: As for Figure 6.37, but with conditions of test 2; left: measured
for the three runs (each symbols for one run); right: one dimensional theory
with c = 0.88m/s.

Figure 6.39: As for Figure 6.37, but with conditions of test 3; left: measured
for the two runs (each symbols for one run); right: one dimensional theory with
c = 1.11m/s.

Chapter 6 Narrow flume experiments: static deflection, free bulge speed and
pressure amplification
127

Figure 6.40: As for Figure 6.37, but with conditions of test 4; left: measured
for the two runs (each symbols for one run); right: one dimensional theory with
c = 0.88m/s.

Figure 6.41: As for Figure 6.37, but with conditions of test 5; left: measured
for the two runs (each symbols for one run); right: one dimensional theory with
c = 1.11m/s.

Figure 6.42: As for Figure 6.37, but with conditions of test 6; left: measured;
right: one dimensional theory with c = 0.88m/s.

Chapter 6 Narrow flume experiments: static deflection, free bulge speed and
128
pressure amplification

Figure 6.43: Surface elevation at the middle point along the length of the
duct, relative to the incident wave amplitude.

Values of the water surface elevation (normalised by the incident wave amplitude)
are shown in Figure 6.43 for tests 1 and 3 and compared to those obtained with
the thick membrane and an internal static pressure head of 7cm. Values of surface
elevation for the thin membrane are much lower than that of the thick one. The
average ratio /A for the thick membrane is 0.84 whereas for the thin membrane,
it is 0.44 and 0.42 for tests 1 and 3, respectively. There is a significant difference
between the two membranes regarding the behaviour of the wave propagating
above them.

Figure 6.44: Ratio of the transmitted wave power, Pt , to the incident wave
power Pi for all the tests and measured with the first three harmonic components
of wave gauge signals.

Chapter 6 Narrow flume experiments: static deflection, free bulge speed and
pressure amplification
129
This can be further observed in Figure 6.44 where the transmitted power (normalised by the incident wave power) is plotted for all the tests carried out. Values
are found to be of a lower magnitude than that for the thick membrane (Figure
6.25), where in the latter case, Pt /Pi was less than 0.5 for the peaks of minimum
transmission.

Figure 6.45: Membrane displacement at thirty instants over one wave period
for configuration of test 1; wave frequency f =0.775Hz; (a) theory; (b) inferred
from laser sensors measurement of membrane displacement.

The membrane displacement measured from the laser sensors is compared in Figure
6.45 with that obtained from the one dimensional model, for the wave frequency
of f = 0.775Hz, for the configuration of test 1. For that one, the free bulge speed
is 1.11m/s and so f = 0.75Hz is the closest available frequency to the resonant
one, that is the 7th harmonic f = 7c/2L = 0.777Hz. It is worth noticing the
7

agreement in amplitude and space from the calculated and measured values of
displacement from x = 0m to the first node x = L/7 = 0.71m. At points further
along the membrane, the second node to node bump is visible, although its magnitude is already seen to be much lower than the predicted one. Further down the
membrane, the magnitude of the displacement vanishes completely and there is no
agreement between theory and measurement. Ripples propagating forth and back
on the membrane were observed after one or two wavelengths of the membrane
displacement. The energy carried in the bulge waves can be transferred to these
ripples in the same way to what happen on ocean surface waves for step gravity

Chapter 6 Narrow flume experiments: static deflection, free bulge speed and
130
pressure amplification
waves, and is thought here to be the cause of the mismatch between calculated
and measured values of membrane displacement.
With regards to the information provided by Figures 6.43, 6.44 and 6.45, an explanation is provided regarding the reasons why the one dimensional model cannot
be applied to the thin membrane.
For free bulge speeds as low as those propagating in the thin membrane, the
bulge wavelength is experimentally observed to be small, the membrane profile
exhibiting, for almost the lowest frequency tested, already six antinodes. In the
case of the thick membrane, where only the third and fourth harmonic (and the
corresponding membrane profile) were observed, it seems appropriate to consider
that the incoming waves propagate over the membrane without being significantly
perturbed, the bulge wavelength remaining larger than the water wavelength. This
way, the incoming waves sees the rubber membrane as a flat bed. Conversely,
the displacement profile exhibited by the thin membrane is made of many more
wavelengths, and this simulates a sort of sinusoidal seabed. Waves propagating
over such a bottom are more prone to lose their energy by various physical phenomena that are beyond the scope of the present work.
Finally, it can be argued that a thin membrane cannot handle great values of
stress. Observations included values of membrane deflection more than 4 times
the incoming wave amplitude, as can be seen in Figure 6.45 over the first metre
where the membrane is still subject to an almost undisturbed wave. Given the
submergence of the membrane for the tests conditions (d < 0.15m), the effect of
such displacements is more than likely to affect the incoming wave significantly.
The effect of increasing the initial stress in the membrane increases the free bulge
speed, so it goes back to the argument on the negative effects of having low free
bulge speeds in the one dimensional model.

Chapter 7
Extractable power
7.1

Introduction

This chapter details the tests carried on the version of the duct equipped with
the PTO system detailed in section 4.5 and placed in the wave basin presented
in section 4.3. For this test, the total length of the duct was reduced to 2m and
the membrane thickness was 1mm. Membrane deflection and pressure in the duct
were measured as in chapter 6. The extractable power was also measured using
(4.28). Results are compared with the predicted values from 5.43 and discussions
follow on the general efficiency of this system as a wave energy converter.

7.2
7.2.1

Experimental layout and test conditions


Description

The experimental sketch is given in Figure 7.1. The duct was placed on the bottom
of the wave basin (in the middle in the width direction) and kept in position using
heavy masses. The same hydraulic pump system used previously maintened the
desired internal static pressure head. Laser sensors (not drawn) were mounted
as in Figure 4.15 to measure the membrane displacement. Resistive wave gauges
(two per position in the width direction) were used to measure incident, reflected
and transmitted wave power. Two air pressure transducers and three wave gauges
were mounted in the bent tube.
131

132

2.0m
0.580m

Wg3

0.5m 0.44m

High
density
foam

Wave paddle

2.7m

Wg1 Wg2

Pt1

2.0m

Air pressure
transducers
Wave
gauges

Copper pipes

Absorbing
beach

Chapter 7 Extractable power

Figure 7.1: Experimental layout for the rubber duct equipped with the PTO.
Dashed lines correspond to different internal static pressure heads.

133

Chapter 7 Extractable power

The test conditions are summarised in Table 7.1 where the range of impedances,
wave frequencies and number of waves amplitudes used are specified. Waves were
generated in the range of [0.45Hz-1.2Hz] with increments of 0.025Hz. For this
range (and this water depth), it was not necessary to change the distance between
Wg1 and Wg2. Pictures of this set-up are provided in Appendix C.

Test no

Internal static
pressure head

Z = ZPTO /Zd

(m)

Wave
frequencies

Number of
wave

Number
of

(Hz)

amplitudes

runs

0.03m

1.03

0.45-1.2

0.03m

1.10

0.5-0.875

0.03m

1.93

0.45-1.075

0.02m

1.73

0.45-1.2

0.02m

3.47

0.45-1.2

0.02m

1.08

0.45-1.2

0.02m

1.02

0.45-1.2

0.02m

0.5-1.1

Table 7.1: PTO experiment: test conditions

As in the experiment carried in chapter 6, the signals obtained from the wave
gauges (in the flume and in the PTO), the pressure transducers (water and air) and
the laser sensors were found to be predominantly monochromatic. Hence the data
analysis carried out hereinafter neglects harmonics other than the fundamental
wave frequency.

7.2.2

Power take-off in experimental conditions

The PTO used was aimed to be viscous, meaning that pressure and flow rate are
in phase, hence the water-air interface displacement and the pressure lagged by
90 . A way to cross check this assumption is to plot one signal against the other.
Such plots are shown in Figure 7.2 for the four tests with Z finite (Table 7.1) and
the duct set with an internal static pressure head of 0.02m.

134

Chapter 7 Extractable power

Figure 7.2: PTO air pressure versus PTO water level for the rubber duct with
an internal static pressure head of 0.02m; f =0.625Hz; (a) test 7; (b) test 6; (c)
test 4; (d) test 5.

PTO water level and air pressure are seen to be in quadrature for tests 6 and 7
when Z=1.02 and Z=1.08 respectively. As the impedance increases, the phase lag
between the water level and the air pressure decreases, meaning that the phase
lag between flow rate and pressure, on the other hand, increases.
The measured phase lag is compared in Figure 7.3 with the theoretically phase lag
calculated using (4.25). The values seem to be in fair agreement and the simplified
model from section 4.5.3 is found to give an accurate explanation of the phase lag
happening in the chamber.
Equation (4.26) can be rewritten as
q0 = (St 0 ),

(7.1)

where the term St 0 represents the flow rate of the air due to the water-air interface displacement. If the air was not compressed, this flow rate would be equal
to q0 . The ratio is plotted in Figure 7.4 for the four previous impedances in the
same frequency range as before. It is seen to be close to 1 for impedances Z =1.02

Chapter 7 Extractable power

Figure 7.3: Phase lag between PTO water level and PTO air pressure as a
function of wave frequency for the duct with an internal static pressure head of
0.02m ; measured, from (4.25); (a) test 7; (b) test 6; (c) test 4; (d)
test 5.

Figure 7.4: Ratio of the flow rate in the copper pipes to that of the air in the
chamber, for the duct with an internal static head of 0.02m.

135

136

Chapter 7 Extractable power

and Z =1.08, while for the two other impedances Z =1.73 and Z =3.47, it is seen
to decrease by up to 20%. For those higher impedances, the flow rate in the pipes
is then slightly different from that due to the water displacement in the chamber.
One of the main explanations for that to occur is that air is being compressed at
those impedances. Power required for this compression is stored in the air (as in the
case of pumping air in a tire) and mostly restored when air expands, the remaining
power being used into heating up the air during compression. This latter power
is assumed to be negligible hereinafter, and even though this power is present at
higher impedances, it is not seen as useful power potentially absorbed by the PTO.
Data from test 8, where the sign stands for the fact that all pipes were closed,

allows measurement of the bulk modulus of the air enclosed in the chamber, that
is
dP
.
(7.2)
KT = V
dV
KT characterises the resistance of the air to a uniform compression. This bulk
modulus is shown in Figure 7.5 and yields an average value of KT = 0.53 105 Pa,
while the bulk modulus of air in the literature is found to be KT = 1 105 Pa.
The main explanation for this difference is the existence of air leaks in the system.
This differential is considered to be of minor influence on the PTO impedance
calculation and therefore will be neglected.
This explanation has been carried out with data from the situation where the
internal static pressure head was 0.02m. The same demonstration can be undertaken with data from the situation where the internal static pressure head is
0.03m. Moreover, a wider range of impedances has been studied in a previous
experiment and the corresponding data are presented in Appendix A.

137

Chapter 7 Extractable power

Figure 7.5: Measured values of the bulk modulus of air, with configuration as
in test 8.

7.3
7.3.1

Matching impedances configuration


Introduction

Tests were carried for the two internal static pressure heads provided in Table 7.1,
with two sets of PTO configurations. In the first case, the impedance of the PTO
was as close as possible to that of the duct and in the second case, a deliberate
mismatch between the PTO and duct impedances was set up to study the influence on the power absorbed by the PTO.
The first configuration (matching) is described in this section. Capture widths
are presented, and in a similar way to what has been done in sections 6.4.3 and
6.4.4, characterisation of the inner flow is made is term of bulge pressure components.

138

Chapter 7 Extractable power

7.3.2

Power capture

Wave gauges recorded the water level e0 in the rigid tube so that the associated
flow rate in the copper pipes could be evaluated, leading to the measured power
absorbed by the PTO following (4.28). Resistive gauges in the basin, ahead of the
duct, provided the measurements of the incident wave power from which capture
widths were calculated.
Capture widths are shown in Figures 7.6 and 7.7 for tests 1-2 and 6-7, respectively. They are compared with the calculated values of capture widths obtained
from (5.43) for those corresponding impedances.
For both internal static pressure heads, the difference in impedances between the
two corresponding tests does not seem to be of a strong influence and can be
thought to represent the case Z = 1. Calculated and measured values seem to
be in fair agreement. Discrepancies are noticeable in term of absolute values of
capture width, but the general trend is similar.

Figure 7.6: Capture width (in duct widths) for a static pressure head of
3cm; lines: computed from theory; symbols: measured; colour legend: test 2
(Z = 1.10) in grey and test 1 (Z = 1.03) in black.

Chapter 7 Extractable power

139

Figure 7.7: Capture width (in duct widths) for a static pressure head of
2cm; lines: computed from theory; symbols: measured; colour legend: test 6
(Z = 1.08) in grey and test 7 (Z = 1.02) in black.

Mathematically the peaks of maximum power capture correspond to frequencies


minimising the denominator of |q|2 where q is obtained with (5.42). A physical
explanation can be given to them, as they correspond to one of several interesting phenomena happening in the duct or the chamber. The theoretical peak of
maximum capture widths in Figure 7.6 was for f =0.725Hz. For this configuration
the length of the slug of water was l =0.683m, yielding a resonant frequency of
p
fslug = (1/2) g/l =0.6Hz.
Two peaks are visible in Figure 7.7. The first theoretical peak was for f =0.68Hz.
The length of the slug of water was, for this configuration, l =0.672m, yielding a
resonant frequency of fslug =0.61Hz. Moreover, at f =0.585Hz, k1 = k2 meaning
that the free bulge speed in the duct and the water wave speed match. This corresponds to a condition of maximum internal velocity at the stern, and hence flow

rate in the pipes. Finally, at f =0.625Hz, k1 = k2 1 + which is a solution of


<(K 2 k12 )2 = 0, maximising |q|2.
The second visible peak in Figure 7.7 occurred for f =1.05Hz. This corresponds
to the second harmonic of the duct as studied in section 6.4.2. The free bulge
speed was c =2.1m/s so f2 =2.1/2=1.05Hz. This is again a condition of maximum

140

Chapter 7 Extractable power

internal velocity at the stern.


