Sie sind auf Seite 1von 99

RESEARCH INTO THE FURTHER

DEVELOPMENT OF THE LIMPET


SHORELINE WAVE ENERGY PLANT
ETSU V/06/00183/REP
URN 02/1487
Contractor
Wavegen
First published 2002
Crown copyright 2002
The work described in this report was
carried out under contract as part of
the DTI Sustainable Energy
Programmes. The views and
judgements expressed in this report are
those of the contractor and do not
necessarily reflect those of the DTI.
Further Renewable Energy information from the Sustainable Energy Programme, and
copies of publications can be obtained from:
Renewable Energy Helpline
Tel: (+44 ) 01235 432450
Email: NRE-enquiries@aeat.co.uk
i
Executive Summary
This is the final report of a 24 month research and development project into the
LIMPET Shoreline OWC wave energy device. The project has focused on a range of
technical issues, some of which are particular to the LIMPET design and others that
have general significance for Shoreline OWCs.
The objectives of the project were as follows:
Task 1 Derive and validate new sea state spectra.
Task 2 New Model tests of Current LIMPET
Task 3 Standardise Evaluation Procedure
Task 4 Pressure Relief Model tests
Task 5 Visual Investigation of Collector Flow Characteristics
Task 6 Investigate Model Loads
Task 7 Conceptual Evaluation of New Device
Task 8 Installation of Power Dumping Equipment.
Task 9 Develop Additional Control Strategies
Task 10 Installation of the Blow Off Valve
Task 11 Structural Load Measurement at the Full Scale
Task 12 Project Management and Reporting
Task 13 Fitment of Acoustic Attenuation
Task 14 Removal of Rock from Gully
Task 15 Fitment of Variable Orifice to the Second Turbine Outlet
The major results and conclusions were as follows:
Difficulties with the pressure transducers attached to the LIMPET limited the
amount of wave data it was possible to collect at the site. This made it impossible
to derive an annual wave climate directly for the site. However, data from a
waverider buoy (further out to sea) were available to produce an annual set of sea-
states, which could be referred to the site using a derived transfer function. While
a better set of sea-states could be produced from other data sets, sea-states
produced in the wave tank have been shown to be representative of the typical
sea-states at the LIMPET site.
Scale model tests have confirmed that water depth has a critical effect on the
performance of a shoreline OWC. They have also shown that the benefits of the
harbour wall effect previously reported for OWCs in deep and intermediate
waters do not occur in shallow water and that the parallel gully is detrimental to
device performance.
A series of standard tests for shoreline wave energy converters has been
developed together with a standard set of test spectra. These tests examine the
device performance with respect to energy capture in productive seas and
structural loads in extreme seas.
Physical and mathematical models have been used to evaluate the performance of
pressure relief valves in OWC collectors. Whilst previous studies of idealised
valves have shown improvements in performance, the present studies, which
include a realistic simulation of valve characteristics, have shown that in general
the application of a blow off valve reduces overall performance.
ii
The use of freeze frame digital video has identified significant non-linear
characteristics in the water motions associated with the LIMPET OWC. The most
significant of these are described as front wall slosh, front wall wash down
and water column slosh. The results have again emphasised the influence of
water depth on device performance and the need to optimise device design for a
particular water depth.
Scale model measurements of structural loads on the LIMPET front wall have
shown that the two most challenging load conditions applied to the design,
external wave slam and peak internal pressure are not independent but are
generally in opposition. If the peak net load is used as the basis of design for the
front wall rather than the peak of the individual loads then the structural
requirement of the front wall is significantly reduced.
The demonstration of extreme loads very much less than those assumed in the
design of LIMPET has permitted a review of the design philosophy and led to a
new concept of collector construction which has reduced the concrete volumes per
unit capture width to approximately one third of that used in LIMPET.
The baseline control strategy has been demonstrated as reliable and secure.
Operational experience has demonstrated the need to change mean turbine speed
with available input power in order to minimise parasitic losses, maximise turbine
efficiency and limit the incidence of stall at times of high wave activity.
Modifications have been made to the control software with the purpose of
achieving these objectives. Longer term monitoring is required to assess the
success of the changes.
Pressure transducers fitted to the inner rear wall of the collector have shown
minimal internal wave loads showing that the internal design of the collector has
been successful in eliminating the incidence of internal wave slam.
Acoustic attenuation fitted to the exit of the turbine system has proved extremely
effective in preventing any noise disturbance from the LIMPET plant. The
particular design adopted has however caused some flow disturbance and has
affected turbine efficiency.
After the first winter of operation it was found that debris from the original
construction remained in the gully. This was successfully removed in the summer
of 2001 using a long reach excavator.
A variable valve was fitted to the second turbine outlet. This enabled the pressure
flow characteristic of the collector to be measured independent of the turbine
operation.
The work tasks undertaken have helped in the understanding of the wave environment
and in the construction and operation of wave energy plant. The experience of
building and operating LIMPET has been invaluable not only in respect of furthering
the development of more shoreline generation but also in respect of developments
offshore. In particular, experience with the LIMPET instrumentation and control
systems together with grid integration issues is directly applicable to the current round
of offshore developments.
LIMPET continues in grid connected operation both as a power generator and as an
equipment test bed.
iii
Contents
1 Introduction..........................................................................................................1
1.1 Project Objectives .......................................................................................... 2
1.2 Parallel Monitoring Project............................................................................ 2
2 Task 1: Derive and Validate New Sea States.....................................................4
2.1 Introduction.................................................................................................... 4
2.2 Review of current sea-state parameters ......................................................... 4
2.2.1 Wave height parameters......................................................................... 4
2.2.2 Wave period parameters ........................................................................ 5
2.2.3 Wave group parameters ......................................................................... 5
2.2.4 Wave power parameters......................................................................... 6
2.3 Wave data sets................................................................................................ 7
2.3.1 Wavetank data........................................................................................ 7
2.3.2 LIMPET site data................................................................................... 7
2.3.3 Waverider buoy data.............................................................................. 8
2.4 Comparison of data sets ................................................................................. 8
2.4.1 Comparison of wavetank data with and without the LIMPET model
and coastline........................................................................................................... 8
2.4.2 Comparison of wavetank data for waves at LIMPET model and 20-
metre contour ......................................................................................................... 9
2.4.3 Comparison of LIMPET site data at different tide levels .................... 10
2.4.4 Comparison of LIMPET site data and wavetank data ......................... 11
2.4.5 Comparison of LIMPET site data and waverider buoy data................ 12
2.5 Specification of sea-states............................................................................ 13
2.6 Discussion.................................................................................................... 14
2.7 References.................................................................................................... 15
3 Task 2 New Model tests of Current LIMPET. .............................................17
3.1 Seabed profiles............................................................................................. 17
3.1.1 Profile 1................................................................................................ 17
3.1.2 Profile 2................................................................................................ 17
3.1.3 Profile 3................................................................................................ 18
3.2 Wave test series............................................................................................ 18
3.3 Wave tank and model set-up........................................................................ 18
3.4 Instrumentation and procedure .................................................................... 19
3.5 Estimation of incident wave power.............................................................. 19
3.6 Power and capture factor ............................................................................. 19
3.7 Incident wave behaviour .............................................................................. 24
3.8 Water column movement ............................................................................. 25
3.9 Influence of the chamber air volume ........................................................... 27
3.10 Discussion & Conclusions ........................................................................... 28
4 Task 3. Standardise Evaluation Procedure .....................................................29
4.1 Modelling and analysis of shoreline and nearshore wave energy converters
29
4.1.1 Ocean wave analysis techniques.......................................................... 29
4.1.2 Modelling ocean waves for shoreline and nearshore wave energy
converters............................................................................................................. 33
4.2 Issues in the wavetank testing of shoreline and nearshore wave energy
converters................................................................................................................. 36
4.2.1 Model scaling....................................................................................... 36
iv
4.2.2 Selection of wavetank facilities ........................................................... 38
4.2.3 Wavetank calibration and testing......................................................... 38
4.2.4 Wave-to-wire modelling ...................................................................... 40
4.3 Specification of a standardised test procedure for shoreline and nearshore
wave energy converters............................................................................................ 40
4.3.1 Baseline tests for shoreline and nearshore wave energy converters .... 41
4.3.2 Additional tests for shoreline and nearshore wave energy converters. 42
4.4 Documentation of shoreline and nearshore wave energy converters testing
44
4.4.1 Experimental set-up ............................................................................. 44
4.4.2 Results section ..................................................................................... 44
5 Task 4. Pressure Relief Model tests..................................................................45
5.1 Analysis of a LIMPET blow-off valve ........................................................ 45
5.1.1 Introduction.......................................................................................... 45
5.1.2 Wavetank simulation of the LIMPET blow-off valve ......................... 46
5.1.3 Results of the wavetank simulation of the LIMPET blow-off valve... 47
5.1.4 Discussion of Wavetank results ........................................................... 49
5.1.5 Numerical simulation of the LIMPET blow-off valve ........................ 51
5.2 Results of the numerical simulation............................................................. 52
5.3 Conclusions.................................................................................................. 53
5.4 References.................................................................................................... 54
5.5 Appendix 5A Estimations of coefficients for numerical simulation ........ 54
5.5.1 K water plane stiffness...................................................................... 54
5.5.2 f wave force....................................................................................... 54
5.5.3 B hydrodynamic damping................................................................. 54
5.5.4 M water column mass ....................................................................... 55
5.5.5 - air spring stiffness ........................................................................... 55
5.5.6 - effective damping coefficient of the turbine and poppet valve ..... 55
6 Task 5 Visual Investigation of Collector Flow Characteristics .....................57
6.1 Introduction.................................................................................................. 57
6.1.1 Experimental set-up ............................................................................. 58
6.2 Identification and discussion of non-linear flow characteristics.................. 60
6.2.1 Characteristic 1: Front wall swash....................................................... 61
6.2.2 Characteristic 2: Front wall down-wash .............................................. 61
6.2.3 Characteristic 3: Water column slosh .................................................. 62
6.3 Conclusions and Implications for Shoreline OWCs .................................... 68
6.4 References.................................................................................................... 69
7 Task 6 Investigate Model Loads .......................................................................70
7.1 Introduction.................................................................................................. 70
7.2 Model testing ............................................................................................... 70
7.3 Internal Wave Slam...................................................................................... 74
8 Task 7 Conceptual Evaluation of New Device.................................................75
8.1 Development of the LIMPET Cross Section ............................................... 75
8.2 Capture Performance of 10m wide Cliff Mounted Device.......................... 77
8.3 Discussion and Conclusions ........................................................................ 77
9 Task 8. Installation of Power Dumping Equipment. ......................................78
10 Task 9. Develop Additional Control Strategies...............................................78
10.1 Introduction.................................................................................................. 78
10.2 Baseline Control Strategy ............................................................................ 78
10.2.1 Pre-start Check..................................................................................... 78
v
10.2.2 Start up................................................................................................. 79
10.2.3 Power production................................................................................. 79
10.2.4 Generator Speed Control...................................................................... 80
10.2.5 Valve Position Control......................................................................... 80
10.3 Limitation of Baseline Controls................................................................... 80
10.4 Revised Control Strategy............................................................................. 81
10.5 Stall Control ................................................................................................. 82
10.6 Discussion and Conclusion.......................................................................... 83
11 Task 10 Installation of the Blow Off Valve...................................................83
12 Task 11- Structural Load Measurement at the Full Scale .............................83
13 Task 13 Fitment of Acoustic Attenuation.....................................................85
14 Task 14 Removal of Rock from the Gully ....................................................86
15 Task 15 Fitment of Variable Orifice to the Second Turbine Outlet..........88
16 Results and Conclusions....................................................................................91
17 Future Developments.........................................................................................92
1
1 Introduction
Figure 1
The LIMPET wave energy converter was built with EU support over the period 1998-2000 by a
consortium led by Wavegen and the Queens University of Belfast. QUB were responsible for the
site selection and development of the collector form whilst Wavegen are the plant owners and
operators, and were responsible for the development, construction and installation of the turbo-
generation equipment. The plant was built to fulfil a number of functions including:
The demonstration of shoreline wave energy as an effective supply mechanism to the UK
national grid.
The supply of power to the grid under a contract under the Scottish Renewables Order.
The testing of previously untried construction techniques.
The operation of a contra-rotating Wells turbine previously untried at the full scale.
The gaining of experience in plant construction and operation.
The provision of a research facility to test new systems and to permit a calibration of model
tests.
Whilst the original EU project was successful in these objectives it became clear as the construction
of the plant approached completion that there were activities outwith the scope of that project
which were of generic value to the wave energy community and which merited support from the
UK national government via the DTI. This recognition and the desire to maximise the benefit of
LIMPET led to the establishment of this project, which in part ran concurrently with the EU project.
The project concentrated on aspects of the construction and operation of shoreline wave energy
plant which were identified during the EU project as requiring further study, and which it was
anticipated could have application on LIMPET, derivative shoreline plant or in many cases on
2
future offshore wave energy generators. Items of generic interest covered by the work programme
include:
Standardisation of Evaluation procedures
Investigation of Structural Loads
Development and evaluation of Control Strategies
Evaluation of power capping via pressure relief valves.
The work was coordinated by Wavegen under the supervision of Dr.T.V.Heath. The plant has been
operated and maintained under the control of Mr.H.Ellen.
The majority of the research studies were performed by the wave energy team at the Queens
University of Belfast under the leadership of Professor T.J.T Whittaker. The prime responsibility
for data collection and organisation was with Mr.C.Boake, and the majority of other analysis and
initial reporting has been performed by Dr. M Folley.
The support of the renewable energy teams at the DTI and ETSU is acknowledged and the project
team would like to express its thanks for their continued interest throughout the project.
1.1 Project Objectives
The project originally consisted of twelve tasks as follows:
Task 1 Derive and Validate New Sea States
Task 2 New Model tests of Current LIMPET
Task 3 Standardise Evaluation Procedure
Task 4 Pressure Relief Model tests
Task 5 Visual Investigation of Collector Flow Characteristics
Task 6 Investigate Model Loads
Task 7 Conceptual Evaluation of New Device
Task 8 Installation of Power Dumping Equipment.
Task 9 Develop Additional Control Strategies
Task 10 Installation of the Blow Off Valve
Task 11 Structural Load Measurement at the Full Scale
Task 12 Project Management and Reporting
In respect of information gathered and analyses performed during the execution of the project,
tasks 8 and 10 were deleted from the scope of work and three new tasks added. These were:
Task 13 Fitment of Acoustic Attenuation
Task 14 Removal of Rock from Gully
Task 15 Fitment of Variable Orifice to the Second Turbine Outlet
1.2 Parallel Monitoring Project
In addition to this project, the DTI Renewable Energy Programme supported a parallel project,
(V/06/00180) to monitor the latter stages of construction, turbo-generation equipment installation,
3
commissioning and operation of the LIMPET. These activities are summarised in a separate report
entitled ISLAY LIMPET PROJECT MONITORING FINAL REPORT.
4
2 Task 1: Derive and Validate New Sea States
2.1 Introduction
The original model tests on LIMPET were based on the bathymetry indicated by an existing survey
and a set of wave spectra derived for the site of the QUB 75kW prototype wave energy converter. A
major objective of the LIMPET project has been and continues to be to correlate site measured data
with tank studies.
The accuracy of wave tank modelling depends amongst other things on the accuracy of the sea-
states used for testing. It has been suggested from anecdotal reports that the waves generated in the
wave tank do not accurately represent the occurrence of groups of waves in typical seas. Sea-states
generated in wave tanks for the analysis of shallow water wave energy converters have typically
been defined using water depth, wave height and incident wave power, using a Bretschneider
spectrum.
The validity of this representation of sea-states has been examined by comparison of data sets
obtained in the wave tank and from the LIMPET site. In particular, the wave spectra, interpreted in
shallow water as wave groupiness, have been investigated and standard parameters compared
between wave tank and LIMPET site data. In addition, the calculations of the wave height, wave
period, and incident wave power parameters have been reviewed to determine their suitability for
shallow water sites.
Following the identification of appropriate representations of a sea-state, a new set of productivity
sea-states has been produced for use with shallow water wave energy converters. This has been
achieved by calibrating waverider buoy data, which provides an extensive data set, with LIMPET
site data and wave tank data.
2.2 Review of current sea-state parameters
2.2.1 Wave height parameters
The significant wave height, H
s
, is the most common parameter used in the definition of sea-states.
Traditionally this has been defined as the average height of the highest one-third waves, H
1/3
, where
the wave height is measured from trough to peak for each wave defined by the zero up-crossing
point [1].