Experimental capture widths of up to 2.2 times (tests 7-8) and 2 times (tests
1-2) the ducts width are obtained with this experimental configuration. Due to
the simultaneous occurrence of physical phenomenon maximising the flow rate in
the pipes around one frequency, maximum power capture is not concentrated at
this frequency. The bandwidth of these non-dimensional capture widths for which
it is greater than one are found to be [0.575Hz;0.75Hz] and [0.65Hz;0.8Hz] for
tests 7-8 and tests 1-2, respectively. Attention on how these resonant peaks are
characterised in term of pressure variation in the duct is given in the next section.

7.3.3

Pressure variation in the duct

The internal pressure was mapped in the manner described in section 6.4.3, using
this time the adjusted pressure from the air pressure transducers for the pressure
at the stern. This is shown in Figures 7.8 and 7.9 for tests 7-6 respectively and
the frequency of maximum power output f =0.625Hz.

Figure 7.8: Pressure in the duct at thirty instants over one wave period with
an internal static pressure head of 2cm; wave frequency f =0.625Hz; (a) theory for Z = 1.02; (b) inferred from laser sensors measurements of membrane
displacement.

141

Chapter 7 Extractable power

Figure 7.9: Pressure in the duct at thirty instants over one wave period with
an internal static pressure head of 2cm; wave frequency f =0.625Hz; (a) theory for Z = 1.08; (b) inferred from laser sensors measurements of membrane
displacement.

Calculated and measured values of this internal pressure follow the same trend,
although the measured values seem slightly smaller in amplitude than the theoretical values. This can be improved in the model by increasing the loss angle
by a couple of degrees. Doing so is justified by saying that a higher hysteresis
coefficient can represent the energy lost in the system by wave radiation from the
membrane, a phenomenon not otherwise taken into account. Unlike in the closed
ends configuration, the pressure is not reflected at the stern and its amplitude is
seen to be increasing along the length of the rubber duct. Figure 7.10 shows a
similar plot for the duct with a internal static pressure head of 3cm and, again,
for the first frequency of maximum power output.
Moreover, Figure 7.11 shows the total internal pressure for the second peak observed in Figure 7.7 (for test 6). Nodes and anti-nodes as observed, in a similar
way to the closed ends configuration. This peak occurred for the second harmonic
of the membrane, as stated earlier.

142

Chapter 7 Extractable power

Figure 7.10: Pressure in the duct at thirty instants over one wave period
with an internal static pressure head of 3cm; wave frequency f =0.75Hz; (a)
theory for Z = 1.03; (b) inferred from laser sensors measurements of membrane
displacement.

Figure 7.11: Pressure in the duct at thirty instants over one wave period
with an internal static pressure head of 2cm; wave frequency f =1.05Hz; (a)
theory for Z = 1.08; (b) inferred from laser sensors measurements of membrane
displacement.

Chapter 7 Extractable power

143

The total internal pressure is seen to be well predicted by the model from section
5.5. As in the case of closed ends, this total pressure was decomposed by a linear
least squares method to verify whether the bulge pressure could be written as the
sum of the three components (5.33)-(5.35).

7.3.4

Bulge components

A least-squares method was used to express the total pressure in the duct as a
sum of three distinct components, as for the case of the duct with closed ends.
An example of values obtained for the amplitude and phase of this inner pressure wave is given in Figures 7.12 and 7.13 for the two peaks of maximum power
capture mentioned above and for the experimental conditions of test 6.
As before, points represent the measured values of amplitude and phase of the
internal total pressure at 10 points along the membrane (and the pressure at the
stern and bow), while the line is the sum of the components obtained with the
least-squares method. In Figure 7.12, points follow the reconstructed internal pressure amplitude and phase and the amplitude is seen to increase along the length
of the membrane as previously noticed in Figures 7.8 and 7.9. The sum of these
three components (plus the wave pressure) were used to map the internal pressure
in a similar way as before, and this is shown in Figure 7.14.
In Figure 7.13, the general shape of the amplitude profile along the length of
the membrane follows that of a natural mode. However, the node and antinodes
of the amplitude obtained from the least-squares seem to be slightly shifted towards points further along the membrane. This can be well appreciated in Figure
7.15 where the location of nodes and antinodes are for slightly different positions.
Also the second minimum is not in fact a node as the phase tends to around
x =1.6m, but does not completely cross the line = .

144

Chapter 7 Extractable power

Figure 7.12: Measured total internal pressure, normalised by the pressure acting on the membrane (a) and phase (b) for the duct with an internal static pressure head of 2cm, plotted as points. Lines represent amplitudes and phases computed as the sum of the components obtained from the linear squares method.
Wave frequency f =0.625Hz, with conditions as in test 6

Figure 7.13: Measured total internal pressure, normalised by the pressure acting on the membrane (a) and phase (b) for the duct with an internal static pressure head of 2cm, plotted as points. Lines represent amplitudes and phases computed as the sum of the components obtained from the linear squares method.
Wave frequency f =1.05Hz, with conditions as in test 6

The magnitudes of the three bulge components obtained at each experimental


frequency were compared with the predicted ones from (5.33)-(5.35) and results are
shown in Figures 7.16-7.18. The choice of the logarithmic axes is to distinguish the
peak occurring in the region f =1.05Hz for Pb+ and Pb , from that at f =0.625Hz.

145

Chapter 7 Extractable power

Figure 7.14: Pressure in the duct at thirty instants over one wave period with
an internal static pressure head of 2cm; wave frequency f =0.625Hz; (a) inferred
from laser sensor measurement of membrane displacement amplitude; (b) sum
of the three bulge wave components (and the wave pressure) obtained for a least
squares method.

Figure 7.15: Pressure in the duct at thirty instants over one wave period with
an internal static pressure head of 2cm; wave frequency f =1.05Hz; (a) inferred
from laser sensor measurement of membrane displacement amplitude; (b) sum
of the three bulge wave components (and the wave pressure) obtained for a least
squares method.

146

Chapter 7 Extractable power

Figure 7.16: Magnitude of Pb+ relative to the wave pressure acting on the
membrane with an internal static pressure head of 2cm and conditions as in
test 6; symbols: experimental values inferred from a least squares method; :
calculated from (5.34).

Figure 7.17: Magnitude of Pb relative to the wave pressure acting on the


membrane with an internal static pressure head of 2cm and conditions as in
test 6; symbols: experimental values inferred from a least squares method; :
calculated from (5.35).

147

Chapter 7 Extractable power

Figure 7.18: Magnitude of PbA relative to the wave pressure acting on the
membrane with an internal static pressure head of 2cm and conditions as in
test 6; symbols: experimental values inferred from a least squares method; :
calculated from (5.33).

The agreement between the three amplitudes obtained from the least-squares
method and the predicted ones can be judged as satisfactory. Unlike the closed
ends configuration from chapter 6, the amplitude of Pb seems much lower than
the amplitudes of Pb+ and PbA , especially around f =0.625Hz. These two last components being those travelling from bow to stern; this is consistent with Figure
7.11 where the pressure in the duct is seen to increase along its length. On the
contrary, for f =1.05Hz, the amplitudes of Pb+ and Pb are the same. Moreover,
those two components are travelling with the same speed, in opposite direction,
giving another justification of the partial standing wave pattern in the duct for
that frequency.

7.4
7.4.1

Mismatch impedances configuration


Introduction

The analysis carried out above stands for the configuration with the impedance
of the PTO as close as possible to that of the duct. More tests were carried with

148

Chapter 7 Extractable power

a mismatch in impedances in order to investigate the applicability of the model


detailed in section 5.5, and a similar analysis as in the previous section is carried
out here.

7.4.2

Power capture

Values of capture widths are shown in Figures 7.19-7.21 for tests 3, 4 and 5, respectively. Measured and calculated power capture widths seem to agree quite well and
any discrepancies in frequencies of maximum power capture can be attributed to
the experimental difficulty of maintaining the internal static pressure head (hence
the free bulge speed) constant through the whole experimental run. Mismatches
between the impedance of the PTO and that of duct do not seem to be of strong
influence since for both internal static pressure heads, the general trend of the
capture width is quite similar to that of the case of matching impedance for the
same configuration.

Figure 7.19: Capture width (in duct widths) for a static pressure head of 3cm;
lines: computed from theory; symbols: measured for the three runs of test 3.

Chapter 7 Extractable power

Figure 7.20: Capture width (in duct widths) for a static pressure head of 2cm;
lines: computed from theory; symbols: measured for the run of test 4.

Figure 7.21: Capture width (in duct widths) for a static pressure head of 2cm;
lines: computed from theory; symbols: measured for the run of test 5.

149

150

Chapter 7 Extractable power

7.4.3

Pressure variation in the duct and bulge components

The total internal pressure can be inferred as before from the displacement of the
membrane. This is shown for a wave frequency f =0.625Hz in Figure 7.22 with
the experimental conditions of test 3. As for the matching case configuration,
the pressure is seen to increase along the length of the second half of the duct.
General trends are similar between calculated and measured values, and the lower
amplitude in the measure internal pressure can be attributed to the influence of
the radiated waves from the membrane displacement affecting the incoming wave
pressure, the latter being assumed not to be disturbed in the theoretical model
(chapter 5).

Figure 7.22: Pressure in the duct at thirty instants over one wave period
with an internal static pressure head of 3cm; wave frequency f =0.625Hz; (a)
theory for Z = 1.93; (b) inferred from laser sensors measurements of membrane
displacement.

As before, a linear squares method was used to separate the three bulge components and a similar mapping of the internal pressure shown in Figure 7.22 was
performed with the values obtained. This is shown in Figure 7.23 for the amplitude
and phase of the internal pressure and in Figure 7.24 for the reconstruction.

151

Chapter 7 Extractable power

Figure 7.23: Measured total internal pressure, normalised by the pressure acting on the membrane (a) and phase (b) for the duct with an internal static pressure head of 3cm, plotted as points. Lines represent amplitudes and phases computed as the sum of the components obtained from the linear squares method.
Wave frequency f =0.625Hz, with conditions as in test 3.

Figure 7.24: Pressure in the duct at thirty instants over one wave period with
an internal static pressure head of 3cm; wave frequency f =0.625Hz; (a) inferred
from laser sensor measurement of membrane displacement amplitude; (b) sum
of the three bulge wave components (and the wave pressure) obtained for a least
squares method.

152

7.5

Chapter 7 Extractable power

Water-air interface displacement and pressure in the PTO

Values of measured normalised pressure in the PTO chamber and of the vertical
displacement of the slug of water are seen in Figure 7.25 and 7.26 for internal static
pressure heads of 3cm and 2cm, respectively and for all the tests carried. They
are compared with those obtained from the one dimensional theory of section 5.5.2.
Measured and calculated values seems to agree, although the measured values
for test 2 are of smaller magnitudes than the calculated ones. Both trends (pressure and amplitude) are similar to those of the capture width plots for the corresponding configuration, in the sense that they have their maximum values in the
frequency range where the power absorbed by the PTO is maximum.
The pressure in the chamber is seen to be increasing as the impedances increases,
while the opposite is true for the amplitude of the water-air interface motion.
Interestingly, there seems to be a slight downwards shift of the theoretical peak
frequency of the maximum pressure as the impedance increases, in the same way
as for the capture width in Figure 7.21, for instance. On the other hand, the
experimental peak of pressure in the chamber (around f = 0.625Hz) seems to be
affected in amplitude, but not in frequency, by the change of impedance.
Measured (and calculated) values of the motion amplitude inside the rigid bent
tube indicates amplitudes of motions up to 3 times (Figure 7.26) and 2.5 times
(Figure 7.25) the incoming wave amplitude. The bandwidth of the frequency range
for which the normalised amplitude of motion is greater than one is seen to become
narrower as the impedance increases. For the duct with an internal static pressure
head of 2cm, this (experimental) range is 0.45Hz < f < 0.8Hz for Z=1.02 and
Z =1.08, 0.45Hz < f < 0.75Hz for Z=1.73 and 0.45Hz < f < 0.65Hz for Z=3.47.
For the duct with an internal pressure of 3cm, this range is 0.45Hz < f < 0.8Hz
for Z=1.03, 0.5Hz < f < 0.8Hz for Z=1.10 and 0.45Hz < f < 0.7Hz for Z=1.93.
On the other hand, as the pressure in the chamber increases for increasing impedances,
the frequency bandwidth for which the normalised pressure is greater than one gets
larger. For the duct with an internal static pressure head of 2cm, this (experimental) range is 0.45Hz < f < 0.775Hz for Z=1.02, 0.45Hz < f < 0.8Hz for Z =1.08,
0.45Hz < f < 0.825Hz for Z=1.73 and 0.45Hz < f < 1.075Hz for Z=3.47. For

153

Chapter 7 Extractable power

the duct with an internal pressure of 3cm, this range is 0.45Hz < f < 0.8Hz for
Z=1.03, 0.5Hz < f < 0.8Hz for Z=1.10 (except the points for f = 0.525Hz and
f = 0.55Hz) and 0.45Hz < f < 0.85Hz for Z=1.93. However, as for the amplitude of motion, the change of bandwidth is significant only for high values of
dimensionless impedance and in this case, for Z =3.47.

Figure 7.25: Pressure (left column) in the PTO and amplitude (right column)
of the water-air interface displacement. Lines are the calculated values from the
theory in section 5.5.2 and points represent measured values for (a) test 1; (b)
test 2; (c) test 3.

154

Chapter 7 Extractable power

Figure 7.26: Pressure (left column) in the PTO and amplitude (right column)
of the water-air interface displacement. Lines are the calculated values from the
theory in section 5.5.2 and points represent measured values for (a) test 7; (b)
test 6; (c) test 4 and (d) test 5.

Chapter 7 Extractable power

7.6

155

Wave analysis

The two pairs of wave gauges placed in front of the duct made it possible to
measure the incident and reflected wave powers Pi and Pr , respectively. The two
pairs of gauges behind the duct enabled the measurement of the transmitted wave
power Pt . Unlike the experiment for the closed ends, the energy conservation
includes a term related to the power captured Pf by the PTO, yielding
Pt + Pr + Pf + Ploss = Pi

(7.3)

Figures 7.27 and 7.28 show the measured values of the transmitted, reflected and
captured power, normalised by the incident wave power, for all the tests taking
only the first harmonic of the wave gauge signals, or the first three harmonics.
The first comment to be made is that the higher wave harmonics do not seem
to be carrying much power, as graphs from left and right column are similar for
both figures. This supports the present approach of neglecting higher harmonics
in the data analysis.
The sum of the normalised transmitted, reflected and absorbed power is seen
to be greater than one (by up to 25%) for both set of duct pressurisation and for
the range of wave frequency where the power output is maximum. For the same
range of wave frequencies, values of normalised transmitted power are seen to be
also close to unity even though neither the normalised reflected nor the capture
widths are zero. This is the result of the poor efficiency of the absorbing beach
for that range of frequency, usually f 0.85Hz as seen in Figure 4.6, where the
reflection coefficient from the foam is seen to be up to 20%. Taking this error into
account, it can be deducted that the magnitude of the power lost (Ploss = 1 )

in the system is small.

156

Chapter 7 Extractable power

Figure 7.27: Wave analysis for the duct in with an internal static pressure
head of 3cm and configurations as in (a) test 1; (b) test 2; (c) test 3 and (d)
test 5. Left column: power carried by the first harmonic; right column: power
carried in the first three harmonics of the signals from the wave gauges.

Chapter 7 Extractable power

Figure 7.28: Wave analysis for the duct in with an internal static pressure
head of 2cm and configurations as in (a) test 7; (b) test 6; (c) test 4 and (d)
test 5. Left column: power carried by the first harmonic; right column: power
carried in the first three harmonics of the signals from the wave gauges.

157

158

7.7

Chapter 7 Extractable power

Conclusions

The information presented in this chapter relates to the the experiment carried
on the duct equipped with a power take-off system. This PTO was investigated
and its characteristics were checked for the experimental conditions tested. It was
observed that increasing the impedance of this PTO results in a phase lag between
pressure in the chamber and the flow rate due to the water-air interface motion.
This yields a difference between this flow rate and the one in the copper pipes that
is of interest for measuring the power absorbed by the PTO. This phase difference
was explained with fair agreement with the simple mechanical model described in
section 4.5.3.
Power data were presented for various configurations, in terms of internal static
pressure heads and PTO impedances. Capture widths up to 2.2 times the width
of the duct were obtained for impedances of the PTO close to that of the duct.
Maximum power output occurred for any physical phenomenon maximising the
velocity at the stern (and hence the flow rate) such as k1 = k2 or k2 = nc/2L. For
the first case, the pressure was seen to increase along the length of the duct, while
in the second case corresponded as in chapter 6 to a natural mode of the membrane. Again, this pressure mapping was enabled by the use of the laser sensors
to measure membrane displacement beneath waves.
The bulge pressure was characterised as the sum of three components using a least
squares method as introduced in chapter 6. For the maximum power absorbed by
the PTO corresponding to the case k1 = k2 , the two components travelling in
the same direction as the wave were of higher amplitudes than the reflected wave
amplitude. However, similar results to that of the closed ends duct were obtained
for the frequency at which k2 = c/L, namely a partial standing wave in the duct
for that frequency.
The amplitudes of these components, together with the capture widths, pressure
in the duct, and amplitude of motion of the slug of water were compared with
the calculated corresponding values from the one dimensional theory such as in
section 5.5. The agreement was satisfactory and discrepancies could be attributed
to the incoming wave being affected by the radiated waves from the membrane, a
phenomenon otherwise not included in the model.

Chapter 7 Extractable power

159

In comparison to the case of the raised duct in the narrow flume, the power lost
in the system was checked to be small. It is more likely that in the latter case,
the power lost in the system was due to the experimental set-up, e.g. the gaps
and sharp corners of the artificial seabed. The pressure inside the duct was of the
same order of magnitude as with closed ends, and the length of the duct (hence the
rubber sheet) was 60% less, making as before the losses in the rubber negligible.

Chapter 8
Two dimensional analytical model
of a duct covered with a tensioned
membrane beneath waves
8.1

Introduction

The results presented for the duct with closed ends and with the two types of
membrane emphasised the limits of the one dimensional theory from chapter 5.
The main hypothesis that waves were undisturbed during their travel over the
membrane was experimentally shown to be invalid for the thin membrane. The
purpose of this chapter is to present a two dimensional analytical model developed
for the interactions of water waves with a submerged rectangular duct, enclosing
water beneath its top rubber face, and placed beneath waves propagation. Results
are compared with those of the experiments in the wide flume.
The idea is to expand the work initiated by Cho and Kim (1998) on a stretched horizontal submerged membrane, mentioned in section 3.3.2. The diffracted and radiated potentials in each region of the fluid are expressed using the same eigenfunction expansion method and the series coefficients obtained with the solid boundary
conditions on the duct wall or by matching the different potentials at the interface.
The membrane equation is solved analytically to obtain the modal amplitude of
membrane displacement.

161

Chapter 8 Two dimensional analytical model of a duct covered with a tensioned


162
membrane beneath waves
Before any comparison with experimental values is made, the model is checked
with previous studies of wave interactions by rectangular submerged obstacles, as
this corresponds to the limiting case of infinite tension in the membrane. The main
problem encountered with the development of such a model for the configuration
of the duct with closed ends is the conservation of the volume of water in it. Three
methods are presented and discussed for that purpose.
This chapter is intended to present the model and the code developed while the
different equations and other mathematics can be found in Appendix B.

8.2

Problem formulation for the closed ends configuration

The interactions of a submerged flexible membrane with monochromatic incident


waves are considered. A definition sketch is given in Figure 8.1. The coordinate
system is (x,y) with y pointing upwards and y = 0 being at the mean free surface
and x = 0 the middle point of the membrane. The length of the membrane is
L = 2s and it is fixed at both ends, x = s. The study is carried out in two
dimensions. The submergence depth of the membrane is d and the total water
depth is h. Tension is applied uniformly in the membrane in the x-direction.

y
O
x
(1)

(2)

d
h

(3)
s
Figure 8.1: Definition sketch of a closed duct with elastic membrane on its
top.

Chapter 8 Two dimensional analytical model of a duct covered with a tensioned


membrane beneath waves
163
The geometry studied by Cho and Kim (1998) does not enclose any fluid, meaning
that there is no wall between regions (1) and (3). In this study, the domain was
closed on this boundary x = s for h y d, leading to the shape of a
submerged step with an elastic top. The aim is to use the method developed by
Cho and Kim (1998) but for this new geometry.
The domain of study is separated into three regions as detailed in Figure 8.1.
Wave amplitude and membrane motions are assumed to be small in order to use
linear theory. The problem can be decomposed into a diffraction problem (in this
case region (3) does not exist and the problem is that of diffraction by a rectangular submerged obstacle) and a radiation problem. The total velocity potential
at point (x,y) can then be written as
(x, y) = D (x, y) + R (x, y),

(8.1)

with
D (x, y) = i (x, y) + S (x, y),

(8.2)

where D , R , i and S represents the diffraction, radiation, incident and scattering potentials respectively.
The incident wave used can be expressed as
i (x, y) =

igA cosh k1 (y + h) ik1 x


e
,

cosh k1 h

(8.3)

where g is the gravitational acceleration, A the wave amplitude (taken as unity)


and i the imaginary unit. Assuming harmonic motion of frequency , the velocity
potential (x, y, t) can be expressed as
(x, y, t) = <((x, y)eit),
where is one of the potentials mentioned above.

(8.4)

Chapter 8 Two dimensional analytical model of a duct covered with a tensioned


164
membrane beneath waves

8.2.1

Diffraction problem

8.2.1.1

Definitions of the potentials

For the diffraction problem, the region (3) does not exist and the duct can be seen
as a concrete block. The boundary conditions for the diffraction velocity potential
are the Laplace equation, the combined linear free surface boundary condition,
the bottom boundary condition of zero velocity on y = d and y = h.
2 D 2 D
+
x2
y 2
D 2
D
y
g
D
y
D
x
D
ik1 D )
lim (
x x
D
y

= 0,

(8.5)

= 0

on y = 0,

(8.6)

= 0

on y = h, x s,

(8.7)

= 0

on x = s, h y d,

(8.8)

= 0,
= 0

(8.9)
on y = d, s x s.

(8.10)

Given the geometry of the problem, the potentials in each region could be expressed
using symmetric functions. However, symmetric potentials alone can not represent
progressive waves, hence the necessity to also include asymmetric potentials. The
diffraction potential is then split into symmetric and asymmetric parts as
D (x, y) = SD (x, y) + A
D (x, y).

(8.11)

The symmetric diffraction potentials are given by Cho and Kim (1998) for their
geometry (region (3) not closed) as
S(1)

S(2)
D

X
ig 1 10 x
{ e
f10 (y) +
aSn e1n (x+s) f1n (y)},
2
n=0

ig X S
b cosh(2n x)f2n (y),
=
n=0 n

(8.12)
(8.13)

with the eigenfunctions f1n and f2n detailed in Appendix B together with the 1n
S(1)
series. The first part of D represents half of the incident wave (the other hlaf
being carried by the assymetric potential for the same region) and the second part,
evanescent waves that vanish at x = . The symmetric potential in region (2)

Chapter 8 Two dimensional analytical model of a duct covered with a tensioned


membrane beneath waves
165
is expressed so that is satisfy the boudary condition at y = d.
S(1)

S(1)

Matching D and D on x = s over [0,d] yields expressions for the bSm


as functions of the aSm . Matching the velocities at the same boundaries and setting zero velocity on [d,-h] leads to a system of equations from which the aSm can
be obtained, hence the bSm . Those governing equations and their solutions are set
out in Appendix B.
Using the same procedure for the asymmetric potentials
A(1)
D

A(2)

X
ig 1 10 x
1n (x+s)
f10 (y) +
aA
f1n (y)},
= { e
ne
2
n=0

ig X A
b sinh(2n x)f2n (y),
n=0 n

(8.14)
(8.15)

the total scattered potentials in each regions can be obtained.

8.2.2

Radiation problem

8.2.2.1

Definition of the potentials

The radiation problem is approached using a series expansion for the membrane
displacement in the form of its natural modes:
(x) =

j fj (x),

(8.16)

j=1

where j is the unknown complex modal amplitude for the j th mode and fj (x) is
the associated modal function given by

S x
2j 1

fjS (x) = cos j , Sj =


j = 1, 2, 3...
s
2
fj (x) =

A x

f A (x) = sin j , A = j j = 1, 2, 3...,


j
j
s

(8.17)

where S stands for symmetric mode and A for asymmetric mode. The symmetry is
about x = 0 as can be seen in Figure 8.1. The radiation potential j,R is governed
by the same boundaries conditions as the diffracted potential and by the dynamic

Chapter 8 Two dimensional analytical model of a duct covered with a tensioned


166
membrane beneath waves
condition
j,R
= ifj (x)
y

on y = d,

(8.18)

stating that the membrane modal normal velocity is the same as the fluid velocity around it. As for the diffraction problem, lR is split into symmetric and
asymmetric parts as
j,R (x, y) = Sj,R (x, y) + A
j,R (x, y),

(8.19)

and the radiation problem is treated using separately these two potentials as in
the diffraction case.
The symmetric modal radiation potentials are expressed as
S(1)

j,R

S(2)
j,R

S(3)
j,R

ig X S 1n (x+s)
{
a e
f1n (y)},
n=0 jn

ig X S
i S(2)
= {
bjn cosh(2n x)f2n (y) + j,R (x, y)},
n=0
g

ig X S
i S(3)
= {
cjn cosh(3n x)f3n (y) + j,R (x, y)},
n=0
g

(8.20)
(8.21)
(8.22)

with the j,R for each region is as defined in Appendix B and used to satisfy the
dynamic condition (8.18).
In a similar way as for the scattered potentials (matching of velocities and potential at the x = s interface) the aSjm and bSjm series can be obtained. The cSlm
series are obtained from the zero velocity condition
S(3)

j,R
(s, y) = 0
x

on x = s, h y d.