=
=
3 /
1 3
3 / 1
1
n
i
i
n
H H
In this relationship n is the number of individual wave heights H
i
in a record ranked highest to
lowest. This definition was adopted to correlate with wave height observation reports that had been
used previously. There is nothing implicitly significant about the highest one-third waves and other
measures such as the average of the highest 1/10 of waves, H
1/10
, have been proposed.
5
Modern analyses have generally used an alternative definition of the significant wave height, H
mo
,
based on the root-mean-square of the surface profile,
rms.
rms mo
H = 4
In deep water with a narrow band spectrum, these two measures of significant wave height are
numerically very similar. However, in shallow water or broad band spectra the two methods can
produce different values for the same wave record.
Other parameters such as the maximum wave height in a wave record, H
max
, or the average wave
height, H
mean
, are sometimes used in the analysis of waves, though the former clearly depends on
the record length and the latter is no less arbitrary than H
1/3
. For the analysis of wave energy
converters, it is proposed that H
mo
is used as the single parameter for wave height since this is
directly related to the energy in the wave.
2.2.2 Wave period parameters
The most commonly used wave period parameter is the zero-crossing period, T
z
, equal to the
average time between up-crossings of the mean water level.

=
=
n
i
i z
T
n
T
1
1
Effective use of this parameter requires some smoothing of the wave profile to eliminate the effect
of ripples on the water surface.
A common parameter used in the analysis of wave energy converters is the energy period, T
e
, which
is equal to the period of a monochromatic wave with the same average power as the wave record
and with the same significant wave height, H
mo,
[2]. In deep water, this can be calculated using
spectral moments, however no such simple method exists for shallow water. The wave spectrum
can also be used to determine the peak wave period, T
p
, equal to the wave frequency with the
highest spectral energy density.
The use of Fourier analysis in shallow water is suspect and so calculating either the energy period
or peak wave period is problematic. Thus, the zero crossing period is the most suitable parameter
for wave period in shallow water. However, both T
e
and T
p
have been used previously to define a
sea-state in shallow water and thus care must be taken to ensure that when comparing sea-states that
the same wave period parameters are used in both data sets.
2.2.3 Wave group parameters
Though there are a number of different methods of measuring the degree of grouping in waves, only
two are used extensively: the average run length, J
1
, and the groupiness factor, GF, [3, 4].
The average run length, J
1
, is the average number of waves in which the wave heights are above a
particular threshold, such as the significant wave height or mean wave height. There is no standard
value for the threshold used to determine average run length and so particular care must be taken
when comparing values of this parameter for different sea-states.
6
The groupiness factor, GF, is a measure of the variance in wave energy. The Smoothed
Instantaneous Wave Energy History (SIWEH), which employs a Bartlett filter to smooth the wave
energy, is used to determine the variance in wave energy.


+ = d Q t
T
t E
p
) ( ) (
1
) (
2
where T
p
is the peak wave period and Q() is the Bartlett filter given by
p
p p
T
T T
Q

<
=

0
/ 1
{ ) (
The groupiness factor is then given by

=
dt t E
dt t E
GF
) (
) (
2
Comparing these two parameters indicates that they do not characterise the wave grouping in the
same fashion. A high groupiness factor implies the existence of relatively large packets of energy,
but does not say anything about their distribution in time. However, a high average run length
implies the existence of long series of relatively large waves, but does not say anything about the
amount of energy in the run. Though both parameters are easily calculated they both have a high
level of variability that reduces only slowly with an increase in record length. To provide a
reasonable estimate the record length for a data set from the North Atlantic typically needs to be
greater than 160 minutes. In shallow water, records of this length are problematic since changes in
tide level during one period mean that the wave record can no longer be considered stationary.
Although in narrow band spectra there is some correlation between average run length and
groupiness factor they do not generally have a high degree of correlation and should be considered
separate and independent. Thus, both parameters are used for defining the sea-state.
2.2.4 Wave power parameters
Although the estimation of incident wave power is critical to the design of shoreline wave energy
converters, no reliable method exists. The currently accepted method is to use a Fourier analysis to
produce a pseudo wave spectrum. Defining the incident wave as the sum of waves from the
spectrum, the power in each wave is calculated using its height and wave number. The total incident
power, P
i
, is then equal to the sum of the power from the individual waves.

+ =


2 2
) (
)
) 2 sinh(
2
1 )( tanh(
4
a
kd
kd
kd
g
P
i
where ) tanh(
2
kd
g
k

=
This method of calculating incident wave power takes no account of the non-sinusoidal nature of
shallow water waves, or the effects of wave breaking. However in the absence of an alternative
method it is often used.
7
2.3 Wave data sets
For the LIMPET project, three sets of data were available for analysis and comparison; wave tank
data, LIMPET site data and waverider buoy data. The first two data sets consist of time series
records of wave conditions, whilst the waverider buoy data set contains only summary data.
2.3.1 Wave tank data
In the calibration of wave tanks, it is standard procedure to calibrate the waves with the model and
coastline removed. This eliminates the influence of reflected waves and isolates the incoming wave.
Fifty-three sea-states, with a range of water depths, significant wave heights and incident wave
powers were used for the assessment of LIMPET. Time series data of the water surface elevation at
the position of the mouth of the LIMPET gully, taken during the wave tank calibration, was
available for analysis. In addition, a smaller set of data, which includes data for when the model and
coastline were replaced in the wave tank, was also available.
The wave surface elevation was measured using a twin-wire wave probe linked to a data-logging
computer collecting data at 32 Hz. The model scale was 1:40, so at full scale this is equivalent to a
sampling rate of 5.06 Hz.
2.3.2 LIMPET site data
Data corresponding to the wave climate incident on LIMPET was collected using seabed pressure
transducers sited approximately 44 and 66 metres from the front wall of LIMPET, with each
transducer sited in approximately six metres of water. Data was collected for approximately 13
minutes each hour, providing a detailed account of the wave climate. Hourly recording of data, as
opposed to the standard three hourly recording of wave data, was used because of the need to
determine the effect that the tide has on the incident wave climate. A sampling frequency of 5 Hz
was used, enabling the collected data to be correlated with data collected from the operation of
LIMPET. This frequency is higher than the standard sampling frequency for wave data of 4 Hz.
The seabed transducers were deployed on the 20
th
November 2001, and data started to be collected
from 5pm. The signal from the transducer at 66 metres was lost following a storm one week later at
9pm on the 27
th
November 2001. During this time, it was possible to collect 171 recordings. The
remaining transducer signal was lost during a storm at 11am on the 18
th
January 2002.
Unfortunately, power cuts on Islay during December meant that during this second period only an
additional 308 recordings were obtained.
The signal from the seabed pressure transducer needed to be transformed to obtain the surface
profile where it is sited. The transformation was achieved using the quasi-period for each half wave,
where the half-period is defined as the time between zero-crossing points. The quasi-period was
converted into a wave number using the water depth at the site of the transducer and then the signal
amplitude was increased to account for the reduction in wave-induced pressure fluctuations with
water depth. It has been found that using the quasi-period, linear theory provides a good estimation
of the attenuation factor. Thus:
p e
kh
=
8
Where is the surface displacement, k is the wave number derived from ) tanh(
2
kh k
g
=