(8.23)

In a similar way with the asymmetric modal radiated potentials


A(1)
j,R

A(2)
j,R

A(3)

j,R

ig X A 1n (x+s)
a e
f1n (y)},
= {
n=0 jn

ig X A
i A(2)
= {
bjn sinh(2n x)f2n (y) + j,R (x, y)},
n=0
g

i A(3)
ig X A
cjn sinh(3n x)f3n (y) + j,R (x, y)},
{
n=0
g

(8.24)
(8.25)
(8.26)

Chapter 8 Two dimensional analytical model of a duct covered with a tensioned


membrane beneath waves
167
A
A
the aA
jm and bjm series can be obtained. The clm series are obtained from the zero

velocity condition as in the symmetric case.

8.2.2.2

Membrane deformation

The equation for the elastic membrane, in absence of physical damping can be
written following Cho and Kim (1998)
Tx

d2
(x) + m 2 (x) = i[(3) (x, d) (2) (x, d)],
2
dx

(8.27)

where Tx is the membrane tension, the fluid density and m the mass of the
membrane per unit length.
The kinematic condition (8.27) can be written in a discrete form as detailed in
Appendix B as

[KijS 2 (MijS + aSij ) ibSij ] jS = FiS


{z
}
|
j=1

i = 1, 2, 3....

(8.28)

ij

Truncating (8.28) to the desired numbers of modes M, the symmetric modal amplitudes jS can be obtained. The same type of system is used for the asymmetric
modes.
The membrane motion can be expressed as
(x) =

(jS fjS (x) + jA fjA (x)),

(8.29)

j=1

and the potential in each region (k) using


(k)

(x, y) =

A(k)
D

S(k)
D

M
X
j=1

S(k)

A(k)

(jS j,R (x) + jA j,R (x))

: k = 1, 2, 3. (8.30)

Chapter 8 Two dimensional analytical model of a duct covered with a tensioned


168
membrane beneath waves

8.2.3

Reflection and transmission coefficients

Combining the symmetric and asymmetric parts, the total potential in region (1)
can be expressed as:
N
ig 10 x
ig X A
(an + aSn )e1n (x+s) f1n (y)
f10 (y)
(x, y) = e

|
{z
}
n=0
|
{z
}
incident wave
scattered wave
M N
ig X X S S
1n (x+s)
(j ajn + jA aA
f1n (y),

jn )e
j=1 n=0
|
{z
}
radiated wave
(1)

(8.31)

where the sum over the eigenvalues has been truncated to N. The reflection
coefficient is therefore




Kr = radiated+scattered
incident

as x

(8.32)

where is the surface elevation of the wave referred to.


Using the dynamic free-surface condition and taking x = 2s,
i
i (x, 0)
g
= e210 s .

incident =

(8.33)

Noticing that for n 1, the 1n are real, the terms for n 1 in the scattered and
radiated waves vanish for x , yielding
radiated+scattered =

((aS0

aA
0)

L
X

(jS aSj0 + jA aA
j0 ))

(8.34)

j=1

and
Kr =

|[(aS0

aA
0)

M
X
j=1

210 s
(jS aSj0 + jA aA
|.
j0 )]e

(8.35)

The transmission coefficient is given by Cho and Kim (1998) as


Kt =

|[(aS0

aA
0)

M
X
210 s
(jS aSj0 jA aA
|.
+
j0 )]e
j=1

(8.36)

Chapter 8 Two dimensional analytical model of a duct covered with a tensioned


membrane beneath waves
169

8.2.4

Conservation of volume in the duct

The volume of water enclosed in the duct should be conserved. Using the definition
of the membrane displacement, this is expressed as

jA fjA (x)dx

(x)dx = 0

(8.37)

jS fjS (x)dx = 0

j.

(8.38)

The first term of the l.h.s of (8.38) is always true due to the definition of the fjA
modal functions in (8.17). This is not the case for the symmetric modes and for
that purpose, a method should be developed so that the symmetric modes satisfy
both the kinematic condition (8.28) and the conservation of volume
W =

M Z
X
j=1

jS cos
s

Sj x
dx
a

M
X
4s cos(j)
j=1

(2j 1)

jS

4s S 2S 3S
=
(
+
+ ...)
1
3
5
= 0.

(8.39)

(8.40)
(8.41)
(8.42)

This change of volume was the challenging part of this model and slowed down its
development. However, three solutions have been attempted and coded and there
follows a description of the methods used to solve this problem of conserving the
inner fluid volume constant.
An ideal solution would be to have a set of trapped modes whose amplitudes
could be adjusted to cancel out the change of volume from the symmetric modes.
Such modes would radiate energy only into the fluid near the membrane, hence
would not affect the energy conservation of the system. However for such mode(s)
to exist at a specific frequency, the determinant of the system (8.28) should be
null. This was checked and found not to be the case, necessitating a search for
other ways to deal with this problem.

Chapter 8 Two dimensional analytical model of a duct covered with a tensioned


170
membrane beneath waves
8.2.4.1

Uniform potential below the membrane

The first idea developed was to solve the system for the symmetric radiated modes
and then look for a set of modes that would cancel this out. In other words, find
a potential that satisfies all the physical boundary conditions of the problem but
whose effect on the membrane would be a displacement triggering a change of
volume opposite to that the one triggered by the symmetric modes.
A change of volume in the duct can be characterised by a uniform potential acting
on the membrane from below. The idea was to solve the kinematic equation (8.27)
with a uniform pressure of unit amplitude beneath the membrane. The change of
volume due to these modes is calculated and the process repeated with the uniform
pressure scaled by a factor so that the new potential obtained implies a change of
volume equal and opposite to that in the original calculation.
This was the first method developed and coded. It will be denoted hereinafter
as the UniP method The two others deal directly with finding symmetric modes
that simultaneously solve (8.28) and that do not imply a change of volume.

8.2.4.2

Weighted least-square method

The idea here is to try solving simultaneously (8.28) and (8.41). This brings an
over-determined system of M+1 equations for M unknowns. Writing this system
explicitly this gives for M = 3 (this will be used for explanation)

11 12 13
F


21 22 23 1 1

2 = F2 .

31 23 33
3
F3
4s
4s
4s
| {z } | {z }

3
5
{z
}
|
F

(8.43)

As previously stated in 6.4.4, such a system does not accept an exact solution. A
best-fit solution is obtained by using a least-square method which looks for the
solution that minimises the sum of the errors. This solution is found to be of the
form
= (Z T .Z)1 .Z T .F

(8.44)

Chapter 8 Two dimensional analytical model of a duct covered with a tensioned


membrane beneath waves
171
The above stands for the case that the errors have the same variance, meaning
that the same importance is given to any equations to be validated. The weighted
least-squares is the extension of the least-squares and a diagonal matrix of weight
W is included in the calculation of
= (Z T .W.Z)1 .Z T .W.F

(8.45)

By tuning the coefficients wii , it is possible to force one condition relative to the
others.

8.2.4.3

Matrix approximation

The last idea developed and coded is purely a mathematical approximation. Consider again the matrix of the system Z and the system obtained (8.43). The reasons
this system is not directly solvable with a traditional Gauss approach is that dimensions Z is not square and its line dimensions do not match the dimension of
. One way to tackle this is to solve the following system

1
F1
11 12 13

21 22 23 2 F2
= .

23
33
3 3
31
4s
4s
4s
0
4
0

3
5

(8.46)

with as small as the computational resource allows it in order to avoid overflow.


This way approaches the case of (8.43) but with a system that is solvable. This
will be denoted the MatApp method.

8.2.5

Code Layout

The previous model was coded on Iridis, the supercomputer at the University of
Southampton. The code, made of several routines detailed in Figure (8.2), was
written in Fortran 90. Inputs were the physical parameters of the problem: water
depth, submergence of the membrane, length of the duct and tension in the membrane. The number of eigenvalues N for the expansion series and the number of
modes M for the membrane deformation were also defined at this stage.
S,A S,A S,A S,A
The matching conditions at x = s used to determine the aS,A
m , bm , alm , clm , clm ,
were numerically checked, together with the zero velocity on the bottom and on the

Chapter 8 Two dimensional analytical model of a duct covered with a tensioned


172
membrane beneath waves

Defintion of the parameters s,h,d, M,N,Tx

Determination of the wavenumbers


series k1n, k2n and k3n and eigenfunctions
series f1n, f2n and f3n.

Calculation of the am and bm coefficients

Evaluation of the diffracted potentials f


for regions (1) and (2)

Calculation of the alm , blm and clm coefficients

Evaluation of the modal radiated potentials


flR for regions (1) and (2)
Calculation of the asymmetric complex
modal amplitudes from (8.22)
Calculation of the
symmetric complex
modal amplitudes
from (8.22)

Calculation of the
symmetric complex
modal amplitudes
from (8.37)

Calculation of the
symmetric complex
modal amplitudes
from (8.38)

Corrective potential
to cancel out the
change of volume

Evaluation of the radiated potentials f


for regions (1), (2) and (3) and the total
potential f
Figure 8.2: Code developed for the analytical model.

Chapter 8 Two dimensional analytical model of a duct covered with a tensioned


membrane beneath waves
173
outer and inner side walls of the duct. The other boundary conditions (kinematic
and dynamic free surface boundary conditions) were also numerically checked. A
example is given with the kinematic condition (8.27) later.

Figure 8.3: Reflection coefficient for h/d = 2 and s/d = 2. from Mei and
Black (1969), diffraction part of the present method.

Figure 8.4: As for Figure 8.3, but s/d = 4.

Chapter 8 Two dimensional analytical model of a duct covered with a tensioned


174
membrane beneath waves

Figure 8.5: As for Figure 8.3, but s/d = 6.

Figure 8.6: As for Figure 8.3, but h/d = 2.78 and s/d = 4.43.

Chapter 8 Two dimensional analytical model of a duct covered with a tensioned


membrane beneath waves
175

8.2.6

Case studies

8.2.6.1

Diffraction problem

The diffraction part of the model can be tested with previous studies made on the
diffraction of water waves by submerged obstacles. One key study is that of Mei
and Black (1969) in which scattering of normal incident waves by either a surface
or submerged rectangular obstacle is investigated. Potentials in different regions
are expressed by means of eigenfunction expansions in which the coefficients are
found using the matching and boundary conditions. Their numerical results for
the reflection coefficients are compared with the present model (N = 10) in Figures 8.3-8.6 and are in fair agreement with the present model.
The main comment from these results is the oscillatory nature of the reflection
coefficients and the effects of the dimensions of the problem on the period and
amplitude of these oscillations. Mallayachari and Sundar (1996) used a boundary
element method for the same kind of problem but with submerged obstacles of
various shapes (rectangular, trapezoidal and half-cylindrical). Their results agree
with those from Mei and Black (1969) for the rectangular obstacle.

8.2.6.2

Complete configuration and comparison of the proposed


methods

One intuitive test to carry out was the one with infinite tension in the membrane.
Figures 8.7-8.10 shows the computed reflection coefficient for that situation with
the complete code (three with correction and no correction) and compares it with
the computed values obtained for the diffraction case alone. The four configurations tested are the same as that of the diffraction situation.
Figures 8.3-8.7, 8.4-8.8, 8.5-8.9, and 8.6-8.10 are the same, the membrane with
infinite tension in the model behaving like a rigid plate. So far, there seems not to
be much difference between the three correction methods and the results without
any of them used.

Chapter 8 Two dimensional analytical model of a duct covered with a tensioned


176
membrane beneath waves

Figure 8.7: Reflection coefficient for h/d = 2 and s/d = 2. diffraction only,
Tx for the following methods of conservation of volume:    uniform potential; + + + least-squares; N N N matrix approximation; no correction.

Figure 8.8: As for Figure 8.7, but s/d = 4.

Chapter 8 Two dimensional analytical model of a duct covered with a tensioned


membrane beneath waves
177

Figure 8.9: As for Figure 8.7, but s/d = 6.

Figure 8.10: As for Figure 8.7, but h/d = 2.78 and s/d = 4.43.

Chapter 8 Two dimensional analytical model of a duct covered with a tensioned


178
membrane beneath waves
The boundary conditions at the free surface and on the duct walls were checked numerically, together with the continuity of potentials and velocities at the interface
between regions (2) and (3). For all the four methods (three for the conservation
of volume concern and the one without), this was seen to be valid.
An example of such numerical check is given in Figures 8.11 to 8.18 for the configuration s = 2m, h/d = 2 with h = 0.4m and Tx /gh2 = 0.15 and a wave frequency
f = 1Hz. The number of eigenvalues was chosen as N = 10 and the number of
modes M = 10. The code used for presentation of these figures is the one with no
correction of internal volume of fluid.

Figure 8.11: Symmetric (left column) and asymmetric (right column) of the
diffracted potentials (upper plots) and velocities (lower plots) at the interface
x = s; s = 2m, h/d = 2 with h = 0.4m and Tx /gh2 = 0.15 and wave
frequency f = 1Hz.

Chapter 8 Two dimensional analytical model of a duct covered with a tensioned


membrane beneath waves
179

Figure 8.12: Symmetric (left column) and asymmetric (right column) of the
radiated potentials (upper plots) and velocities (lower plots) at the interface
x = s; s = 2m, h/d = 2 with h = 0.4m and Tx /gh2 = 0.15 and wave
frequency f = 1Hz.

Numerical verification of the matching of the potentials and velocities at the interface x = s and 0 y d are shown in Figures 8.11 and 8.12 with
S,A
D
,
x
S,A
R
=
,
x

S,A
D,x =

(8.47)

S,A
R,x

(8.48)

for the diffracted and radiated potentials, respectively. The potentials (or velocities) are calculated at x = s using the expression for region (2) (points) and for
region (1) (line).

Chapter 8 Two dimensional analytical model of a duct covered with a tensioned


180
membrane beneath waves
Diffracted and radiated potentials are seen to be continuous at the interface between the two regions. The diffracted and radiated velocities are null on the
outside wall of the duct (x = s and h y d) and continuous at the interface, except over a small distance above it. This is typical of a singularity due to
the presence of the sharp corner, where the velocity at this point needs to use an
infinite number of eigenvalues to minimise the error.

Figure 8.13: Combined free surface dynamic condition (8.6) for the symmetric
(left plot) and asymmetric (right plot) diffracted potentials in region (1); s =
2m, h/d = 2 with h = 0.4m and Tx /gh2 = 0.15 and wave frequency f = 1Hz.

Figure 8.14: Combined free surface dynamic condition (8.6) for the symmetric
(left plot) and asymmetric (right plot) radiated potentials in region (1); s = 2m,
h/d = 2 with h = 0.4m and Tx /gh2 = 0.15 and wave frequency f = 1Hz.

Chapter 8 Two dimensional analytical model of a duct covered with a tensioned


membrane beneath waves
181
Figures 8.13 and 8.14 show the the dynamic condition (8.6) at the free surface in
region (1) for the diffracted and radiated potentials, respectively, where
S,A
D
,
y
S,A
R
=
,
y

S,A
D,y =

(8.49)

S,A
R,y

(8.50)

and Figures 8.15 and 8.16 show the same verification for region (2).

Figure 8.15: Combined free surface dynamic condition (8.6) for the symmetric
(left plot) and asymmetric (right plot) diffracted potentials in region (2); s =
2m, h/d = 2 with h = 0.4m and Tx /gh2 = 0.15 and wave frequency f = 1Hz.

Figure 8.16: Combined free surface dynamic condition (8.6) for the symmetric
(left plot) and asymmetric (right plot) radiated potentials in region (2); s = 2m,
h/d = 2 with h = 0.4m and Tx /gh2 = 0.15 and wave frequency f = 1Hz.

Chapter 8 Two dimensional analytical model of a duct covered with a tensioned


182
membrane beneath waves
The boundary condition at the free surface (8.6) is validated for both potentials
and both regions of interest. The symmetric and asymmetric pattern of the potentials are visible for the plots concerning the region above the membrane, e.g.
region (2).
Figures 8.17 and 8.18 show the numerical verification on the dynamic condition
(8.18) for regions (2) and (3), respectively.

Figure 8.17: Dynamic condition (8.18) for the symmetric (left plot) and asymmetric (right plot) for the first modal (l = 1) radiated potential in region (2);
s = 2m, h/d = 2 with h = 0.4m and Tx /gh2 = 0.15 and wave frequency
f = 1Hz.

Figure 8.18: Dynamic condition (8.18) for the symmetric (left plot) and asymmetric (right plot) for the first modal (l = 1) radiated potential in region (3);
s = 2m, h/d = 2 with h = 0.4m and Tx /gh2 = 0.15 and wave frequency
f = 1Hz.

Chapter 8 Two dimensional analytical model of a duct covered with a tensioned


membrane beneath waves
183
All those previous numerical verifications (and figures) were carried for all the
methods described above. At this stage, the UniP method was seen not to satisfy
the continuity of potentials and velocities at the interface x = s. This is likely
to be due to the inclusion of the uniform potential in the three regions. The
comparison of methods continues by taking into account the two main remaining
conditions, e.g. the energy conservation and the kinematic boundary condition on
the membrane (8.27), together with the volume conservation with the aim to try
a best fit (or compromise) to the problem.

Figure 8.19: Kinematic condition (8.27) for the symmetric (left plot) and
asymmetric (right plot) with no volume correction; s = 2m, h/d = 2 with
h = 0.4m and Tx /gh2 = 0.15 and wave frequency f = 1Hz.

Figure 8.20: Kinematic condition (8.27) for the symmetric (left plot) and
asymmetric (right plot) with the WLS volume correction; s = 2m, h/d = 2
with h = 0.4m and Tx /gh2 = 0.15 and wave frequency f = 1Hz.

Chapter 8 Two dimensional analytical model of a duct covered with a tensioned


184
membrane beneath waves
Figures 8.19, 8.20 and 8.21 show the numerical verification of the kinematic equation (8.27) for the code with no correction of volume, with the WLS method and
with the MatApp method, respectively.

Figure 8.21: Kinematic condition (8.27) for the symmetric (left plot) and
asymmetric (right plot) with the MatApp volume correction ( = 1015 ); s =
2m, h/d = 2 with h = 0.4m and Tx /gh2 = 0.15 and wave frequency f = 1Hz.

The asymmetric verification is unchanged as the different methods for the volume
conservation problem deals only with the symmetric modes. For both symmetric
and asymmetric parts, discrepancies occur around x = s and this is consistent

with the numerical verification of the continuity of potential and velocities where
it was seen that the potentials at these locations needed an infinite number of
eigenvalues to counter the singularity due to the sharp corner. For all three cases,
the membrane equation is verified along the length of the membrane, although
discrepancies centred around x = 0 and over a length of x = s/20 are visible for
the MatApp method.

Chapter 8 Two dimensional analytical model of a duct covered with a tensioned


membrane beneath waves
185

Figure 8.22: Comparison of the change of volume induced by the symmetric


and efficiency of the proposed methods; (a) no correction of volume, (b) WLS
method; (c) Unip method; (d) MatApp ( = 1015 ); s = 2m, h/d = 2 with
h = 0.4m and Tx /gh2 = 0.15.

Figure 8.22 shows the change of volume induced by the symmetric radiated modes
of the membrane vibration, as a percentage of the undisturbed volume (in two
dimensions). The MatApp method seems to be the one treating the volume conservation issue with best efficiency, for which almost no change of volume due to
the symmetric modes can be obtained. As can be seen in Figure 8.23, the energy
relation
Kr2 + Kt2 = 1
(8.51)
is numerically verified for the range of studied wave frequencies with error of less
than 1% for methods (a), (b) and (c) with the notations of the same figure. The
Unip method, on the contrary, does not satisfy it, as the inclusion of a (uniform)
potential corresponds to the inclusion of more energy in the system. The MatApp
correction method was then the one chosen to compare the results of the analytical
model with experimental data.

Chapter 8 Two dimensional analytical model of a duct covered with a tensioned


186
membrane beneath waves

Figure 8.23: Comparison of the error on the energy conservation in the system
for the proposed methods; (a) no correction of volume, (b) WLS method; (c)
Unip method; (d) MatApp ( = 1015 ); s = 2m, h/d = 2 with h = 0.4m and
Tx /gh2 = 0.15.

8.3

Experimental verification

Outputs from this model were compared with results of the experiment carried
out in the wide basin for the configuration (test 8) for which all pipes of the PTO
were closed, tending to the limit case of a submerged closed duct as for the present
modelled geometry. The internal static pressure head was p0 = 0.02m, determining
the coefficients cSj0 in (8.22). The membrane submergence d was approximated as
the equivalent membrane submergence depth for that internal static pressure head,
so for that case, d = 0.363m. The tension Tx used to run the code was the one
needed to counter the Poissons effect from the elongation in the y-direction due
to the internal pressure applied. Values of N=10 and M=10 were chosen. More
runs with higher values for both of these parameters were seen not to change the
results presented below.

Chapter 8 Two dimensional analytical model of a duct covered with a tensioned


membrane beneath waves
187
Figure 8.24 shows the predicted reflection coefficient and the measured one with
conditions such as test 8. The orders of magnitude are similar, but there is no
clear evidence of agreement between calculated and measured values.

Figure 8.24: Ratio of the reflected power, Pr , to incident wave power Pi for
the duct with an internal static pressure head of 2cm. Line: present model;
symbols: first harmonic analysis of wave gauge signals.

/ A
40
20
0
0
0.5
f (Hz)

1
1

0.5

0
x (m)

0.5

Figure 8.25: Response of the membrane (/A) as a function of wave frequency


f and horizontal coordinate x, parameters such as in test 8.

Chapter 8 Two dimensional analytical model of a duct covered with a tensioned


188
membrane beneath waves

6
4
2
0
0
0.5
f (Hz)

1
1

0.5

0
x (m)

0.5

20
15
10
5
0
0
0.5
f (Hz)

1
1

0.5

0
x (m)

0.5

4
2
0
0
0.5
f (Hz)

1
1

0.5

0
x (m)

0.5

Figure 8.26: Modal response amplitude as function of wave frequency f and


horizontal coordinate x; upper plot: first symmetric mode, middle plot: second
symmetric mode; lower plot: third symmetric mode.

Chapter 8 Two dimensional analytical model of a duct covered with a tensioned


membrane beneath waves
189

40
30
20
10
0
0
0.5
f (Hz)

1
1

0.5

0
x (m)

0.5

15
10
5
0
0
0.5
f (Hz)

1
1

0.5

0
x (m)

0.5

3
2
1
0
0
0.5
f (Hz)

1
1

0.5

0
x (m)

0.5

Figure 8.27: As for Figure 8.26 for asymmetric modes.

Chapter 8 Two dimensional analytical model of a duct covered with a tensioned


190
membrane beneath waves
The other available data was the membrane displacement. It is shown in Figure
8.25, together with the modal responses of the first three symmetric (Figure 8.26)
and asymmetric (Figure 8.27) modes, for all frequencies. The free bulge speed
for that configuration was c = 2.10m/s, leading to a natural frequency of f2 =
1.05Hz as detailed earlier. This is verified by mapping the experimental membrane
displacement for that frequency, as shown in Figure 8.28. Three distinct bulge
patterns are visible as expected, and the difference in their shapes comes from the
fact that this configuration is not exactly the same as one with purely closed ends.
Again there is no agreement with the calculated values of membrane displacement.

Figure 8.28: Membrane displacement (normalised by the incident wave amplitude A) for configuration of test 8 and f =1.05Hz.

8.4

Conclusions

An analytical model based on an expansion series for the different potential concerned for the geometry shown in Figure 8.1 has been developed, following the
work initiated by Cho and Kim (1998). One of the important issue to be addressed was that of the conservation of volume in this closed geometry. For that
purpose, three methods have been proposed to tackle the problem of change of
volume induced by the symmetric modes of the membrane vibration. With regards to the analyses carried out, the MatApp method seemed to be the best fit
for this problem. All the boundary conditions were numerically checked, together
with the energy conservation and the change of volume. This is why this method
was the one chosen for comparison with experimental results. Other methods were
seen to be invalid for one or more of these numerical tests.

Chapter 8 Two dimensional analytical model of a duct covered with a tensioned


membrane beneath waves
191
One disadvantage of this method is that increasing the number of modes M of the
membrane vibration increases the dimension of the matrix of the system (8.46).
This yields an increase of error when approximating the membrane equation (8.27)
by the linear system(8.46). This can be avoided by taking a smaller coefficient,
if computational resources allow it.
Results of reflection coefficient and membrane displacement obtained from this
model did not agree with the experimental results obtained with the configuration
of chapter 7 where all copper pipes were closed. This configuration was the closest
one to that of Figure 8.1.
The mapped displacement in Figure 8.25 suggests that the bulge wavelengths
obtained with this model are much shorter than those measured. One possible
explanation is that the membrane stress used to run this code was too low, even
though it has a physical meaning in a two dimensional analysis. Three dimensional
effects may have to be considered, and in particular the fact that the restoring force
due to the bending of the membrane in the width direction yields a much higher
tension in the rubber than that used so far.

/ A

20
15
10
5
0
0
0.5
f (Hz)

1
1

0.5

0
x (m)

0.5

Figure 8.29: Response of the membrane (/A) as a function of wave frequency


f and horizontal coordinate x, parameters such as in test 8, with the longitudinal
tension used 2.7 times greater than that due to Poissons ratio effect only.

Chapter 8 Two dimensional analytical model of a duct covered with a tensioned


192
membrane beneath waves

6
/ A
4
2
0
0
0.5
f (Hz)

1
1

0.5

0
x (m)

0.5

Figure 8.30: As for Figure 8.29 with a the tension multiplied by a factor of
11.

Figures 8.29 and 8.30 show the displacement obtained by multiplying the tension
used before by factors of 2.7 and 11 respectively. The membrane displacement
profile exhibits longer wavelengths, approaching the experimental case. However
there is so far no clear understanding of the longitudinal tension needed to characterise the experimental case so far. One possibility could be to work out an width
averaged membrane model for the determination of a more appropriate tension to
run the code with, as part of future work.