, h is the
water depth and p is the variation from mean in recorded pressure in metres of water.
2.3.3 Waverider buoy data
The nearest Met Office waverider buoy is buoy 62106, sited at 057.00 N, 009.90 W. This is
approximately 250 km from the LIMPET site. This buoy gives a range of meteorological data;
however, the only wave data it provides is the significant wave height and wave period. Two years
of data from 1
st
Feb 2000 31
st
Jan 2002 was available for the project. The data was provided by
the Met Office without any filtering and thus is contaminated by false readings. Data with reported
significant wave heights greater than 20m, and data with reported zero crossing periods greater than
20 seconds was removed from the data set when used in this analysis.
2.4 Comparison of data sets
Five separate comparisons of the data sets have been performed. These enable the validity of the
wave tank sea-states to be assessed and production of the basic elements of a representative set of
sea-states.
2.4.1 Comparison of wave tank data with and without the LIMPET model and coastline
Figure 1 compares the calculated wave parameters for the wave record with and without the
model/coastline. The first record of each couple, e.g. Te07Pi10L175.txt, is without the
model/coastline and the second record, e.g. D5Te07Pi10L175-2.txt, is with the model/coastline.
Record name H
mo
H
s
H
max
H
mean
T
e
T
p
T
z
GF J
i
P
i
Te07Pi10L175.tx
t
1.55 1.4
5
2.40 1.05 7.27 8.87 6.21 0.6
8
1.6
8
9.08
D5Te07Pi10L17
5-2.txt
1.61 1.4
6
2.66 1.05 7.20 8.68 6.07 0.6
4
1.8
1
9.81
Te07Pi20L175.tx
t
2.15 2.0
1
3.32 1.45 7.37 9.27 6.25 0.6
6
1.6
5
17.5
8
D5Te07Pi20L17
5-2.txt
2.22 2.0
1
3.67 1.47 7.32 8.68 6.02 0.6
5
1.6
9
18.6
4
Te10Pi10L175.tx
t
1.41 1.2
9
1.68 0.87 10.5
4
12.3
6
8.27 0.6
9
1.7
2
8.78
D5Te10Pi10L17
5-2.txt
1.43 1.3
2
1.77 0.88 10.2
2
11.0
3
7.31 0.6
9
1.6
6
8.93
Te10Pi20L175.tx
t
1.94 1.7
8
2.36 1.19 10.7
5
11.6
6
8.19 0.6
9
1.7
4
16.5
4
D5Te10Pi20L17
5-2.txt
1.98 1.8
2
2.47 1.19 10.2
7
11.6
6
7.59 0.6
8
1.7
0
17.1
2
Te10Pi40L175.tx
t
2.71 2.5
2
3.38 1.69 10.5
0
12.0
0
8.62 0.6
6
1.7
7
32.3
4
D5Te10Pi40L17 2.74 2.5 3.52 1.68 10.2 10.7 7.73 0.6 1.5 32.7
9
5-2.txt 5 6 4 6 9 0
Te13Pi40L175.tx
t
2.48 2.2
9
2.90 1.63 13.5
5
16.3
2
11.5
4
0.5
8
1.9
4
28.9
3
D5Te13Pi40L17
5-2.txt
2.52 2.2
0
2.97 1.50 13.4
7
14.0
7
9.42 0.6
1
1.7
9
29.3
6
Te10Pi20L187.tx
t
2.00 1.8
7
2.45 1.29 10.5
9
11.0
3
8.45 0.6
8
1.5
2
18.0
2
D5Te10Pi20L18
7-2.txt
2.04 1.8
8
2.50 1.26 10.3
1
14.0
7
7.88 0.6
5
1.6
3
18.6
6
Te10Pi60L187.tx
t
3.35 3.2
0
4.22 2.15 10.5
9
12.3
6
8.71 0.6
6
1.6
0
50.5
5
D5Te10Pi60L18
7-2.txt
3.39 3.1
9
4.25 2.09 10.3
3
12.7
5
8.11 0.6
4
1.6
7
51.3
3
Te13Pi30L187.tx
t
3.12 2.8
7
3.72 2.05 13.7
0
13.6
0
11.0
7
0.6
2
1.8
7
46.8
7
D5Te13Pi60L18
7-2.txt
3.16 2.9
0
3.70 2.03 13.2
0
13.6
0
10.1
1
0.5
9
1.9
0
47.5
1
Figure 2 Comparison of wave parameters with and without model/coastline.
The data shows that in general the addition of the model/coastline has a minimal effect on the wave
parameters. On average the incident power increases by 3%, the significant wave height, H
mo
,
increases by 2%, the maximum wave height increases by 4% and the zero crossing period decreases
by 8%. All the other parameters remain approximately constant. These results indicate that there is
minimal coherence between the incident and reflected waves at the point of measurement, i.e. the
reflected wave is as likely to destructively interfere with the incident wave as constructively
interfere.
It is significant that seabed pressure transducers deployed at the LIMPET site to measure wave
climate were positioned at an equivalent location to the wave probe used to derive the wave
parameters in these wave tank tests. Thus, it is reasonable to assume that the wave parameters
derived from data obtained from these transducers are approximately equal to the incident wave
parameters.
2.4.2 Comparison of wave tank data for waves at LIMPET model and 20-metre contour
This comparison was required to translate the wave climate measured at the LIMPET site to the 20-
metre depth contour that is used to define the sea-states. Unsurprisingly, the wave period and wave
group parameters are not significantly affected by the change in water depth. Wave height
parameters are affected due to wave breaking between the 20-metre depth contour and the LIMPET
site. However, if only non-breaking waves are considered then the significant wave height varies
little between the two measuring positions. Figure 3 shows the correlation between the deep and
shallow water significant wave heights.
This surprising result is due to the shoaling coefficient being initially less than 1.0 before increasing
as the water depth decreases. The effect of shoaling on each wave will be complex and unique,
however it would appear that at the water depths and wave periods typical for the sea around
LIMPET these effects do not cause the significant wave height to change.
10
0
0.5
1
1.5
2
2.5
3
0 0.5 1 1.5 2 2.5 3
Shallow water significant wave height (m)
D
e
e
p

w
a
t
e
r

s
i
g
n
i
f
i
c
a
n
t

w
a
v
e

h
e
i
g
h
t

(
m
)
Figure 3 Comparison of significant wave heights, H
mo
, at the 20-metre depth contour and
LIMPET site
2.4.3 Comparison of LIMPET site data at different tide levels
Figure 3 shows the variation of significant wave height, H
mo
, and tide level for the period 24
th
November 2001 to 30
th
November 2001. The change in significant wave height with tide can be
clearly seen as a modulation at the same frequency as the tide level superimposed on a changing
significant wave height due to prevailing wave conditions. However, the tide level and modulation
of significant wave height are not in phase, with the peak significant wave height lagging the peak
tide level by approximately 2.5 hours.
-2
-1
0
1
2
3
4
5
6
S
i
g
n
i
f
i
c
a
n
t

w
a
v
e

h
e
i
g
h
t

/

T
i
d
e

l
e
v
e
l

(
m
)
Tide level (m)
Significant wave height (m)
24 Nov '01 25 Nov '01 26 Nov '01 27 Nov '01 28 Nov '01 29 Nov '01 30 Nov '01
Figure 4 Variation of significant wave height with tide
11
This time lag implies there are additional factors associated with the tide, other that water depth,
influencing the significant wave height. Observation of the site indicates a large amount of wave
breaking occurring in a region due south of the Frenchmans Rocks (see Figure 5). This is
associated with a strong tidal current that runs north south off the Frenchmans Rocks due to the
filling and emptying of the Irish Sea. It is the combination of tide level and energy dissipation due
to this wave breaking that causes the change in significant wave height.
LIMPET
Frenchmans
Rocks
Region of
wave breaking
Figure 5 Frenchmans Rocks and wave breaking
2.4.4 Comparison of LIMPET site data and wave tank data
Wave tank simulations of LIMPET define the sea-state using three parameters; the water depth, the
incident power and the significant wave height. These are produced using the Bretschneider
spectrum and thus defining the incident wave power and significant wave height defines the wave
period. In order to compare waverider data with LIMPET site data these parameters must be
appropriately matched. Thus, the comparison between LIMPET site and wave tank data essentially
consists of examination of the wave group parameters; groupiness factor and average run length.
This is the time-based equivalent of comparing the frequency-based wave spectra and is preferred
due to the problems already identified regarding the use of wave spectra in shallow water. Figure 6
shows the average wave group parameters for all of the data sets. This shows a high degree of
similarity between the wave group parameters for the LIMPET site data and the wave tank data.
This implies that on average the wave tank waves have similar wave groups to the LIMPET site.
Wave tank data LIMPET site data
Groupiness factor 0.55 0.53
Average run length 1.86 1.85
Figure 6
Two types of analysis were performed on the data to examine it in more detail. A correlation
analysis was used to identify whether either wave group parameter is correlated to another wave
parameter. In addition, an analysis of variance was used to determine the variation in the wave
group parameters.
12
The average run length is poorly correlated with all other wave parameters for both the wave tank
and LIMPET site data. Comparing the variance of average run length between the wave tank data
and LIMPET site data indicates that the LIMPET site data has a slightly higher variance than the
wave tank data, though only 0.14 compared to 0.11.
The groupiness factor is correlated to the significant wave height, with a low groupiness factor
associated with a high significant wave height. This reduction in the groupiness factor has
previously been linked to wave breaking and saturation of the waves. A similar effect is evident in
both the wave tank and LIMPET site data, indicating that the effect of significant wave height is
similar for both data sets. The variance in the groupiness factor is again slightly higher for the
LIMPET site data compared to the wave tank data.
This comparison indicates that the wave trains used for testing in the wave tank have on average
similar wave group parameters to those at LIMPET. Moreover, the change in groupiness factor due
to significant wave height is similar for both data sets. The LIMPET site data shows a greater
variation in the group parameters, which is associated with a broader range of spectra, however in
general the differences are relatively small.
2.4.5 Comparison of LIMPET site data and waverider buoy data
Because the seabed pressure transducers were only operating for a short time, it was not possible to
use the data collected from them directly to produce an annual wave climate for the site. However,
correlating this data with waverider buoy data has enabled an estimation of the annual wave climate
to be produced.
Figure 7 shows the comparison between wave heights recorded for the period 24
th
Nov 2001 to 30
th
Nov 2001 by the waverider buoy and those determined using the seabed pressure transducers, with
the effect of the tide eliminated. This period is chosen as the longest contiguous set of data from the
seabed pressure transducers. Figure 8 is a plot of all the wave height data for the waverider buoy
and seabed pressure transducer.
0
2
4
6
8
10
12
W
a
v
e
h
e
ig
h
t
(
m
)
Waverider buoy
Seabed PT
25 Nov '01 24 Nov '01 26 Nov '01 27 Nov '01 28 Nov '01 29 Nov '01 30 Nov '01
Figure 7 comparison of wave heights
from waverider buoy and seabed PT
for the period 24
th
Nov 01 to 30
th
Nov 01
0
0.5
1
1.5
2
2.5
3
3.5
4
4.5
5
0 1 2 3 4 5 6 7 8 9 10
Waverider buoy - Hs (m)
S
e
a
b
e
d
P
T
-
H
m
o
(
m
)
Figure 8 Comparison of wave heights
from waverider buoy and seabed PT
for all data
Although there is a large amount of scatter in the data, a clear relationship between the wave heights
from the waverider buoy and seabed pressure transducer can be seen. At small wave heights, there
is an approximately linear relationship between the waverider buoy and seabed pressure transducer,
whilst at larger wave heights the wave height derived from the seabed pressure transducer saturates
13
to a maximum of about five metres. This saturation is due to larger waves breaking before they
reach where the seabed pressure transducer is sited. It is worth noting that this saturation limit
corresponds well with the wave height breaking limit in about 6.5 metres of water; the water depth
in which the pressure transducers are deployed. The line shown on Figure 8 has the following
characteristics:
H
pt
= 0.6 H
wr
H
wr
< 5.0
H
pt
= 5.0 2.0 e
-(Hwr 2.0)/3
H
wr
> 5.0
0
2
4
6
8
10
12
14
W
a
v
e
p
e
r
io
d
(
s
e
c
s
)
Waverider buoy
Seabed PT
25 Nov '01 24 Nov '01 26 Nov '01 27 Nov '01 28 Nov '01 29 Nov '01 30 Nov '01
Figure 9 Comparison of waverider buoy
and seabed PT wave periods for period
24
th
Nov 01 to 30
th
Nov 01
0
2
4
6
8
10
12
14
0 2 4 6 8 10 12 14
Waverider buoy wave period (secs)
S
e
a
b
e
d
P
T
z
e
r
o
c
r
o
s
s
in
g
w
a
v
e
p
e
r
io
d
(
s
e
c
s
)
Figure 10 Comparison of waverider
buoy and seabed PT wave periods for
all data
Figure 9 shows the comparison between the wave period from the waverider buoy and the zero
crossing period calculated from the seabed pressure transducer data for the period 24
th
Nov 2001 to
30
th
Nov 2001. The wave period from the waverider buoy is only reported to within 1 second, so a
5-hour moving average has been used to improve the data. Figure 10 is a scatter diagram containing
these two wave periods for all of the data collected.
The data of Figure 10 shows that the zero crossing period calculated from the seabed pressure
transducer data follows the waverider buoy wave period reasonably well. However, on the 29
th
/30
th
Nov 01 the two periods can be seen to diverge, which together with similar incidents cause the
large amount of scatter shown in Figure 9. This is most likely due to the wind direction changing
and the source of waves for the seabed pressure transducer and waverider buoy being different. This
illustrates the limit of validity for using waverider buoy 62106 to derive the wave climate at
LIMPET. The linear fit, necessarily going through the origin has the relationship below.
T
pt
= 1.1 T
wr
2.5 Specification of sea-states
It is elsewhere argued that sea-states for testing shoreline and nearshore wave energy converters
should be specified at the 20-metre depth contour. No wave data is available at this location and so
it is necessary to infer the data using the relationships produced above.
The only data set that is sufficiently large to provide an annual set of sea-states is from the
waverider buoy. Thus, it is necessary to derive a transfer function to convert the wave parameters of
the waverider buoy to estimates of parameters at the 20-metre depth contour. This in effect means
using a transfer function from the waverider buoy to the LIMPET site and then from the LIMPET
14
site to the 20-metre depth contour. At the 20-metre depth contour typical waves will not have
experienced any depth-induced wave breaking, thus the influence on significant wave height due to
saturation should be ignored. Considering the one-to-one relationship between significant wave
height at the 20-metre contour and LIMPET site, the significant wave height at the 20-metre depth
contour is estimated to be equal to 0.6 of the significant wave height reported by the waverider
buoy. The one-to-one relationship between wave periods at the LIMPET site and 20-metre depth
contour means that the zero crossing period is estimated to be equal to 1.1 of the wave period
reported by the waverider buoy. Figure 10 shows a scatter diagram for the significant wave height
and zero crossing period at the 20-metre depth contour.
Zero crossing period
Hs <7 7-9 9-11 11-13 >13
Weighting 0.27 0.33 0.03
<2
Estimated Energy Contribution 5% 14% 2%
Weighting 0.17 0.11 0.05
2-4
Estimated Energy Contribution 22% 22% 14%
Weighting 0.004 0.02 0.003
4-6
Estimated Energy Contribution 2% 11% 3%
Weighting 0.002
>6
Estimated Energy Contribution 4%
Figure 11 Scatter diagram for 20-metre depth contour at Islay
Only sea-states that make at least a 2% contribution to the annual wave energy are included in the
scatter diagram. The estimated average incident wave power at the 20-metre contour is 24 kW/m.
From this the following test sea-states at the 20-metre depth contour are defined.
Spectrum Hs (m) Tz (s)
1 1.5 6
2 1.5 8
3 1.5 10
4 3.0 8
5 3.0 10
6 3.0 12
7 5.0 10
8 5.0 12
9 5.0 14
Figure 12, Representative Sea States at the 20m Contour
These ten sea-states would be used with three tide levels: high tide, low tide and mean water level,
with an equal amount of time assumed to be spent at each tide level.
2.6 Discussion
Although it has been possible to derive a suitable set of test sea-states, this has been achieved using
a relatively large amount of extrapolation and conjecture. This was necessary in respect of the
15
limited availability of data. A more accurate and secure set of test sea-states could be derived using
more extensive data.
This improved set of test sea-states could be produced using a hindcast model together with site
data from the 20-metre contour. A hindcast model, which derives sea-states from wind and
meteorological data, is generally accurate provided that there are no rapid changes in bathymetry
and the site is beyond the wave-breaking zone. Such a model can provide data not only on the wave
height and period, but also the wave spectrum and directionality. Unfortunately, hindcast data is
relatively expensive to obtain (compared to waverider buoy data) and its cost would have to be
included in the production of representative sea-states. The site data would be used to calibrate the
hindcast data, and thus monitoring would only be required for a short period; such calibration being
necessary to identify the extent to which tidal streams and other localised effects not accounted for
in hindcast model affect the local wave climate. However, this would not remove the need for
measurement of wave activity adjacent to the device to help determine instantaneous device
efficiency.
Whilst a better set of sea-states could be produced using other data sets, the sea-states produced in
the wave tank have been shown to be representative of typical sea-states at the LIMPET site. For
illustration, Figure 13 shows recordings of wave tank data (13a to 131c) compared to LIMPET site
data (13d to 13f). Whilst each of the traces differs in detail, there is no apparent systematic
difference between the wave tank and site data, which is confirmed by the wave group parameters.
Thus, the performance of LIMPET as determined from wave tank trials is representative of full-
scale performance.
This is contrary to anecdotal reports of the seas produced in the tank at Wavegen not being
representative. Two reasons for this observation are suggested. Firstly, there will be instances when
the sea-state at the LIMPET site is different from those in wave tank tests due to a different incident
wave spectrum, though the typical sea-state may not differ. This has more significance for singular
events such as maximum wave loads and less significance for productivity testing, however,
provided that the wave spectrum was known, the sea-state could be reproduced. Secondly, temporal
and spatial scaling can influence perceptions so that it is difficult to compare the sea-states using
direct observation. Combining this with the difficulty of estimating a sea-state from direct
observation the authority of anecdotal evidence cannot be considered as very high. Thus, the sea-
states produced in a wave tank are acceptable for most scenarios provided there is recognition that
other time-series can occur because of differing wave spectra. How to deal with these different
spectra and any other wave parameters is discussed later in Section 5, Standardised Evaluation
Procedure.
2.7 References
1. US Army Corp of Engineers, Shore Protection Manual. 1977, Coastal Engineering Research
Centre.
2. Mollison, D. The prediction of device performance. in Power from Sea Waves. 1980.
Edinburgh: Academic Press. p. 135-168.
3. Mansard, E.P.D. and Sand, S.E. A comparative evaluation of wave grouping measures. in
24th International Coastal Engineering Conference. 1994. Kobe, Japan: ASCE. p. 832-846.
4. Funke, E.R. and Mansard, E.P.D. On the synthesis of realistic sea states. in 17th
International Coastal Engineering Conference. 1980. Sydney, Australia: ASCE. p. 2974-
2991.
16
Figure 13a: Sea 22 GF = 0.59, Ji = 1.78
Figure 13b: Sea 2 GF = 0.67, Ji = 1.74
Figure 13c: Inverness D5Te10Pi10L150-2 GF = 0.78, Ji = 1.67
Figure 13d: WAV090102 1600.txt GF = 0.69, Ji = 1.82
Figure 13e: WAV100102 1000.txt GF = 0.57, Ji = 1.87
Figure 13f: WAV100102 2100.txt GF = 0.53, Ji = 1.67
17
3 Task 2 New Model tests of Current LIMPET.
During the construction of the LIMPET collector it became clear that there was a variance between
the survey on which the tank floor bathymetry was based and the actual sea bed profile at site. In
general it was observed, both from dipping at the shoreline and from vessel mounted echo sounder,
that the sea was shallower than anticipated and that it was substantially flat for some distance from
the shore. Whilst the initial survey indicated a 7m water depth at the coastline with an immediate
1:25 gradient to seawards, an interim survey
during the construction phase showed a depth of
approximately 5m close to shore, remaining
substantially constant for some 80m before
starting to shelve at the expected 1:25 slope.
These profiles are compared in figure 14, where
it is seen that the delayed start of the slope
significantly reduces the water depth in which
waves approach the device. A much smaller
percentage of the offshore wave climate may
thus be expected to reach the shore with a
corresponding fall in the performance of
LIMPET. In order to quantify this power loss a
series of model tests were performed in which
the same wave maker settings were used but different sea floor profiles were established in the
wave tank.
Because the study was essentially comparative, the majority of waves used to test the model were
monochromatic. Monochromatic waves have the advantages that they are more determinate than
wave spectra and show resonance more clearly. However, monochromatic waves may not produce a
number of non-linear effects that are significant to the performance of the LIMPET. Thus a small
number of random seas were also tested.
3.1 Seabed profiles
3.1.1 Profile 1
The first profile used has a plateau at 6m below the site datum for 17m from the device, followed by
a 1:25 slope to deep water. This profile best matches that used previously to test the LIMPET
device.
3.1.2 Profile 2
The second profile has a plateau at 6m below the site datum for 17m from the device, followed by a
1:25 slope for 18m and a small step (0.3m). There is then an 80m wide plateau at 7m below datum,
followed by a 1:25 slope to deep water.
Figure 14
18
3.1.3 Profile 3
The third profile has a plateau at 6m below the site datum for 16m from the device, followed by an
abrupt step to a plateau at 5m below the site datum for 48m. A slope of 1:25 continues from this
plateau into deep water. This profile best matches the estimate of the actual sea bed profile, as per
the information available at the time of test.
3.2 Wave test series
The following wave periods and amplitudes were chosen to provide data about the full range of
waves experienced at the LIMPET site.
Full scale wave periods (secs) 7.0, 8.0, 9.0, 10.0, 11.0, 12.0, 13.0, 14.0, 15.0, 16.0
Full scale wave amplitudes (m) 1.0, 2.0
Three water depths were chosen to represent three distinct tide levels of approximately high tide,
mean tide and low tide.
Full scale water depths (m) 8.0, 7.0, 5.5
1
At each water depth two seas with a Bretschneider spectrum were used to represent random seas.
The seas were chosen to represent a sea of average energy period and incident power, together with
another sea with a larger incident power
2
.
Depth (m) Energy period (s) Incident Power (kW/m)
8.0 10 20
8.0 13 60
7.0 10 20
7.0 13 60
5.0 10 20
Figure 15 Irregular Test Seas
3.3 Wave tank and model set-up
Wavegens canal tank was used for the trials. It is 15m long and 3.8m wide. A single paddle creates
waves that are absorbed by a beach at the opposite end of the tank. Baffles containing a double
thickness of netting are positioned 0.35m from both sides of the tank, from the paddle to the beach,
to minimise the influence of transverse resonance. The tank floor rises at approximately 45 from
the bottom of the paddle to meet a 2.3 (1:25) slope that then rises for 7 metres to a plateau that
extends a further distance of approximately 4m to the beach. The three profiles were achieved by
placing the front of the model on the edge of the plateau, moving the model back 2m (80m FS
3
) and
then building the tank floor up with plywood to achieve the final desired profile. The model was set
up so that the site datum was 175mm (7m FS) above the plateau.