Chapter 9
Conclusions and future work
9.1

Achievements and perspectives

This thesis reported experiments carried out on a rectangular duct covered with
a rubber membrane, set on the seabed and aligned with the wave propagation
direction. Two different versions were investigated: the first one with closed ends
and the second one with a power take-off attached at the duct stern. The main
goal of this work was to verify the one dimensional model used for the Anaconda
wave energy converter for a novel configuration, and explain any mismatch due to
its limitations.
As mentioned in section 3.3.5, the available experimental data regarding behaviour
of such a structure beneath waves was poor. The methodology developed within
this study has filled this gap by means of pressure transducers placed at the bow
and stern of the duct, together with laser sensors carefully set along the length of
the membrane. For the first time to the authors knowledge, a complete experimental characterisation of the membrane displacement and pressure variation was
obtained for such a structure.
The experiments carried out on the closed end version of 5m in a narrow flume
with the thick (1mm) membrane exhibited resonant behaviours for frequencies
fn = nc/2L, at which the bulge wavelength is an integer times 2L. The membrane
displacement was made of nodes and antinodes characteristic of such resonance.
This measured displacement yielded the measurement of the amplitude of the pressure in the duct, following the load-deflection relationship. The magnitude of this
193

194

Chapter 9 Conclusions and future work

pressure at the stern for these resonant frequencies was seen to reach up to 2.8
times that of the incident wave at a depth equal to the membrane submergence.
By means of a least squares method, the internal pressure was decomposed into
three components and the pressure wave explained as a partial standing wave near
resonance. This was explained with a one dimensional model where the incoming
(water) wave was assumed to be undisturbed along its travel above the membrane.
On the contrary, this model failed to predict the membrane displacement and
pressure behaviour for the thinner (0.2mm) of the tested membranes. For that
membrane, values of free bulge speeds were much lower than those of the thicker
membrane, yielding a profile of membrane displacement like a ripple bed. This
measured displacement agreed with the predicted one over a length of the duct
equal to the bulge wavelength. After that, the incoming wave lost its energy by its
interactions with this non uniform bed and this was confirmed by measuring the
surface displacement at mid-length of the duct. The main conclusion from this
disagreement was that the one dimensional model used was not effective for devices
whose length is of a much higher order of magnitude than the bulge wavelength,
that is, the hypothesis of non disturbance of the incoming wave is no longer valid.
Moreover, the experimental observation of ripples, of much higher amplitude than
those propagating on the thick membrane, were found to play an important role
in the way the energy carried by the bulges is dissipated during their travel.
The same duct, reduced to a length of 2m, was equipped with a PTO and tested
in a wide basin with the thick membrane. Capture widths of up to 2.2 times the
duct width were obtained and explained with the one dimensional theory. Maximum power outputs were obtained for frequencies maximising the flow rate that
went through the pipes of the PTO. These frequencies corresponded to either the
case that the water wave speed and bulge speed matched, or the wave frequency
matched a natural frequency fn . For the first situation, the amplitude of the
internal pressure was seen to increase along the length of the duct, while as before, the second situation exhibited characteristic patterns of a resonant behaviour.
This showed that the theory previously tested on the Anaconda rubber tube works
for another configuration. Using a seabed configuration (hence in shallow waters)
broadens the possible usage of such a concept. Bottom-mounted, this device would
not interfere with commercial and leisure activities present nearshore. Because of
wave refraction, there is no concern in capturing power contained in waves coming
from many different directions. Practically, this rectangular structure could be

Chapter 9 Conclusions and future work

195

towed on site and then sunk by adding ballast. The rubber membrane could then
be installed.
To go one step forward in the general study of the interactions of water waves
with a submerged duct covered with a stretched membrane, a two dimensional
analytical model was developed. It is based on expansion series for the diffracted
and radiated potentials in each regions, the series coefficients being obtained by
matching the potentials and velocities at the boundaries. The problem of conserving the inner water volume was addressed and several solutions were proposed.
One of them was chosen as a best compromise and its outputs compared with the
experimental data obtained from the tests carried on the configuration presented
in chapter 7 with all pipes of the PTO closed. There was very poor agreement with
regards to membrane displacement and reflection coefficient. Calculated profile of
the membrane displacement was seen to be many more wavelengths greater than
the experimental case. The tension in the membrane used to run the code was
seen to be of importance in this number of wavelengths, suggesting that results
obtained when using a higher tension in the membrane could potentially match
the experimental ones.

9.2

Future work

Testing a new wave energy converter at a laboratory scale is only the first step
within the overall development, and this is what has been done in this study. The
tests carried on the duct equipped with the PTO showed that the bulge wave
concept can be used to harness wave power when used on a seabed configuration.
However, such potential needs to be further examined if a commercial full-scale
prototype is intended for design. Particular issues to be addressed are:
Scale effects if longer devices are intended to be used, in particular the ratio
of the water wavelength to that of the bulge wave to avoid situations observed
in the thin membrane test. An useful experiment would be one with the thin
membrane clipped on top of a closed ends duct shorter than the one tested
in this study. The ratio of the free bulge speed in the thick membrane to its
length was around
c
0.3
2L

(9.1)

196

Chapter 9 Conclusions and future work


so an experiment with the same ratio for the thin membrane would induce
a length of L = 0.6m. This would be a way to check the validity of the
argumentation developed in section 6.5 regarding the influence of the ripplelike profile of the membrane displacement on the incident water wave.

Design of a more appropriate (bottom-mounted and fully submerged) power


take-off. The PTO used in this study was to quantify the power that could
be absorbed by a linear dashpot but not to use it. A new design, aimed at
harnessing this power, needs to be bottom-mounted and fully submerged for
taking full advantages of a non-visible wave energy converter.
Rubber life in seawater compared to freshwater and the role of this salty
environment on rubber properties. Will the rubber properties and behaviour
be the same in seawater? What is the influence of the micro-organisms
growing on the membrane? Will this affect the cost of the rubber, initially
thought to be cheap? These questions are among those that need to be
addressed by further research.
Another aspect that needs to be addressed by further studies is that of ripples
propagating on the tensioned membrane. For the thinner of the tested membranes,
faster travelling ripples were experimentally observed ahead of bulge waves. Such
ripples have been widely studied for the case of surface waves as for instance by
Munk (1955); Longuet-Higgins (1963, 1995); Jiang et al. (1999) and many more.
When gravity waves reach their maximum steepness, the surface tension is unevenly distributed and localised near the wave crests and this travelling stress is
the source of capillary waves triggered ahead of the crests. The quoted previous works were carried out with respect to the surface tension at the free surface
(Ts =0.0744N/m) yielding a characterisation of maximum steepness, ripple amplitude and wavelength based on that value. A similar study for different surface
tensions, closer to the thinner membrane ought to be carried out to understand
specifically the influence of ripples on bulge propagation. The displacement of
the thinner membrane was of a higher order of magnitude than the thicker membrane, leading to the generation of steeper bulges and capillary waves. Because
the displacement of the thicker membrane was lower, no capillary waves were experimentally observed.
Finally, attention has to be given on to theoretical model with more experiments
carried out with the exact same configuration. The main task to be undertaken is
that of finding the correct longitudinal tension to run the code with. The effects

Chapter 9 Conclusions and future work

197

of including the third dimension in such a model was mentioned in section 3.3.4.
The advantages of this inclusion might be similar for the present model. Working
out a width averaged membrane model to determine an equivalent tension to use
in the membrane can be the step before that.

References
A.G. Abul-Azm. Wave diffraction through submerged flexible breakwaters. Ocean
Engng., 23(5):403422, 1996.
AW-Energy. http://www.aw-energy.com/.
AWS Ocean. http://www.awsocean.com.
H.F. Bauer. Hydroelastic vibrations in a rectangular container. International
Journal of Solids and Structures, 17(7):639652, 1981.
C. E. Blenkinsopp. Air Entrainment, Splash and Energy Dissipation in Breaking
Waves. PhD thesis, School of Civil Engineering and the Evironment, University
of Southampton, UK, 2007.
R. Boud. Status and Research and Development Priorities, Wave and Marine Accessed Energy. Dti Report # FES-R-132, AEAT Report # AEAT/ENV/1054,
UK Dept. of Trade and Industry (DTI), 2003.
M. Briggs, W. Ye, Z. Demirbilek, and J. Zhang. Field and numerical comparisons
of the RIBS floating breakwater. Journal of Hydraulic Research, 40(3):289301,
2002.
M.J. Briggs. Analytical and numerical models of the RIBS XM99 ocean-scale prototype. Technical report, US Army Corps of Engineers-Coastal and Hydraulics
Laboratory, 2001.
L.L. Broderick and C.H. Jenkins. Experimental investigation of fluid-filled membrane breakwater. Journal Waterway, Port, Coastal Ocean Engng., 119(7):
639656, 1993.
L.L. Broderick and J.W. Leonard. Nonlinear response of membranes to ocean
waves using boundary and finite elements. Ocean Engng, 22(7):731745, 1995.

199

200

REFERENCES

L. Cameron, R. Doherty, A. Henry, K. Doherty, J. Vant Hoff, D. Kaye, D. Naylor,


S. Bourdier, and T. Whittaker. Design of the next generation of the oyster wave
energy converter. In 3rd International Conference on Ocean Energy, Bilbao,
Spain, 6-8 October 2010.
H. Chanson. Some hydraulic aspects during overflow above inflatable flexible membrane dam. Report ch47/96, Departement of Civil Engineering, The University
of Queensland, Australia, May 1996.
J. R. Chaplin, F. J. M Farley, M. E. Prenctice, R. C. T. Rainey, S. J. Rimmer, and
A. T. Roach. Development of the ANACONDA all-rubber WEC. In European
Wave and Tidal Energy Conference, Porto, Portugal, 2007a.
J.R. Chaplin, F.J.M. Farley, and R.C.T. Rainey. Power conversion in theANACONDA WEC. In International Workshop on Water Waves and Floating
Bodies, Plitvice, Croatia, 15-18 April 2007b.
J.R. Chaplin, V. Heller, F.J.M. Farley, G.E. Hearn, and R.C.T. Rainey. Laboratory testing the Anaconda. Philosophical Transactions of the Royal Society of
London, 370:403424, 2012.
Checkmate SeaEnergy Ltd. http://www.checkmateseaenergy.com.
I.H. Cho, S.T. Kee, and M.H. Kim. The performance of flexible-membrane wave
barriers in oblique incident waves. Applied Ocean Research, 19(3-4):171182,
1997.
I.H. Cho, S.T. Kee, and M.H. Kim.

Performane of dual flexible membrane

wave barrier in oblique waves. Journal of Waterway, Port, Coastal Ocean


Engineering., 124(1):2130, 1998.
I.H. Cho and M.H. Kim. Interactions of a horizontal flexible membrane with
oblique incident waves. J. Fluid Mech., 367:139161, 1998.
I.H. Cho and M.H. Kim. Wave deformation by a submerged flexible circular disk.
Applied Ocean Research, 21:263280, 1999.
I.H. Cho and M.H. Kim. Interactions of horizontal porous flexible membrane with
waves. Journal of Waterway, Port, Coastal Ocean Engineering, 126(5):245253,
2000.
J. Chu, W. Guo, and S.W. Yan. Geosynthetic Tubes and Geosynthetic Mats:
Analyses and Applications. Geothechnical Engineering Journal of the SEAGS
& AGSSEA, 42(1):5665, 2011.

REFERENCES

201

A.M. Cornett. A global wave energy resource assessment. In Proceedings of the


Eighteenth International Society of Offshore and Polar Engineers Conference,
pages 318326, Vancouver, BC, Canada, 6-11 July 2008.
S. Das and K.F Cheung. Coupled boundary element and finite element model for
fluid-filled membrane in gravity waves. Engineering Analysis with Boundary
Elements, 33:802814, 2009.
Robert G. Dean and Robert A. Dalrymple. Water wave mechanics for engineers
and scientists. World Scientific, Advanced Series on Ocean Engineering-Volume
2, 1991.
K. den Boer. Estimation of incident and reflected wave characteristics of perpendicular wave action. Research report s43411981, Delft Hydraulics Laboratory,
April 1981.
F.D.E. Dewi, S.I. Liapis, I. Stergios, and R.H. Plaut. Three-dimensional analysis
of wave attenuation by a submerged, horizontal, bottom-mounted, flexible shell.
Ocean Engineering, 26:813839, 1999.
Antonio F. de. O. Falcao. Wave energy utilization: A review of the technologies.
Renewable and Sustainable Energy Reviews, 14:899918, 2010.
F.J.M Farley and R.C.T. Rainey. Anaconda the bulge wave sea energy converter.
November 2006.
Finavera Renewables. http://www.finavera.com.
M. Folley and T.J.T. Whittaker. Analysis of the nearshore wave energy resource.
Renewable Energy, 34:17091715, 2009.
M. Folley, T.J.T. Whittaker, and A. Henry. The effect of water depth on the
performance of a small surging wave energy converter. Ocean Engineering, 34
(8-9):12651275, 2007a.
M. Folley, T.J.T Whittaker, and J. Vant Hoff. The design of small seabed-mounted
bottom-hinged wave energy converters. In European Wave and Tidal Energy
Conference, Porto, Portugal, 2007b.
H.D Frederiksen. Wave attenuation by fluid-filled bags. Journal of the Waterways,
Harbors and Coastal Engineering Division, 97:7390, 1971.

202

REFERENCES

M.R. Hann, J.R. Chaplin, and F.J.M Farley. Assessment of a multi-cell fabric
structure as an attenuating wave energy converter. In World Renewable Energy
Congress, Linkoping, Sweden, 8-13 May 2011.
W.R. Hawthorne.

The early development of the dracone flexible barge.

Proceedings of the Institution of Mechanical Engineers, 175:5283, 1961.