1
These water depths are measured from the base of the model, which is 7.0m below the sites datum.
2
The energy period and incident power is approximate for deep water the actual power available at the device is
smaller.
3
FS full scale equivalent
19
The LIMPET model, including the device and gully, was positioned in the middle of the wave tank.
The gully was widened by 25mm (1m FS) at the southern end of the device and 50mm (2m FS)
northern end, so it was a total of 0.6m (24m FS) wide. Breezeblocks covered in a double layer of
netting were placed orthogonal to the direction of wave propagation to extend the shoreline either
side of the gully by 0.5m. Beyond this the waves were allowed to pass unhindered to the beach at
the back of the tank.
3.4 Instrumentation and procedure
The wave tank was instrumented with 6 wave gauges. Three wave gauges were placed in deep
water approximately 3 metres from the paddle; one on the centreline of the tank and the other two
0.6 metres either side of this. The three other gauges were placed along the centreline of the tank at
0m, 1m and 2m from the mouth of the gully. The mouth of the gully was 0.8m from the front lip of
the device.
There were two wave gauges in each of the three chambers of the LIMPET; one placed 20mm
parallel to the back wall and the other 20mm parallel to the front wall. Each chamber was also
connected to a pressure transducer.
The wave gauges were re-calibrated whenever the water depth changed and at least once per day.
The pressure transducer calibrations were checked against the factory setting.
All data readings were taken with 4096 data points using a sampling frequency of 32 Hz. A delay
prior to each test allowed the tank to settle. The length of delay was determined by observation,
waiting for a regular output from the wave gauges; the typical delay was approximately 1 minute.
With the Bretschneider seas the output is not regular and a period of approximately 2 minutes was
allowed for the seas to settle.
3.5 Estimation of incident wave power
To calculate the performance of the LIMPET it was necessary to determine the incident wave
power. In these tests this was achieved by assuming that the incident wave is constructed of
sinusoidal waves and then using a Fourier analysis to deconstruct the time series into a spectral
energy density function. It was then assumed that each individual wave transmits energy at its group
velocity and that the total incident power was the sum of these individual waves. Whilst this is a
reasonable approximation for deep water it is less accurate for shallow water.
As a monochromatic wave enters shallow water the wave changes into a cnoidal form. A Fourier
analysis of this wave will result in sinusoidal waves at the primary wave frequency together with its
harmonics. The above procedure then applies different group velocities for each harmonic.
3.6 Power and capture factor
20
Power for the LIMPET is calculated by integrating the force and velocity for each water column
and then adding these together. The force on each water column is equal to the air pressure in the
column multiplied by the water plane area of the column, which is assumed to be constant and equal
to 0.033m
2
(53m
2
FS). The capture factor was calculated by dividing the power capture by the
incident power at the mouth of the gully with seabed profile 1 over a width of 0.525m (21m FS).
Power capture and capture factor for the monochromatic waves are shown in figures 14 19.
0.0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1.0
6 8 10 12 14 16
Wave period (secs)
Profile 1, a=1m Profile 2, a=1m Profile 3, a=1m
Profile 1, a=2m Profile 2, a=2m Profile 3, a=2m
Figure 16 Capture factor at high water level
0.0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1.0
6 8 10 12 14 16
Wave period (secs)
Profile 1, a=1m Profile 2, a=1m Profile 3, a=1m
Profile 1, a=2m Profile 2, a=2m Profile 3, a=2m
Figure 17 Capture factor at mean water level
21
0.0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1.0
6 8 10 12 14 16
Wave period
Profile 1, a=1m Profile 2, a=1m Profile 3, a=1m
Profile 1, a=2m Profile 2, a=2m Profile 3, a=2m
Figure 18 Capture factor at low water level
0
200
400
600
800
1000
1200
1400
1600
1800
2000
6 8 10 12 14 16
Wave period (secs)
Profile 1, a=1m Profile 2, a=1m Profile 3, a=1m
Profile 1, a=2m Profile 2, a=2m Profile 3, a=2m
Figure 19 Power capture at high water level
22
0
200
400
600
800
1000
1200
1400
1600
6 8 10 12 14 16
Wave period (secs)
Profile 1, a=1m Profile 2, a=1m Profile 3, a=1m
Profile 1, a=2m Profile 2, a=2m Profile 3, a=2m
Figure 20 Power capture at mean water level
0.00
100.00
200.00
300.00
400.00
500.00
600.00
700.00
6 8 10 12 14 16
Wave period (secs)
Profile 1, a=1m Profile 2, a=1m Profile 3, a=1m
Profile 1, a=2m Profile 2, a=2m Profile 3, a=2m
Figure 21 Power capture at low water level
Whilst there is significant scatter of both power capture and capture factor a number of observations
can be made from the data. As expected the profiles with more shallow water result in less power
being captured by the device. The reduction in power is most significant for profile 3, whilst profile
2 only appears to reduce the power capture for long wave periods. There is no apparent peak in the
capture factor that would indicate resonance of the water column or harbour. The resonant period of
23
the water column has been calculated at approximately 5 seconds (assuming an effective water
column length of 6 metres) which is lower than the minimum test period of 7 seconds. The primary
resonant period of the harbour, (that is the period where the harbour length equals of a
wavelength), is however approximately 10 seconds. Since there is no well defined peak in the
response at this period it may be concluded that either this resonance does not occur, or it is so
small that the scatter in the capture factor obscures it.
Lower tide levels result in lower power captures for similar wave periods and amplitudes, though
the capture factor is similar since these waves have less power in shallow water. The larger wave
amplitudes result in lower capture factors for all the cases studied. A doubling of the wave
amplitude caused an approximate halving of the capture factor. Or alternatively the power captured
increases linearly with incident wave amplitude. Thus a significant amount of power is being lost in
column motion; probably due to vortex creation.
An interesting and significant result is obtained if the capture factor for the LIMPET with seabed
profile 3 is plotted using the incident power measured at the equivalent place for profile 1. The
comparison of capture factor for this and profile 1 is shown in
Figure 22
4
. This shows that the reduction in power capture due to seabed profile is largely, if not
completely, due to a reduction in the incident power.
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1
6 8 10 12 14 16
Wave period (secs)
Profile 3 Profile 1
Figure 22: Comparison of device capture factor for profiles 1 and 3
The results for the Bretschneider seas are shown in Figure 23. This show a similar reduction as the
monochromatic seas but interestingly these results may be better conditioned than the
monochromatic results because the frequency spread avoids the existence of strong nodes and anti-
nodes.

4
This also illustrates the potential effect of reflections when calibrating the incident wave on the capture factor. The
divergence seen at a period of 13 seconds is due to changes in the incident power and could be the result of just a 1% of
the energy being reflected; the standing wave created interfering constructively for profile 1 and destructively for
profile 3.
24
0.00
0.10
0.20
0.30
0.40
0.50
0.60
0.70
Profile 1 Profile 2 Profile 3
Te =10, Pi =20 LWL Te =13, Pi =40 LWL Te =10, Pi =20 MWL
Te =13, Pi =60 MWL Te =10, Pi =20 HWL Te =13, Pi =60 HWL
Figure 23 Capture Factor In Bretschneider Seas
3.7 Incident wave behaviour
As the waves approach the LIMPET they change from a reasonable approximation of a sinusoidal
wave to a cnoidal wave due to a reduction in water depth. The waves may also spill or break prior
to reaching the LIMPET gully, also due to the water depth. It was found that this only occurred
when the wave amplitude was 2 metres, with no wave spilling/breaking occurring with 1 metre
amplitude waves.
The table below shows the full-scale wave period at which wave spilling started to occur. At
periods longer than these, the point at which the wave spilling/breaking occurs moves further away
from the gully. It would appear from the splash when the wave hits the device that the wave
spilling/breaking results in a reduction of incident wave power reaching the device.
Water depth
8.0m 7.0m 5.5m
Position 1 - - 12s
Position 2 - 16s 12s
Position 3 12s 7s 7s
Figure 24
In addition to general wave spilling/breaking some localised wave spilling and/or splashing occurs
when the reflected/radiated wave from the device meets the incoming wave.
Incoming waves run-up the front wall of the LIMPET, inclined at 40 to the horizontal. The water
on this wall then either shoots over the top of the device, or falls back down the wall into the sea.
This falling water causes visible water shear immediately in front of the front wall and a wave to be
sent away from the device. In addition to this the falling water entrains a large number of bubbles,
some of which enter the southern water column if the water level is sufficiently low. This has only
been observed when the water level is 5.5 metres. This flow regime is discussed more in respect of
Task 5.
25
3.8 Water column movement
Apart from the wave amplitude and frequency, the movement of the water columns appears to be
affected by the seabed profile and water depth (tide level). There were also differences between the
amplitudes measured at the front and back of each column and in some tests between the columns
themselves.
The movement of the water at the back of the column was generally larger than the motion at the
front, the difference being largest at higher wave frequencies. This is due to inertia of the water
entering the column carrying it to the back. In addition to this the water in the column sloshes fore
and aft, which appears to have a natural period of approximately 0.5 secs (3 secs FS). This slosh is
most evident with low tide and large amplitudes. An example of the traces for the fore and aft
gauges is shown in Figure 25. With seabed profiles 1 & 2 the slosh is only significant at low tide.
With profile 3 the slosh in the columns is evident at all depths, though it is still more significant at
low tide.
It appears that this non-linear motion is due to the parametric excitation of the water column from
the cnoidal or spilled/broken waves. The waves effectively drive the column with a periodic series
of impulses, rather than the sinusoidal varying forcing function of linear theory. With an impulse
excitation the water column resonates at its resonant frequencies both in heave and slosh to give the
motion shown.
-20
-15
-10
-5
0
5
10
15
20
25
0.000 1.000 2.000 3.000 4.000 5.000 6.000 7.000
Time (secs)
Rear Front
Figure 25: Motion at rear and front of central water column for profile 3 at low tide,
Period = 10secs, Amplitude = 1m
Differences in the water motion between the columns are generally small, though the motion of the
southern column is generally slightly larger than the other columns. This appears to be linked to a
tendency for the waves to be pushed across the bay by the northern headland, illustrated, when the
incident power is sufficient, by the water shooting off the southern end of the device.
26
-30
-20
-10
0
10
20
30
0.000 1.000 2.000 3.000 4.000 5.000 6.000
Time (secs)
Southern column Central column Northern column
Figure 26: Water column motion for profile 3 at mean tide,
T = 10 secs, a = 1m
Figure 26 shows typical water column motions, calculated by taking the average of the front and
back gauges. Extreme difference between the water column motions is illustrated in Figure 27.
-20
-15
-10
-5
0
5
10
15
20
25
0.000 1.000 2.000 3.000 4.000 5.000 6.000
Time (secs)
Southern column Central column Northern column
Figure 27: Water column motion for profile 3 at low tide,
T = 7 secs, a = 2m
27
If the water columns move out of phase from each other air will be cycled between the columns
rather than through the turbine, rendering the water column motion less effective. If the air is
assumed to be incompressible then typically this cycled flow is about 1-2% of the turbine flow. The
largest differences in water column motion result in a cycled flow of approximately 20% of the
turbine flow, though this rarely occurs.
Figure 28: Comparison of capture factors
3.9 Influence of the chamber air volume
A model scaled by its linear dimensions does not accurately model the effect of air compressibility.
Differing relationships can be used for this, however a common relationship is to reduce the air
volume by the scale factor squared (rather than the scale factor cubed as a Froude scaled model
would), which keeps the same equivalent spring rate at full and model scales. Thus an airbox is
used to achieve a more accurate modelling of the effects of the air above the water columns.
However previous studies have shown that the use of an airbox has little effect on the power capture
of the LIMPET and adds an additional area where problems may occur in the testing (Dustan 1999).
Consequently all of the results referred to above relate to the model tests without an airbox,
however a small number of tests were performed using an airbox to ensure that similar results
would be obtained.
A 0.6m
3
chamber was used for the tests. Correct scaling would require an airbox of approximately
1.2m
3
, however this size was not available and the smaller receiver was used since it would give an
indication of the effect. The comparison between the capture factors (Cf) for the LIMPET at
Position 1 with a water depth of 7.0m is shown in Figure 28.
Figure 28 shows a change in Cf, depending upon whether an airbox is used or not. The effect is
similar to that achieved with a reduction in the damping applied to the column, a higher Cf at high
frequencies together with a smaller Cf at lower frequencies. The compressibility of the air reduces
the airflow through the dampers and thus smaller pressures are created.
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
5 7 9 11 13 15 17
Wave period (secs)
Airbox, nD=5 No airbox, nD=5
28
The use of an airbox also changes the
dynamics of the water column motion. This
is most evident if column velocity and
displacement are plotted as shown in figure
29 for seabed profile 1, full-scale wave
period of 14 seconds and full-scale wave
amplitude of 1 metre.
The phase is moving in a clockwise
direction and the major difference occurs
as the water column is falling. The velocity
of the column shows a higher peak and
distinct drop when an airbox is attached.
This appears to be due to the air spring,
driven by the incident cnoidal wave,
allowing more free motion of the water
column.
3.10 Discussion & Conclusions
This short study illustrates that there are major uncertainties associated with determining the power
available in the incoming wave, both in mathematical and tank modelling. However it does confirm
the importance of the seabed profile in the approach to the device, which can change the power
capture by more than a factor of two. Moreover this reduction in power capture appears largely due
to power losses as a result of wave shoaling or breaking and not a reduction in device efficiency.
Our best estimate is that the energy actually incident at the LIMPET collector is 66% of that
predicted by the pre-construction study. Further studies and a better understanding of the change in
wave power as it approaches the shoreline will enable the performance of the LIMPET to be more
accurately determined. In addition, this understanding would help to identify a suitable method and
position to measure the wave climate at the Islay LIMPET site.
The study has also illustrated that in certain circumstances the motion of the water column is
significantly non-linear. The non-linear motion is apparently due to the parametric excitation of
cnoidal waves. In addition to this it was also seen that air compressibility had an effect on column
motion. Whilst the air compressibility did not appear to affect the potential power capture in the
situation tested, its effect has not been determined for cases where the forcing function is
significantly non-linear. This topic merits further study.
-0.1
-0.05
0
0.05
0.1
0.15
0.2
-50 -40 -30 -20 -10 0 10 20 30 40
Displacement (mm)
Airbox No airbox
Figure 29 Phase diagram for water column motion
Airbox No airbox
29
4 Task 3. Standardise Evaluation Procedure
There is a recognised need for a common standard within the wave energy community for
evaluating device performance. Work task 3 draws on the experience of the LIMPET project and
that of the previous 75kW QUB prototype to propose a methodology for assessing shoreline
systems.
4.1 Modelling and analysis of shoreline and nearshore wave energy converters
4.1.1 Ocean wave analysis techniques
For the modelling and analysis of wave energy converters, it is initially necessary to examine how
ocean waves can be represented. It is clear to anybody who has spent time looking at ocean waves
that they are complex, with the same conditions never reoccurring exactly. However, it is also clear
that the waves during any particular period have some average or common characteristics. The
challenge for the modelling of ocean waves is to decide what characteristics or properties of the
waves should be reproduced and how this is to be achieved. Having decided this, it is possible to
compare the response of wave energy converters to ocean waves with the response of model wave
energy converters to modelled waves to determine the extent to which the ocean waves are
appropriately modelled. There are three fundamental methods ways that ocean waves can be
analysed for characterisation: spectral analysis, statistical analysis and phenomenological analysis.
4.1.1.1 Statistical analysis
T1 T2 T3
H1
H2
H3
Figure 30: A typical ocean wave surface profile
Figure 30 shows a typical surface profile for an ocean wave. The surface profile can be reduced into
a discrete series of individual waves using the up crossing of the surface through the mean still
water level. Each wave can be assigned a period, T
i
, equal to the time between up crossings, and a
height, H
i
, equal to the distance from the minimum of the wave trough and the maximum of the
wave crest. Statistics can now be used to define characteristics of the sea.
Defined this way, the mean wave period, known more commonly as the zero crossing period, T
z
, is
the average of the wave periods, T
i
. The mean wave height is defined as the average of the wave
heights, H
i
. More commonly however, the significant wave height, defined as the average height
of the largest 1/3 of waves, has traditionally been used. Though there is no analytical relationship
between the mean wave height and significant wave height, generally the significant wave height,
H
s
, is approximately equal to 1.6 times the mean wave height. Thus statistically, the significant
wave height and zero crossing period are given by the equations below.
30

=
=
n
i
i n s
H H
1
1
6 . 1 (1)

=
=
n
i
i n z
T T
1
1
(2)
A statistical analysis can also be used to produce probability distribution functions (PDFs) for the
wave period, wave height or a two-dimensional distribution of wave period and height. An idealised
distribution of wave height in deep water is shown in Figure 31. Statistical analysis of the wave
climate is often used to define the probability of an extreme wave. This is achieved by matching the
actual PDF to a theoretical PDF and then extrapolating to larger wave heights. In deep-water the
Rayleigh PDF has been found to be successful in predicting the probability of large waves, however
in shallow water this function has been found to overestimate the probability of the expected wave.
0
0.01
0.02
0.03
0.04
0.05
0.06
0.07
0 0.5 1 1.5 2 2.5 3 3.5
Wave height (m)
P
r
o
b
a
b
i
l
i
t
y
Figure 31: Probability density function for deep-water wave height
It is quite easy to see why the Rayleigh PDF is inappropriate for shallow water. The Rayleigh PDF
has a finite probability for any wave height, however in shallow water the maximum height of any
wave is limited by the water depth. Typically, waves break when their height is approximately 78%
of the water depth. Though waves converging from different directions may result in larger waves,
there is still a very definite limit to wave height in shallow water and thus the Rayleigh PDF is
inappropriate for shallow water. There is no currently accepted statistical method for determining
the probability of extreme waves in this environment, though the Rayleigh PDF is often used to
provide an upper bound to the probability (i.e. the probability of a particular extreme wave
occurring in shallow water is lower than predicted by the Rayleigh PDF).
4.1.1.2 Spectral analysis
The standard spectral method for ocean waves is to produce a Fourier analysis of the surface
profile. A Fourier analysis defines the surface profile as the linear superpositioning of discrete
31
sinusoidal waves of different amplitudes, frequencies and starting phase. Conveniently, in deep
water this is a reasonable approximation of the physical situation. That is, each individual wave
identified in the analysis could be attributed to particular events such as storms or winds that have
occurred elsewhere in the ocean. The output from this analysis can easily be converted into a
spectral energy density function since the energy for each individual wave can be considered
independently. Analysis of a typical deep-water surface profile produces the spectral energy density
function shown in Figure 32.
0 0.05 0.1 0.15 0.2 0.25
Wave frequency (Hz)
S
p
e
c
t
r
a
l

e
n
e
r
g
y

d
e
n
s
i
t
y
Figure 32: typical spectral energy density for deep-water ocean wave
Having determined the spectral energy density function, S(f), it is possible to calculate
characteristics of the sea using moments of the spectrum. The n
th
moment, m
n
, of the spectrum is
defined as