C. Huang, W. Lou, C. Tsai, T C. Wu, and H Y. Lin. Mechanical properties of
polymer thin film measured by the bulge test. Thin Solid Films, 515:72227226,
2007.
T. Ijima and T. Uwatoko. Wave Interception by Sea-Ballon Breakwater. Memoirs
of the Faculty of Engineering, Kyushu University, 45(4), 1985.
L. Jiang, H.J. Lin, W. Schultz, and M. Perlin. Unsteady ripple generation on steep
gravity-capillary waves. Journal of Fluid Mechanics, 386:281304, 1999.
T.H.G. Jongeling and N.H. Rovekamp. Storm surge barrier Ramspol. In XXVIII
IAHR Congress, Graz, Austria, 22-27 August 1997.
S.T. Kee, D.S. Kim, K.N. Hwang, J.O. Han, and N.S. Hong. The performance evaluation of a floating pontoon with triple vertical porous-membrane breakwater
in the oblique seas. In Thirteenth International Offshore and Polar Engineering
Conference, pages 783789, Honolulu, USA, May 25-30 2003.
S.T. Kee and M.H. Kim. Flexible membrane wave barrier. part 2. floating/submerged buoy-membrane system. Journal of Waterway, Port, Coastal Ocean
Engng., 123:8290, 1997.
M.H. Kim, B.B. Edge, S.T. Kee, and L. Zhang. Performance evaluation of buoymembrane wave barriers. In 25th Int. Conf. on Coast. Engrg., 1996.
M.H. Kim and S.T. Kee. Flexible membrane wave barrier. part 1. analytic and
numerical solutions. J. Waterway, Port, Coastal Ocean Engng., 122:4653, 1996.
M. Larson, C. Simonson, R. Besant, and P. Gibson. The elastic and moisture
transfer properties of polyethylene and polypropene membranes for use in liquidto-air-energy exchangers. Journal of Membrane Science, 302:136149, 2007.
J.F. Lee and C.J. Chen. Wave interaction with hinged flexible breakwater. Journal
of Hydraulic Research, 28(3):283295, 1990.
W.K. Lee and E.Y.M. Lo. Surface-penetrating flexible membrane in wave barriers
of finite drafts. Ocean Engineering, 29:17811804, 2002.

REFERENCES

203

S.I. Liapis, L.S. Constantine, and J.D. Trowbridge. On the use of a flexible structure as a breakwater. In 15th International Conference on Ocean, Offshore
Mechanics and Arctic Engineering, pages 4147, Florence, Italy, 1996.
J. Lighthill. Waves in fluids. Cambridge Mathematical Library, 1978.
E.Y.M Lo. Flexible dual membrane wave barrier. Journal of Waterway, Port,
Coastal Ocean Engineering, 124(5):264271, 1998.
M.S. Longuet-Higgins. The generation of capillary waves by steep gravity waves.
Journal of Fluid Mechanics, 16(1):138159, 1963.
M.S. Longuet-Higgins. Parasitic capillary waves: a direct calculation. Journal of
Fluid Mechanics, 301:79107, 1995.
V. Mallayachari and V. Sundar. Wave Transformation Over Submerged Obstacles
in Finite Water Depths. Journal of Coastal Research, 12(2):477483, 1996.
S.R. Massel. Harmonic generation by waves propagating over a submerged step.
Coastal Engineering, 7:357380, 1983.
M. McIver. Diffraction of water waves by a moored, horizontal, flat plate. J.
Engng. Maths, 19:297320, 1985.
C.C. Mei and J.L. Black. Scattering of surfaces waves by rectangular obstacles in
waters of finite depths. J. Fluid Mech., 38(3):499511, 1969.
G. Mrk, S. Barstow, M.T. Pontes, and A. Kabuth. Assessing the global wave energy potential. In 29th International Conference on Ocean, Offshore Mechanics
and Arctic Engineering, pages 447454, Shanghai, China, 6-11 June 2010.
W. Munk. High frequency spectrum of ocean waves. Journal of Maritime Research,
14:302314, 1955.
S. Neelamani, D. Hauder, H.E. Krogstad, S. Lehner, J.A.J. Monbaliu, and L. Wyatt. Wave transmission and reflection characteristics of a rigid surface and
submerged horizontal plate. Ocean Engineering, 19(4):327341, 1992.
T. Ohyama, T. Kiyokawa, T. Uda, and Y. Murai. Transmission and reflection
characteristics of waves over a submerged flexible mound. Coastal Engineering
in Japan, 32(1):5368, 1989.
M. Patarapanich and H.F. Cheong. Reflection and transmission characteristics
of regular and random waves from a submerged horizontal plate. Coastal
Engineering, 13:161182, 1989.

204

REFERENCES

A.C. Phadke and K.F. Cheung. Response of bottom-mounted fluid-filled membrane in gravity waves. Journal of Waterway, Port, Coastal Ocean Engineering,
125(6):294303, 1999.
A.C. Phadke and K.F. Cheung. Resonance and response of fluid-filled membrane
in gravity waves. Applied Ocean Research, pages 1528, 2001.
A.C Phadke and K.F Cheung. Nonlinea Response of Fluid-Filled Membrane in
Gravity Waves. Journal of Engineering Mechanics, 129(7):739750, 2003.
M. Rahm. Underwater substation system for wave energy converters. Acta Universitatis Upsaliensis 711, Digital Comprehensive Summaries of Uppsala Dissertations from the Faculty of Science and Technology, 2010.
A.W. Stahl. The utilization of the power of ocean waves. Trnasactions of the
American Society of Mechanical Engineers, 13:438506, 1892.
D.G. Stamos. Experimental analysis of the interaction of water waves with flexible
structures. PhD thesis, Virginia Polytechnic Institute and State University,
Blacksburg, Virginia, USA, 2000.
D.G. Stamos, M.R. Hajj, and P. Demetri. Performance of hemi-cylindrical and
rectangular submerged breakwaters. Ocean Engineering, 30:813828, 2003.
M. Tanaka, T. Ohyama, T. Kiyokawa, T. Uda, and A. Omata. Characteristics
of wave dissipation by flexible submerged breakwater and utility of the device.
Coastal Engineering, 2:16131624, 1992a.
M. Tanaka, T. Ohyama, T. Kiyokawa, T. Uda, and A. Omata. Wave control by
flexible mound. Technical Report 11, Shimizu Tech. Res. Bull., March 1992b.
J. Vlassak. New experimental techniques and analysis methods for the study
oh mechanical properties of materials in small volumes. PhD Thesis, Stanford
University, 1994.
T.J.T. Whittaker, D. Collier, M. Folley, M. Osterried, A. Henry, and M. Crowley.
The development of oyster-a shallow water surging wave energy converter. In
European Wave and Tidal Energy Conference, Porto, Portugal, 2007.
A.N. Williams. Flexible floating breakwater. Journal Waterway, Port, Coastal
Ocean Engng., 117(5):429450, 1991.
X. Xiang, X. Chen, and J. Vlassak. Plane-strain bulge test for thin films. J. Mater.
Res., 20(9):23602370, 2005.

REFERENCES

205

X. Yu and A.T. Chwang. Analysis of wave scaterring by submerged disk. J. Engng.


Mech., 119:18041817, 1993.
R. Zhao. A complete linear theory for a two-dimensional floating and liquid-filled
membrane structure in waves. Journal of Fluids and Structures, 9:937956,
1995.
R. Zhao and J.V. Aarsness. Numerical and experimental studies of a floating and
liquid-filled membrane structure in waves. Ocean Engng, 25(9):753765, 1998.

Appendix A
Power take-off test in the narrow
flume
A.1

Introduction

The first attempt at using the power take-off system described in section 4.5 was
for the 5m duct in the narrow flume. For this test, the slug of water was enclosed
in a circular rigid tube of cross-sectional area Sd = 0.021m2 , yielding a ratio of
cross-sectional area = 2.75. Figure A.1 shows this and the power take-off installation. The complete set-up is similar to that of Figure 6.1, the only difference is
that the PTO was connected at the duct stern.
Power capture values were very low for this test. This is explained fairly well
with the 1D analytical model. This test with a wider range of impedances than
those used in the wider basin helped to provide a better understanding of the
PTO. This appendix shows plots of phase lag in the PTO and capture widths
obtained for the test conditions detailed in Table A.1. The general explanation
has been given in section 7.2.2.

A1

A2

Appendix A Power take-off test in the narrow flume

membrane

copper
pipes

circular
rigid tube

Figure A.1: Power take-off set-up in the narrow flume.

A.2

Experimental set-up and test conditions

Internal static

Wave

Number of

Number

Test n

pressure head
(m)

frequencies
(Hz)

wave
amplitudes

of
tests

0.07m

12.33

0.7-1.7

0.07m

6.16

0.7-1.7

0.07m

4.11

0.6-1.7

0.07m

3.08

0.6-1.7

0.07m

1.54

0.6-1.7

0.07m

1.02

0.6-1.7

0.07m

0.77

0.7-1.7

0.07m

0.7-1.7

Table A.1: PTO experiment in narrow flume: test conditions

Appendix A Power take-off test in the narrow flume

A3

As before, the signals obtained from the wave gauges (in the flume and in the
PTO), the pressure transducers (water and air) and the laser sensors were found to
be predominantly monochromatic. Hence the data analysis carried out hereinafter
neglects higher harmonics.

A.3

Linearity of the power take-off

Figure A.2: Air pressure in the chamber as a function of the water-air interface
elevation for several dimensionless impedances Z. The wave frequency is f =
0.95Hz and incident wave amplitude A =0.025m.

A4

Appendix A Power take-off test in the narrow flume

Figure A.3: Phase lag between PTO water level aPTO and PTO air pressure
pPTO as a function of dimensionless impedance Z for several wave frequencies;
measured, from (4.25).

Figure A.2 shows the air pressure in the chamber as a function of the water-air
interface displacement for various dimensionless impedances. As the impedance
increases, the air pressure and water displacement are no longer in quadrature.
For the case where all pipes are closed, the magnitude of both values seems higher
than for lower impedances.
The phase lag between water-air displacement and pressure is plotted for various
impedances and wave frequencies in Figure A.3 and compared with the theoretical
values calculated from 4.25) with again fair agreement.

Appendix A Power take-off test in the narrow flume

A.4

A5

Power capture

Capture widths for the tested impedances are presented in Figure A.4 and compared with the values calculated with (5.43). The magnitudes of the captured
power are very low and can be explained with the 1D model used in this study,
although the agreement between calculated and measured values is poorer in this
case than for the experiment in the wide flume.

Figure A.4: Capture widths for the duct equipped with the PTO and placed in
the narrow flume for various dimensionless impedances Z; Grey line: calculated
from (5.43), line and symbols: measured.

Appendix B
Analytical model of duct covered
with a tensioned membrane: the
maths
B.1

Introduction

This section is to provide the details of the analytical model from chapter 8.

B.2
B.2.1

Closed ends configuration


Diffraction problem

The symmetric diffraction potentials are given by Cho and Kim (1998) for their
geometry (region (3) not closed) as
S(1)

S(2)
D

X
ig 1 10 x
aSn e1n (x+s) f1n (y)},
{ e
f10 (y) +
2
n=0

ig X S
=
b cosh(2n x)f2n (y),
n=0 n

with
1n =

ik1

n=0

k1n

n1

B1

(B.1)
(B.2)

(B.3)

Appendix B Analytical model of duct covered with a tensioned membrane: the


B2
maths

2n =

ik2

n=0

k2n

n1

3n = k3n

(B.4)
(B.5)

n0

where the eigenfunctions f1n and f2n are defined as


cos k1n (y + h)
cos k1n h
cos k2n (y + d)
=
cos k2n d

f1n =

n 0,

(B.6)

f2n

n 0,

(B.7)

and the eigenvalues k1n and k2n are solutions of


2
k1 tanh(k1 h) = ,
g
2
k1n tanh(k1n h) =
g

k10 = ik1
k20 = ik2

(B.8)

(B.9)

n 1,

2
k2 tanh(k2 d) = ,
g
2
k2n tanh(k2n d) =
g
and k10 and k20 are set as

n 1,
.

(B.10)

The coefficients aSn and bSn can be calculated by matching the two solutions at
x = s. The continuity of potential at x = s leads to

X
X
1 10 s
S
bSn cosh(2n s)f2n (y)
an f1n (y) =
e f10 (y) +
2
n=0
n=0

d y 0. (B.11)

Multiplying (B.11) by f2m (y) and integrating over [d,0] allows determination of
bSn as:

bSm =

1 10 s
e Cm0
2

(2)

(2)
Nm

m=n

cosh(2m s)Nm

where
Z

aSm Cmk

k=0

f2n (y)f2m (y)dy =

0 m 6= n

(B.12)

(B.13)

Appendix B Analytical model of duct covered with a tensioned membrane: the


maths
B3
(2)
Nm
=

Cmn =

d
2 cos2 (k

2m d)

{1 +

sin(2k2m d)
},
2k2m d

(B.14)

f1n (y)f2m (y)dy


d

2k1n sin[k1n (h d)]


+
2
2
k2m
k1n
sin(dk2m + hk1n ) sin(dk2m hk1n )
+
}.
k2m + k1n
k2m k1n
1

= (2 cos(k1n h) cos(k2m d)) 2 {

(B.15)

The continuity of velocity SD /x and the boundary conditions (8.8) at x = s


leads to the equality:

X

2n bSn sinh(2n s)f2n (y)

X
1
10 s
1n aSn f1n (y) =
10 e f10 (y) +

n=0

n=0

h y d

d y 0
(B.16)
(B.17)

Defining

nm =

f1n (y)f1m(y)dy

(B.18)

and
Z

f1n (y)f1m (y)dy =

(1)
Nm
=

h
2 cos2 (k

1m h)

{1 +

(1)
Nm

m=n

(B.19)

0 m 6= n

sin(2k1m h)
},
2k1m h

(B.20)

the aSm coefficients can be written after multiplying (B.17) by f1m (y) and integrating it over [h,d] as:

X
1 10 S00
1k k0 S

e ( 1 1) m = 0
a
=

(1) k
m
2
N0

N
10
0
k=0

1k km S
1 10 e10

a
=

m
(1) k
(1) 0m
2

N
m
m
1m
1m
k=0

m1

(B.21)

Appendix B Analytical model of duct covered with a tensioned membrane: the


B4
maths
Next, multiplying (B.16) by f1m (y) and integrating it over [d,0] using (B.12)
leads to

S
S
X
1
00
F0k
1k k0 S
1 10 F00

1) e10 a (1)
e
(
am +
a
=

1
(1)

2
N0
2
10 N0
N0
k=0

m=0

S
X

1 e10
Fmk
1k km S

S
S

a
=

(Fm0
+ 10 0m ) m 1
a
+
k
m
(1)
(1)
2

N
m
m
1m
1m
k=0

(B.22)

where

nm =

f1n (y)f1m (y)dy

(B.23)

and
S
Fmk

X
2n tanh(2n s)Cnm Cnk
(2)

Nn

n=0

(B.24)

The coefficients aSm make the two systems consistent. Adding (B.21) and (B.22)
leads to:

(1)
S
S
X
1 10 F00
F0k
1k k0 Nk S

e
(
a
=

1) m = 0
2a
+
k
m

(1)
(1)

N
10
10
0
0
k=0
(B.25)
(S1)

(1)
S
S
X

1
F

N
F

1k km k
m0
mk
S

aSk = e10
m1
2am +
(1)
(1)
2

N
m
m
1m
1m
k=0
where km is the usual Kronecker symbol:
km =

1 k=m
0 k 6= m

By solving (S 1) the aSm (m = 0,1,2...)