=
0
). ( . df f S f m
n
n
(3)
An average wave height can be derived from the average energy in the sea, which is proportional to
the zeroth moment. Confusingly, this is also often called the significant wave height, though
conveniently is normally numerically very similar to the significant wave height defined by
statistical methods (the subscript mo is used to distinguish this significant wave height from the
statistical significant wave height).
0
4 m H
mo
= (4)
A measure of the average wave period can be obtained from the spectral energy density function by
identifying the wave period that has the highest energy density. This is known as the peak period,
T
p
. However, whilst this is easy to do for the spectra shown in Figure 32, a two-peaked wave
spectrum would present more difficulties to identifying the peak period. Such two-peaked wave
spectra often occur where the waves derive from two separate sources, such as two distinct weather
32
systems. A common alternative used in wave energy is to define a wave period that would have the
same power per unit width in deep water as the sea, assuming that it was a sinusoidal wave of
height, H
mo
. It is relatively easy to show that this energy period, T
e
, is given by
0
1
m
m
e
T

= (5)
It can also be shown, though more problematically, that the zero crossing period is given by
2
0
m
m
z
T = (6)
A spectral analysis is typically used to reproduce the wave climate at a particular instance for
productivity modelling. This is achieved by using a standard spectral shape together with particular
values of the significant wave height and energy period. Most commonly, the Bretschneider
spectrum is used for this purpose since it permits the significant wave height and peak frequency
(which assuming a Bretschneider spectrum can easily be converted to another measure of wave
period) to be used directly. The use of this spectrum is also simplified by its inclusion in standard
software for wave tank paddles.
Bretschneider spectrum,
|
|
.
|

\
|
|
|
.
|

\
|

|
|
.
|

\
|
=
4 5
2
4
5
16
5
) (
f
f
Exp
f
f
f
H
f S
p p
p
mo
(7)
Whilst the Bretschneider spectrum is a good approximation for the waves generated by a single
storm and that have reached full maturity, there are many instances when the spectrum is not so
appropriate. For example, the waves at a particular instance could be the combination of two
separate storms, or the waves are generated too close to where they are measured for them to reach
full maturity.
Wave spectra can also include a parameter associated with the direction that particular waves are
travelling. This is generally called the spreading parameter, s. The effect of the shoreline is to
reduce the value of this parameter as the waves will tend to travel perpendicular to the coastline.
) ( ) ( ) , ( f f S f S =
Where
) 1 2 ( ..... 7 5 3 1
2 ..... 8 6 4 2 1
) (


=
s
s
f

for
2

< , otherwise 0 ) ( = f
The decomposition and construction of waves based on linear superpositioning is effective because
of its physical applicability in deep water. However, in shallow water the application of Fourier
analysis will not provide a spectrum related to the individual waves, but a quasi-spectrum
containing harmonics from the non-sinusoidal wave profiles. The reconstruction of the wave profile
cannot be achieved by simply adding together the individual wave components. Thus, parameters
calculated using a linear spectral analysis have a limited validity in shallow water.
The Korteweg de Vries equations have been used to produce a spectral analysis using cnoidal
waves, making the results more applicable to shallow water. However, these analyses are extremely
complex and have not been used extensively. Thus whilst they may become appropriate with further
33
development they are not currently suitable for the definition of wave climates in shallow water for
the analysis of wave energy converters.
4.1.1.3 Phenomenological analysis
A phenomenological analysis examines events that occur in ocean waves. A classic example of
phenomenological analysis is the estimation of maximum wave loads derived from experiments
using solitons, i.e. a single wave in otherwise still water. These tests are designed to model the
single salient event, namely a wave breaking onto the structure. Thus, the significance of the test
depends on a particular phenomenon, the breaking wave, as opposed to a general characteristic of
the ocean waves. It is generally difficult to quantify a phenomenological analysis because the
frequency or magnitudes of events are not easily calculated. However, the analysis of phenomena is
important for the classification of non-linear characteristics.
Whilst most phenomena are related to critical failure scenarios, there are also circumstances where
they can be considered in the analysis of productivity. Wave groupiness is a phenomenon that could
affect device performance and would not be identified by either a statistical or a spectral analysis.
Wave groupiness is the tendency for waves to arrive in groups, with periods of large waves
followed by lulls in the wave activity. The hydrodynamics of this process are poorly understood and
its occurrence has not been analysed in any detail. When modelling ocean waves the simulated sea
should have a similar degree of groupiness to the ocean. Idiosyncratic evidence from comparing
observations of waves in wave tanks and in the ocean indicates that wave groupiness is not always
correctly modelled.
4.1.2 Modelling ocean waves for shoreline and nearshore wave energy converters
Whilst it is possible to envisage a number of supplementary circumstances where ocean waves
would need to be modelled, e.g. determining the plant availability for repair and maintenance tasks,
defining wave climates suitable for installation, the most significant requirements to model ocean
waves for wave energy converters are in determining the maximum expected wave loads and the
average productivity. These two requirements are considered separately since they are likely to
involve modelling very different sea conditions.
4.1.2.1 Modelling ocean waves for device productivity
Device productivity can be defined as the average amount of power captured by the wave energy
converter and converted into useful energy. Typically, this average is calculated over a year with
the implicit assumption that each year will have approximately the same amount of energy. To
calculate an average power capture, P , a set of n seas is modelled, with the power capture, P
i
, from
each sea weighted depending on its relative frequency of occurrence, w
i
.

=
=
n
i
i i
P w P
1
(8)
Fundamentally, this procedure requires seas that have occurred at different times to be treated as
equal. That is, the power captured by the wave energy converter is the same for all the seas grouped
together. In the analysis of deep-water wave energy converters, this has been achieved by using a
scatter diagram of the annual sea states based on significant wave height and energy period.
34
Whilst the significant wave height and wave period are fundamental in determining the
performance of a wave energy converter, this procedure assumes that other factors associated with
the ocean waves have an insignificant effect on performance. Thus, the same power capture for a
sea with a particular wave height and period would be obtained irrespective of the spectral shape,
wave groupiness or other particular phenomena. However, adding additional parameters for
defining a sea state would increase the number of test seas rapidly and so any additional parameters
must be fully justified.
Evidence from prototype shoreline devices, as well as wave tank tests, indicate that for the analysis
of wave energy converters in shallow water an additional index of water depth should added to the
scatter diagram. Thus, each sea would be defined by significant wave height, wave period and water
depth. The introduction of an additional parameter multiplies the number of seas to be tested and
given the requirement to provide a relatively rapid assessment of device performance, there appears
little scope to increase the number of parameters defining the sea states beyond the three already
identified.
It would appear sensible that the waves should be defined as those that are incident on the wave
energy converter, and indeed this is the case for the analysis of deep-water wave energy converters.
Unfortunately, in shallow water this approach cannot be adopted due to the occurrence of wave
breaking. Waves break near the shoreline as the ratio of the wave height to water depth increases.
The wave heights and water depths at typical sites for shoreline wave energy converters mean that
breaking waves are not rare or extraordinary events and thus cannot be ignored or marginalised.
This occurrence of broken waves causes difficulties with defining the wave climate at the position
of the device.
Defining the sea state using the surface elevation at the position of the wave energy converter
causes two difficulties. Firstly, a waves surface profile cannot be easily used to distinguish the
condition of the wave, i.e. whether it has or has not broken. The particle velocities in broken waves
are very different from those in unbroken waves and waves with the same surface profile will not
interact with a wave energy converter in the same way. Secondly, broken waves cannot be
generated directly by a wave tanks wave maker, but shoaling or superpositioning are required to
cause the waves to break. It is not clear how a wave maker could be driven to achieve this.
A sufficient depth of water to define the wave climate is required to avoid wave breaking and where
linear superpositioning of the waves can be used with reasonable accuracy. The Ursell number can
be used as a measure of non-linearity. If the Ursell number is less than 5.0 then each wave
component will deviate less than 10% from that of a sinusoidal wave, which for typical waves
equates to a water depth of 20m. A water depth of 20m also ensures that there would be minimal
depth-limited wave breaking.
Defining the wave climate at the 20m contour means that for comparative analysis the bathymetry
must also be specified. This is because the bathymetry influences shoaling, wave breaking and the
wave energy incident on the device. The bathymetry at the LIMPET site is used as the standard
bathymetry since data for other sites is not available. As a simplified profile this consists of a 70
metre wide plateau, followed by a 1:20 slope to deep water. This profile is shown in Figure 33.
35
Water Depth
for Model
1:20 Slope
Position of Wave Probe
70m
2
0
m
Figure 33: Bathymetry for comparative model testing
4.1.2.2 Modelling ocean waves for extreme loads
Three different techniques of modelling ocean waves for the determination of extreme loads on
wave energy converters may be considered. Each technique has different advantages and
disadvantages, their significance depending on the situation. However, it should be noted that none
of these modelling techniques could guarantee finding the maximum loads. This is a common
shortcoming to all extreme event analyses.
The first two methods rely on defining the worst sea state that is expected to occur once every 100
years. Though in general it is reasonable to expect that this will occur in seas with the largest waves,
it is not possible to be definitive about the significance of other parameters such as the wave period,
spectra shape or spreading parameter. Indeed, within coastal engineering there is an implicit
assumption that these parameters are not significant since in general the worst sea state is defined
only by the significant wave height.
In the first method, having defined the extreme sea-state, the waves are modelled for the equivalent
of 3 hours at full scale. The loads experienced during this test are used to generate a cumulative
distribution function. This function is extrapolated to identify the design load, where the probability
of a higher load occurring is below a particular value e.g. 0.1%. This procedure has the advantage
that at typical model scales used for extreme loading, e.g. 1:100, the test lasts about 20 minutes.
This provides the opportunity to test a range of different sea-states containing different spectra or
other parameters. The disadvantage with this method is that the statistical extrapolation of the test
data to an unmeasured load is generally unsafe.
The second method uses the same extreme sea-state as in the first method, but runs the test for a
significantly longer time. The advantage of this approach is that the design load does not have to be
36
extrapolated from the data set. The disadvantage is that the longer testing time means that other sea-
states or scenarios cannot so easily be explored.
The third method uses a phenomenological approach by generating a single event, typically a large
breaking wave, to occur. A range of single events can be created, the largest loads found providing
the design loading. The advantage of this method is that the loading is associated with a definite
single event and thus is more easily analysed with respect to the hydrodynamics. The disadvantage
of this method is that these events may not be able to occur naturally and may only be created in a
wave tank. Moreover, it is difficult to assign a probability to any event. Thus, the design would
allow for loads that could never happen, or have an extremely low probability of occurrence. The
final disadvantage of this method is that the generation of these types waves requires specialised
knowledge and understanding, which may not be available.
The chosen method of estimating extreme wave loads is required to be both relatively fast and
simple. These requirements favour the first method, which does not have the specialist requirements
of the phenomenological approach, whilst being less time consuming than the second method. The
main weakness of the method is that it uses a statistical method to estimate the extreme load.
However, it would be expected that any extrapolation of the data would tend to result in a slight
over-estimation, thus providing some degree of confidence that they represent conservative
estimates. For consistency, the same bathymetry and site of wave climate be used as for the
productivity testing.
4.2 Issues in the wave tank testing of shoreline and nearshore wave energy converters
Researchers experienced in the use of wave tanks for the analysis of wave energy converters may
consider many of the issues raised here as self-evident or obvious. No apology is made for this. By
ensuring that these minimal considerations are made regarding the use of wave tanks for testing, it
is hoped that results obtained will be more reliable and thus one will be able to compare results
between devices tested at different institutions with increased confidence.
4.2.1 Model scaling
Of fundamental importance in wave tank testing are the scale and the
scaling laws used for the tests. Experience has shown that normally
in hydraulics the hydrodynamics is dominated by the interplay of
two types of forces. For wave energy converters, and most marine
structures, this has been found to be the inertial and gravitational
forces. Maintenance of a constant ratio of these forces is known as
the Froude criterion and is a fundamental requirement for the testing
of wave energy converters. Thus, it is assumed that forces due to
viscosity, surface tension and compressive stress are negligible.
Cases where this is not the case will be discussed later.
Froude criterion - =
gL
V
constant
Conveniently, the Froude criterion is maintained by using a
geometrically scaled model, i.e. all dimensions are simply reduced
Parameter Scaling
ratio
Length S
Angle 1
Mass S
3
Force S
3
Pressure S
Time S
1/2
Velocity S
1/2
Flow rate S
5/2
Power S
7/2
Figure 34 Scale ratio for
standard parameters used
with wave energy
converters
37
by the same scale factor. Figure 34 shows that appropriate scaling ratios, where S is the dimensional
scale factor.
For particular cases in the analysis of wave pressures the effect of compressive stresses are
significant. These occur when air trapped by a breaking wave exerts pressure on a solid structure. In
these cases, the wave pressures need to be scaled using the Cauchy criterion.
Cauchy criterion - =
E
V
2

constant
However, these impact pressures are complicated by the different aeration quantities of fresh and
salt water, resulting in differing values for the bulk modulus of elasticity, E. Loads due to entrapped
air in breakers can be identified by very high pressures of short duration. This spike is often
followed by an oscillation of the pressure due to the resonant frequency of the entrapped bubble.
Loads of this type need to be scaled using the Cauchy criterion.
The scaling of breaking waves is of particular importance to the modelling of shoreline and
nearshore wave energy converters. The point that waves break and the type of breaker that occurs
depends mostly on the ratio of wave height to water depth and the slope of the seabed. Both of these
are maintained using geometric scaling, indicating that the breaking of waves is dominated by the
interplay of gravitational and inertial forces, and thus the use of the Froude criterion for scaling is
appropriate. The dissipation of energy in a broken wave is complex and not fully understood,
however it is thought that the rate of energy dissipation, as a proportion of total energy, is correctly
modelled with geometric scaling [Hughes Steven A., 1993 #221].
For oscillating water columns (OWCs), the compressibility of the air in the plenum chamber must
also be scaled correctly. Assuming that the change in volume is only a small proportion of the
volume of the plenum chamber, the compressibility of the air can be adequately scaled by using a
chamber volume scaled to in proportion to S
2
(as opposed to S
3
that geometric scaling would
produce). Whilst provision of a geometrically inflated air chamber is relatively unproblematic for a
fixed OWC, it creates significant problems with the modelling of floating OWCs. Either the
additional volume must move with the OWC, with problems of stability and inertia, or the
additional volume is stationary and connected to the OWC with a flexible coupling, with problems
regarding the constraint of the OWCs motion and losses of energy in the coupling.
Linear analysis would predict that the effect of the air compressibility is to change the resonant
frequency of the water column, potentially dramatically influencing device performance. However,
in modelling the LIMPET OWC it was found that the overall performance was insensitive to the
effective compressibility of air in the plenum chamber. Whilst this indicates that air compressibility
does not need to be accurately modelled for LIMPET, this may be due to the high natural frequency
of the LIMPETs water column and may not be valid for OWCs that have resonant frequencies
closer to the wave frequencies.
The productivity modelling of a wave energy converter depends not only on the correct modelling
of the hydrodynamics, but also on the modelling of the dynamics of the power take-off mechanism.
At model scale, power levels are particularly small, where the power extracted is often only a
fraction of a watt. Moreover, the power is halved for every 18% reduction in scale. For this reason it
is best to use as large a scale as possible for productivity modelling. In addition, particular attention
should be paid to any mechanisms used for the power take-off that could be the source of parasitic
losses. Therefore, it is advised that measurements of force and velocity (or their equivalent) should
38
be made as close to the primary wave/structure interface as possible; intermediate bearings, belts
and pulleys should be eliminated wherever possible.
The design of the full-scale power take-off mechanism may make it difficult to replicate accurately
at model scale. For example, a Wells turbine running at constant speed has a linear relationship
between the flow rate and pressure drop across the turbine, though due to its simplicity an orifice is
commonly used to model the turbine. However, the pressure drop across an orifice is proportional
to the flow rate squared, clearly different from a Wells turbine. Moreover, even systems used to
more accurately model the Wells turbine such as tubes containing a honeycomb lattice do not match
exactly the turbine characteristics. Other power take-off mechanisms are likely to present similar
problems and their design should be considered carefully. However, because the energy extracted
from a wave energy converter in any one wave-cycle is generally small in comparison to the energy
stored in inertial and gravitational energy the method of extracting energy used is not generally
critical.
The requirements for extreme sea-state testing of models are not as demanding. Indeed, it is likely
that the model will be of a smaller scale to enable the same wave tank to be used for the
productivity and extreme sea-state tests.
4.2.2 Selection of wave tank facilities
It is assumed that in general the largest and most sophisticated wave tank available will be used for
any particular test. However, there are a number of circumstances where it may be more appropriate
to use a 2D wave flume.
2D wave flume tests are particularly suited to parameter studies where the objective is to understand
an aspect of the underlying hydrodynamics. In this respect, 2D wave flumes are particularly suited
for analysis in conjunction with flow visualisation and PIV systems. However, as with all
idealisation the dependability of the results should be checked against data from 3D scenarios.
Results from 2D or quasi-2D tests, i.e. tests where the wave energy converter is a significant
proportion of the width of the tank, should be used to provide productivity data only with extreme
caution. The effect of the narrow tank on the device productivity is indeterminate; depending on the
hydrodynamics, the power capture either may be over or under-estimated. Moreover, the influence
may not be consistent for a particular device, but may depend on the particular waves being used for
testing.
4.2.3 Wave tank calibration and testing
Calibration of the wave tank is performed to ensure that the wave energy converter will experience
the desired sea-states. Normally, this is achieved by removing the model and any coastline from the
tank and measuring the waves close the point that the device will be sited. A reflection analysis can
be used to isolate the incident wave from the reflected wave.
In section 4.1.2.1 it was argued that the wave climate for testing of shoreline and nearshore wave
energy converters should be defined at the 20m contour and that the wave tanks seabed profile is
used to appropriately transform the waves as they move into shallower water. With this procedure,
it is particularly important that the effect of the bathymetry is correctly modelled. This can be
checked using the shoaling coefficient for monochromatic waves at different frequencies and
amplitudes.
39
The shoaling coefficient relates the change in wave height due to changes in the water depth. The
most commonly used shoaling coefficient is based on linear wave theory and derived using the
conservation of energy. This is given by Tucker [Tucker 1991, pg 272] as,
( ) | |
2 / 1
0
tanh 2 sinh / 2 1
1
h h h
H
H
s
s
+
= (9)
Where, h is the water depth, k is the wave number derived from ) tanh(
2
kh k
g
=

This relationship appears to be reasonably accurate except in, or near, the surf zone. Prior to wave
breaking the shoaling coefficient becomes non-linear and no generally accepted expression for the
shoaling coefficient exists in this region. Away from the surf zone, deviation from this shoaling
coefficient indicates a parasitic energy loss. Losses due to boundary layer friction (BLF) will occur
at both full and model scale, however this will generally only represent a minimal energy loss.
Attenuation due to BLF,
( )
( )
|
|
.
|