(m = 0,1,2...) determined using (B.12).

(B.26)

can be evaluated and hence the bSm

The asymmetric parts of the diffracted potentials used are the ones from Cho
and Kim (1998)
A(1)
D

A(2)
D

X
ig 1 10 x
1n (x+s)
f10 (y) +
aA
f1n (y)},
= { e
ne
2
n=0

ig X A
b sinh(2n x)f2n (y).
=
n=0 n

(B.27)
(B.28)

Appendix B Analytical model of duct covered with a tensioned membrane: the


maths
B5
In a similar way to what was done for the symmetric part, the aA
m (m = 0,1,2...)
can be evaluated by solving the system

(1)
A
A
X
1 10 F00
F0k
1k k0 Nk A

a
=

1) m = 0
2a
+
e
(
k
m

(1)
(1)

N
10 0
10 0
k=0
(B.29)
(S2)

(1)
A
A
X

1 10 Fm0
Fmk 1k km Nk A

ak = e
m1
2am +
(1)
(1)
2

N
m
m
1m
1m
k=0

where

A
Fmk

X
2n coth(2n s)Cnm Cnk
(2)

Nn

n=0

(B.30)

and the remaining bA


m using

bA
m =

B.2.2

1 10 a
e
Cm0
2

aA
m Cmk

k=0
(2)

sinh(2m a)Nm

(B.31)

Radiation problem

The symmetric radiation modal potentials are expressed as


S(1)
lR

S(2)
lR

S(3)
lR

ig X S 1n (x+s)
= {
a e
f1n (y)},
n=0 ln

ig X S
i S(2)
= {
bln cosh(2n x)f2n (y) + lR (x, y)},
n=0
g

i S(3)
ig X S
cln cosh(3n x)f3n (y) + lR (x, y)},
= {
n=0
g

(B.32)
(B.33)
(B.34)

with
i cos(Sl x/s)[m1 cosh(m1 y) + sinh(m1 y)]
S(2)
lR (x, y) =
,
m1 [m1 sinh(m1 d) + cosh(m1 d)]
i cos(Sl x/s) cosh[m1 (y + h)]
S(3)
lR (x, y) =
,
m1 sinh[m1 (h d)]

(B.35)
(B.36)

and
Sl
,
s
2
.
=
g

m1 =

(B.37)
(B.38)

Appendix B Analytical model of duct covered with a tensioned membrane: the


B6
maths
In a similar way to the diffraction potential, aSlm can be determined by solving the
system
(S3)

2aSlm +

(1)
S
X
Fmk
1k km Nk

aSlk =

(1)

1m Nm

k=0

S
Xlm
(1)

1m Nm

m 0,

(B.39)

where
S
Xlm

i
=
g

lR
(s, y)f1m (y)dy.
x
S(2)

(B.40)

The bSlm are obtained using the continuity of potential on x = s as done earlier

with

bSlm =

aSlk Cmk

k=0
(2)

cosh(2m s)Nm

(B.41)

Finally using the condition of vanishing normal velocity (8.8) for region (3)
S(3)

lR
(s, y) = 0
x

on x = s, h y d

(B.42)

leads to

S(3)
i lR
(s, y).
sinh(3n s)f3n (y) =
g x

cSln 3n

n=1

(B.43)

Multiplying (B.43) by f3m (y) and integrating over [-h,-d] allows the calculations
of cSlm as
cSlm

i
(3)

g3m Nm sinh(3m s)

with
f3n =
and
Z

f3n (y)f3m (y)dy =

(3)
Nm

S(3)

cos k3n (y + h)
cos k3n (h d)

lR
(s, y)f3m (y)dy,
x

n0
(3)
Nm

m=n

0 m 6= n

(h d) m = 0
.
= 1
(h d) m =
6 0
2

(B.44)

(B.45)

(B.46)

(B.47)

Appendix B Analytical model of duct covered with a tensioned membrane: the


maths
B7
The asymmetric radiated modal potentials are given as
A(1)
lR

A(2)

lR

A(3)
lR

ig X A 1n (x+s)
a e
f1n (y)},
= {
n=0 ln

(B.48)

ig X A
i A(2)
{
bln sinh(2n x)f2n (y) + lR (x, y)},
n=0
g

ig X S
i A(3)
= {
cln sinh(3n x)f3n (y) + lR (x, y)},
n=0
g

(B.49)
(B.50)

with
i sin(A
A(2)
l x/s)[m2 cosh(m2 y) + sinh(m2 y)]
,
lR (x, y) =
m2 [m2 sinh(m2 d) + cosh(m2 d)]
i sin(A
A(3)
l x/s) cosh[m2 (y + h)]
lR (x, y) =
m2 sinh[m2 (h d)]

(B.51)
(B.52)

and
A
l
.
s

m2 =

(B.53)

In a similar way to solving the symmetric problem, the aA


lm coefficients can be
calculated by solving
(S4)

2aA
lm

(1)
A
X
Fmk
1k km Nk
(1)

1m Nm

k=0

aA
lk =

A
Xlm
(1)

1m Nm

m0

(B.54)

where
A
Xlm

i
=
g

0
d

lR
(s, y)f1m (y)dy
x
A(2)

(B.55)

and hence the bA


lm , using the continuity of potential at x = s,
bA
lm =

k=0

aA
lm Cmk
(2)

sinh(2m s)Nm

(B.56)

The zero velocity condition is again used to determine the cA


lm leading to
cA
lm

(3)

g3m Nm cosh(3m s)

d
h

lR
(s, y)f3m (y)dy.
x
A(3)

(B.57)

Appendix B Analytical model of duct covered with a tensioned membrane: the


B8
maths

B.2.3

Membrane deformation

The equation for the elastic membrane, in absence of physical damping can be
written as
Tx

d2
(x) + m 2 (x) = i[(3) (x, d) (2) (x, d)],
dx2

(B.58)

where Tx is the membrane tension, the fluid density and m the mass of the
membrane per unit length. (2) (x, y) and (3) (x, y) are the total potential in
regions (2) and (3) respectively and can be written as
(2)

(x, y) =

(2)
D (x, y)

(2)

j lR (x, y)

(B.59)

j=1

(3)

(x, y) =

(3)

j lR (x, y)

(B.60)

j=1

and the membrane displacement (x) is expressed as


(x) =

j fj (x).

(B.61)

j=1

Splitting potentials and displacement into symmetric and asymmetric modes,


(B.58) can be solved successively for both modes. Substituting

S(2)

(x, y) =

S(2)
D (x, y)

S(2)

jS lR (x, y),

(B.62)

j=1

S(3)

(x, y) =

S(3)

jS lR (x, y),

(B.63)

j=1

and
S (x) =

jS fjS (x)

(B.64)

j=1

into (B.58) leads to

X
j=1

jS [T

d2 fjS
m 2 fjS (x) pSjR (x)] = pSD (x).
2
dx

(B.65)

Appendix B Analytical model of duct covered with a tensioned membrane: the


maths
B9
where
S(3)

S(2)

pSjR (x) = i[jR (x, d) jR (x, d)],


pSD (x) =

S(3)
i[D (x, d)

S(2)
D (x, d)].

(B.66)
(B.67)

Multiplying (B.65) by fiS and integrating over [-s,s] yields

X
j=1

[KijS 2(MilS + aSil ) ibSij ]jS = FiS

i = 1, 2, 3...

(B.68)

with
KijS
aSij
FiS

Z s
d2 fj
S
mfj (x)fi (x)dx
(B.69)
=
T 2 (x)fi (x)dx, Mij =
dx
s
s
Z
Z s
1 s S
1
S
S
S
p (x)fi (x)dx) (B.70)
p (x)fi (x)dx), bij = =(
= <( 2
s jR
s jR
Z s
=
pSD (x)fi (x)dx.
(B.71)
Z

Appendix C
Pictures of the experimental
set-ups

Figure C.1: The duct, covered with the rubber 1mm membrane, in the wave
basin.

C1

C2

Appendix C Pictures of the experimental set-ups

Figure C.2: Bow of the duct, with the connection for the water hose, in the
wave basin

Figure C.3: Stern of the duct, with the connection to the bent tube, in the
wave basin. Heavy masses are visible and were used to hold the bottom plate
still.

Appendix C Pictures of the experimental set-ups

Figure C.4: Power take-off system.

Figure C.5: Ramp and bow of the duct, in the narrow flume.

C3

C4

Appendix C Pictures of the experimental set-ups

Figure C.6: The duct on the artificial seabed, covered with the 1mm membrane, in the narrow flume.

Appendix D
Determination of rubber
properties
D.1

Introduction

A stripe of rubber of length L0 = 0.4m, width W0 = 0.2m and thickness at rest v0


(1mm or 0.2mm) was clamped on one its top side to a fixed item while the other
side was fashioned to receive successive loads, as shown in Figure D.1.

L0

W0

h0

Figure D.1: Sketch of the apparatus for testing rubber properties.

D1

D2

Appendix D Determination of rubber properties

(a)

(b)

(c)

Figure D.2: Three different loads applied to the strip of rubber; (a) F = 0N,
(b) F = 78.86N, (c) F = 98.48N.

This apparatus was aimed for determining Youngs modulus of the rubber and its
damping properties. Notations used in this appendix are specific to it.

D.2

Youngs modulus

Measurement of the Youngs modulus was carried by applying successive loads on


the stripe of rubber. Time interval between each new load was two minutes so
the load had to time to be completely effective. At each step, the new length L
was measured manually and electronically by using a laser sensor recording the
displacement of the bottom frame. The new thickness v (averaged over three positions lengthwise) and the new width W (averaged over three positions lengthwise)
were also measured manually. A view of the apparatus with the rubber strip subject to three different loads is shown in Figure D.2.
The stress in the rubber F/A where A is the cross-section of the strip (A = W v)
was then measured together with the elongation and the stress-strain relation
plotted. The Youngs modulus was calculated as the slope of the best fit passing
trough the origin as shown in Figure D.3.

Appendix D Determination of rubber properties

Figure D.3: Stress-strain relationship for successive loadings with two minutes
intervals.

Figure D.4: Stress-strain relationship for successive loadings with thirty seconds intervals.

D3

D4

Appendix D Determination of rubber properties

The same process was repeated but with time intervals between each loads being
thirty seconds. Result is shown in Figure D.4 where a slightly higher value of the
Youngs modulus was obtained as for the same stress, the rubber did not have
time to stretch completely, yielding lower values of elongation. However, the two
obtained values for the Youngs modulus from this test were close enough not to
make any difference in the chosen value.

D.3

Incompressibility of rubber

One assumption used while developing the load-deflection model in (5.2) was that
rubber is incompressible. This was checked in the same time than measuring
the Youngs modulus. Figure D.5 shows the measured volume V of the strip
(V = W vL) for the different stresses applied. This total volume is seen to be
constant, hence justifying the incompressibility of the rubber used in this work.

Figure D.5: Volume of the rubber strip for the different stresses applied.

Appendix D Determination of rubber properties

D.4

D5

Dynamic damping

The coefficient used is related to the energy dissipated in the mechanism of


deformation of the rubber. It can be calculated, with the notations from Figure
D.1 as
cL0
=
(D.1)
W0 v0 E
where c is a decay coefficient.
Consider a oscillating mass m attached to the rubber strip. The amplitude of
motion d is recorded as shown in Figure D.6 from which several following peaks
can be determined as d1 > d2 > ... > d9 . This decay in amplitude can be characterized by a coefficient such as
d1+n
= exp(n) n = 1, 2...8
d1

(D.2)

This coefficient is then calculated as the slope of the plot (n,ln(d1+n /d1 )). The
coefficient c can then be calculated as
c = 2mf
yielding the calculation of from (D.1)

Figure D.6: Damped free oscillations of a rubber strip.

(D.3)

Potrebbero piacerti anche