\
|
|
|
.
|

\
|
+
+
=
L
h
L
h
L
B
L
h
p
T BC
X
Exp

4 4
2 4
sinh
sinh
2
(10)
Where
- percentage reduction in wave height
X
p
horizontal distance in wave flume
B - tank width
L wavelength
C wave celerity
T wave period
h water depth
- kinematic viscosity
Typically, this is less than 5%. A similar percentage of losses would be expected at full scale for
most coastlines. Losses that are more significant can occur at model scale due to flexure of the
seabed and/or leakage of water past the seabed; if identified these would need to be eliminated.
Transverse waves do not generally occur during calibration, because there is no source to generate
them, however when the model and coastline are replaced waves propagating across the tank may
begin to present a problem. The significance of the transverse wave should be determined by
measuring its amplitude using a range of monochromatic wave tests. If these transverse waves are
found to be significant, it may be necessary to re-locate to a wider wave tank that has multiple wave
makers that can be used to absorb transverse waves. Alternatively, beaches can be used along the
sides of the wave tank. If these are used the wave tank will have to be re-calibrated to ensure that
they do not remove significant energy from the waves incident on the wave energy converter. Some
estimation of the influence of transverse waves could be achieved by assuming an infinite array of
devices is being modelled. However, this requires knowledge of the devices radiation pattern and
uses linear wave theory assumptions, which have limited validity in shallow water.
40
4.2.4 Wave-to-wire modelling
In the modelling of wave energy converters in wave tanks the results are often given purely in the
terms of hydrodynamic performance. This is understandable since a wave tank is used to examine
the hydrodynamics. However, whilst the details of the particular power take-off mechanism may
not affect the hydrodynamic performance (see section 4.2.1), the performance of the power take-off
mechanism may be affected by the constitution of the captured energy. For example, the efficiency
of the Wells turbine depends solely on the turbine flow coefficient; with a clear stall point at high
flow coefficients resulting in a rapid reduction in efficiency. Thus, to correctly model the average
performance of an OWC and Wells turbine the instantaneous flow rate must be used to determine
turbine efficiency and not the average flow rate. The configuration and associated performance of
the electrical generator, or other power take-off component, should also be included in the model
since these will also influence overall performance.
In addition to component efficiencies, the fixed losses of the power take-off mechanism should also
be considered. These can influence the overall sizing of the power take-off mechanism and thus the
overall design of the device. This has been made particularly acute for wave energy converters by
the high degree of variability in the wave climate, where for the majority of the time the device will
have to operate with relatively low power levels.
The wave-to-wire model, or more accurately the primary wave interface-to-wire model, does not
have to be extremely complex, however it is necessary that the method in which power is to be
collected has been defined and that sensible estimates of energy losses have been produced.
4.3 Specification of a standardised test procedure for shoreline and nearshore wave energy
converters
A standardised test procedure is required in order to provide comparative evaluation of alternative
devices. Thus, the test procedure should allow the performance of devices to be compared,
identifying the most promising for further study. To avoid investigation into areas without
significant promise, this procedure should be carried out in the earlier stages of development, and
thus it should be relatively simple and quick to perform. This implies that the number of tests
performed must be kept to a minimum.
Effective evaluation also means that the procedure chosen should be unbiased. Thus, the particular
tests chosen to estimate the annual average performance of the device should not favour one
particular design configuration above another. Unfortunately, the wave and/or site parameters that
define the performance of a particular device may not be known a priori and so it is unclear which
of these parameters are critical prior to testing. Moreover, extensive parameter studies are not
compatible with the requirement for a rapid evaluation.
This problem is resolved by defining a single set of tests for baseline performance, together with
additional tests that illustrate the sensitivity of the particular device to other parameters. These
sensitivity tests would examine parameters known to be potentially significant, together with any
other parameters considered important by the device team. These team-devised tests would help to
account for any unexpected biasing against the device in the baseline tests, allowing a more genuine
performance of the device to be demonstrated.
41
Though this procedure allows a relatively simple testing procedure to be used, which can account
for bias within the test set, it does not always permit a definitive comparative analysis to be
performed. That is, a device cannot always be declared as superior or inferior from the results of
this procedure. Consideration shows that this is reasonable to expect for two reasons. Firstly, during
the testing, abstractions of actual sea-states are used. If the method of abstraction is found to
influence performance, as demonstrated by any additional tests, the baseline test results must be re-
evaluated. Secondly, shorelines are not a homogeneous group, and it is possible that particular
device concepts are suited to particular locations. The procedure analyses an abstract situation and
so cannot be used definitively to compare devices.
4.3.1 Baseline tests for shoreline and nearshore wave energy converters
4.3.1.1 Productivity tests
The primary parameters for defining the incident wave climate have already been identified in
section 4.1.2.1. These are significant wave height and wave period at the 20m-depth contour and the
water depth at the device. For these baseline tests, it is assumed that the waves are long-crested and
have a Bretschneider spectral distribution. The following table contains the test sea states, taken
from a set containing 3 incident wave-power levels, 4 energy periods and 2 water depths.
Test number
Incident wave
power (kW/m)
Energy period
(secs)
Water depth
(metres)
1.1 15 7 5
1.2 15 9 5
1.3 15 11 5
1.4 40 9 5
1.5 40 11 5
1.6 40 13 5
1.7 65 11 5
1.8 65 13 5
1.9 15 7 7
1.10 15 9 7
1.11 15 11 7
1.12 40 9 7
1.13 40 11 7
1.14 40 13 7
1.15 65 11 7
1.16 65 13 7
Figure 35 Baseline productivity tests
The final column in the table, frequency, is an estimate of the percentage occurrence of the similar
sea states at the site of LIMPET on Islay. This can be used to provide an estimation of average
productivity if that is required. In general, such a simplification should be used sparingly since it
hides the relative performance of the device with respect to the three parameters.
Baseline productivity tests should last the equivalent of approximately 15 minutes at full scale.
Thus at 1:40 scale each test would last approximately 2 minutes.
42
4.3.1.2 Extreme loading tests
Traditionally, extreme sea conditions are defined by the significant wave height only, with the peak
wave period derived from the expression for a fully developed sea. This is used for these tests with
a single variation in water depth and spreading parameter to give four baseline tests for extreme
loading. These are shown in the table below.
Test number
Significant
wave height
(m)
Energy period
(secs)
Water depth
(metres)
Spreading
parameter
2.1 10 14 5 -
2.2 10 14 5 15
2.3 10 14 7 -
2.4 10 14 7 15
Figure 36 Baseline extreme loading tests
Each of these test sea-states should last for the equivalent of three hours at full scale. At 1:40 scale,
this is equivalent to a 20-minute test. These tests must be repeated for each possible state of the
wave energy converter, e.g. for a floating device, the testing should be repeated where one or more
of the moorings have failed. The test series should be designated as 2.1a, 2.2a, etc.
4.3.2 Additional tests for shoreline and nearshore wave energy converters
4.3.2.1 Productivity tests
For productivity testing additional tests are used to determine the sensitivity of particular
parameters as a measure of how much confidence can be applied to the baseline results (shown grey
in Figure 37). Test 1.14 (Pi = 40kW/m, Te = 11s, d = 7m) is used as a pivotal test to examine the
effect of wave period and water depth. Both of these, in different situations, have been found to be
critical to the performance of wave energy converters. Examination of these parameters in more
detail should help to determine whether the baseline tests are skewed by the particular values of
wave period and water depth adopted. These additional tests are shown in the table below, together
with the tests in the baseline set that form the series (though obviously these tests do not need to be
repeated).
Test number
Incident wave
power (kW/m)
Energy period
(secs)
Water depth
(metres)
3.1 40 7 7
3.2 40 8 7
1.12 40 9 7
3.3 40 10 7
1.13 40 11 7
3.4 40 12 7
1.14 40 13 7
Figure 37 Examination of wave period tests
43
Test number Test number Test number Test number
1.1 1.1 1.1 1.1
1.2 1.2 1.2 1.2
1.3 1.3 1.3 1.3
1.4 1.4 1.4 1.4
1.5 1.5 1.5 1.5
1.6 1.6 1.6 1.6
1.7 1.7 1.7 1.7
Figure 38 Examination of water depth tests
Whilst it is expected that the device performance will be significantly affected by both wave
frequency and water depth, additional tests are required to determine whether it has been reasonable
to assume that the influence of other factors is relatively insignificant. For this the effect of spectral
shape, dominant wave direction and spreading parameter should be tested.
Test
number
Incident wave
power
(kW/m)
Energy
period
(secs)
Additional parameter
Parameter
value
5.1 40 11 Spectral shape Twin-peaked
5.2 40 11 Spectral shape JONSWAP
5.3 40 11 Dominant wave direction +15
5.4 40 11 Dominant wave direction +30
5.5 40 11 Spreading parameter s = 8
5.6 40 11 Spreading parameter s = 15
Figure 39 Examination of other parameters
If the power capture is found to be particularly sensitive to any of these parameters, an estimate will
have to be made and justified regarding its influence on baseline performance. If necessary,
additional tests should be performed to provide sufficient information.
Additional tests may also be required to determine the influence on performance of other
parameters. These are more difficult to specify and identification of them will depend on the insight
and ingenuity of the device team. If an additional significant parameter is identified then sufficient
tests need to be performed so that its effect on the baseline performance can be estimated and
justified. These would form the test series 6.1, 6.2, etc
4.3.2.2 Extreme loading tests
Additional tests for extreme loading may be required where it is suspected that, due to the device
design the particular tests chosen, i.e. tests 2.1-2.4, will not present the worst case scenario. This
may occur where the device has a pronounced resonance that could be excited in particular seas
resulting in extreme excursions and subsequent loadings. Alternatively, the device may be
particularly susceptible to waves approaching from a particular direction. These tests should be
designated 7.1, 7.2, etc. with additional specification of device state as described in section 4.3.1.2,
e.g. 7.1a, 7.1b, etc.
44
4.4 Documentation of shoreline and nearshore wave energy converters testing
This standardised test procedure has been designed to support both comparative and developmental
analyses. This requires consistency in the application of the testing together with reporting of
results. A standard expectation is that the tests should be performed with sufficient rigour and
reported in sufficient detail so that the same results would be obtained if they were repeated by
different investigators, using only the report details as guidance. Thus, it is clear that the report must
include details of the experimental arrangement in addition to the results obtained.
4.4.1 Experimental set-up
Documentation of the experimental set-up needs to include physical dimensions and arrangements,
together with details on the instrumentation used and their calibration. Engineering drawings of
both the wave tank and model should be supplemented by photographs, and if possible video, of the
testing facilities. The relatively easy availability of digital cameras makes this a less arduous task
than it would have previously been. These photographs (and video if available) enable the
experimental set-up that is not detailed in engineering drawings to be critically examined and
reviewed.
Details of the wave tank calibration should be included so that the accuracy of the waves used for
testing can be confirmed. Details should include the reflection and shoaling coefficients for the
wave tank, together with the spectral shape of the test waves. The spectral shape should be
supplemented with wave gauge recordings of the surface profile at both the 20m-contour and the
device location. Where appropriate the directional distribution of the incident waves should be
recorded. The recording of the surface profile will enable other parameters of the incident wave to
be derived is necessary. To ease access to the data it should be saved in comma-delimited format,
with clear notes detailing the contents of the files. Ideally, this and all other data should be
contained on a CD that accompanies the report.
A description of the transducers used and their calibration should also be described and data sheets
included where appropriate. Finally, a description of any software (and its associated code) used to
manipulate and/or analyse the data collected should be included. In particular, details of any data
conditioning processes used to improve the quality of the data should be fully described.
4.4.2 Results section
Numerical results for all of the tests should be summarised in a table to provide easy access for
comparison and analysis. Each test should be cross-referenced to a data file containing the raw data
from the test, which is contained on the accompanying CD. An explicit description of any data file
referencing system should be included, together with how the data files are structured. Where
appropriate qualitative descriptions of tests should be included, focusing initially on the phenomena
observed and then providing conjecture regarding its effect on device operation and performance.
45
5 Task 4. Pressure Relief Model tests
5.1 Analysis of a LIMPET blow-off valve
5.1.1 Introduction
A relatively small amount of literature has been published regarding the use of blow-off valves with
Oscillating Water Columns (OWCs) (Falcao and Justino 1998; Alcorn 1999). Blow-off valves
have been proposed for use with OWCs to control the flow through its turbine and thereby increase
performance. This can be understood with reference to Figure 40, which shows the relationship
between the turbine flow coefficient and the turbine torque coefficient.
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5
0
0.005
0.01
0.015
0.02
0.025
0.03
0.035
0.04
0.045
0.05
flow coefficient, fi
t
o
r
q
u
e

c
o
e
f
f
i
c
i
e
n
t
,

T
*
Figure 40 Variation of turbine torque coefficient with turbine flow coefficient
Flow coefficient,
r A
Q
a

= , Torque coefficient,
5 2
*
r
T
T

=
Where, Q is the flow through the turbine (m
3
/s), A
a
is the turbine annulus area (m
2
), is the turbine
speed (rad/s), r is the turbine tip radius (m), is the air density (kg/m
3
) and T is the turbine torque
(Nm).
Figure 40 shows that to maximise turbine torque the amount of turbine flow needs to be limited to
avoid an excess flow coefficient and thus low turbine efficiency. For a Wells turbine the flow rate is
proportional to the pressure drop across the turbine and thus this condition is equivalent to limiting
the pressure in the plenum chamber. Thus, a perfect blow-off valve would operate so that the
chamber pressure was never higher than a critical value, P
ct
, defined by the turbine characteristics.
46
Falcao and Alcorn have both investigated the operation of perfect blow-off valves and shown,
unsurprisingly that the performance increases when compared to an OWC without a blow-off valve.
The amount of increase in performance depends on the time that the turbine would have
experienced excessive flow rates and their magnitude. Falcaos results indicate that in extreme cases
the system efficiency can be more than doubled.
In general, it is not possible to produce a perfect valve and thus it is important to examine the
influence on performance of a practicable valve. For manufacturing and control simplicity, a poppet
blow-off valve is proposed for the blow-off valve. A poppet valve can be in one of two distinct
states, open or closed. Characteristics of a poppet valve operation include the pressure at which the
valve opens, P
o
, the pressure at which the valve closes, Pc, the valve response time, t
v
, and the loss
coefficient of the open valve, C
l
(assuming that the closed valve does not leak).
The performance of a poppet blow-off valve has been investigated using both wave tank tests and a
numerical simulation. The wave tank tests used a 1:40
th
scale model of LIMPET in a narrow wave
tank. The wave tank tests were used to produce initial results on the performance of a blow-off
valve and to calibrate the numerical simulation. The numerical simulation was then used to
investigate the valve performance in a wider range of circumstances. The numerical simulation was
produced using Labview.
5.1.2 Wave tank simulation of the LIMPET blow-off valve
The model of a poppet-type LIMPET blow-off valve has been tested in the narrow wave tank at
QUB. The LIMPET model spans the full width of the narrow tank and thus is essentially a 2D
model. Such an arrangement will only provide relative changes in hydrodynamic performance due
to valve use, however it should accurately model the influence of the blow-off valve on pneumatic
performance.
An appropriately sized airbox was attached to the LIMPET plenum chamber to provide the correct
air spring stiffness. The Wells turbine was simulated using an orifice. The poppet valve was
modelled using a solenoid-activated plunger with side vents. The plunger is returned to the closed
position using a spring. Testing was performed with a total valve area of 1000mm
2
, which equates
to 1.6m
2
at full scale. The opening time for the valve is approximately 0.07 secs, which equates to
0.45 secs at full scale. The closing times were of a similar value.
The valve was controlled using two pressure values; a threshold opening pressure and a hysteretic
closing pressure. The valve would open at the threshold pressure and remain open until the pressure
dropped below the hysteretic pressure. The use of two controlling pressures was considered
necessary to avoid valve bounce. Valve bounce could occur as an oscillation between the drop in
pressure when the valve opened and the increase in pressure as it closed.
Data on device performance was obtained using four wave gauges measuring the water column
motion within LIMPET and a pressure transducer measuring the stagnation pressure (relative to
atmosphere) within the plenum chamber. The wave input power was calculated as the product of
chamber pressure and rate of volume change due to motion of the water column surface.
The orifice used to represent the turbine was calibrated using tests with the poppet valve
permanently closed. With the valve closed, it was assumed that the average input wave power was
equal to the average power loss across the orifice. Assuming that the pressure drop across the
47
orifice is proportional to the flow rate squared then an appropriate damping coefficient for the
orifice is given by
Orifice damping coefficient,
2 / 3
pressure
Power
B
o
=
Provided the same orifice is used, this expression can be used to calculate the duct power, which is
the power available to the turbine, when the blow-off valve is operating. Finally, the turbine shaft
power can be calculated using a flow coefficient derived from an equivalent pressure experienced
by the turbine, derived from the duct power, and the turbine speed.
5.1.3 Results of the wave tank simulation of the LIMPET blow-off valve
-500
-400
-300
-200
-100
0
100
200
300
400
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
Time (secs)
C
h
a
m
b
e
r

p
r
e
s
s
u
r
e

(
P
a
)
Valve closed P = 290 Pa P = 260 Pa P = 200 Pa P = 170 Pa P = 140 Pa
Figure 41 Effect of blow-off valve on chamber pressure for different control pressures
The effect of the blow-off valve on the chamber pressure and consequently turbine flow can be
clearly seen in
. This figure also illustrates the consequences of the finite time for the valve to open and close. For
example at an opening pressure of 200 Pa it takes until the pressure is approximately 270 Pa before
it has an effect on the chamber pressure and similarly the effect of valve closing at a closing
pressure of 200 Pa is only evident once the pressure has dropped to approximately 150 Pa. Because
of this effect, which eliminates the potential for valve bounce, all tests were performed with equal
values for the opening and closing pressures.
Clearly, a reduction in chamber pressure will result in a reduction in wave input power unless there
is an associated increase in flow rate, due to increased water column motion, that is sufficient to
compensate for it. The dynamics/hydrodynamics of LIMPET will depend on the wave frequency
and amplitude and well as the effective turbine damping coefficient and control strategy for the
blow-off valve. Figure 42 shows the effect of operating the blow-off valve on input wave power. In
all of the circumstances investigated, the effect of the blow-off valve was to reduce the input wave
48
power, with the amount of reduction differing in each case. This illustrates that the overall
performance of the blow-off valve will depend not only on its effect on the turbine flow, but also on
the water column hydrodynamics.
Opening/closing pressure (Pa)
Period
(secs) Amp (mm) 140 170 200 260 290
1.1 37.5 66% 70% 74% 80% 81%
1.6 37.5 82% 86% 87% 89% 89%
2.1 37.5 65% 67% 67% 65% 76%
1.1 50 64% 67% 71% 77% 78%
1.6 50 80% 84% 87% 86% 85%
2.1 50 60% 60% 60% 62% 60%
Figure 42 Effect of blow-off valve on input wave power as percentage of power with turbine
only operating
The effect of the blow-off valve on turbine performance will depend on the turbine characteristics
and turbine speed. These, together with the chamber pressure define the flow coefficient through the
turbine and thus the turbine efficiency. Using the LIMPET turbine characteristics (damping
coefficient and efficiency) produced by Curran and Gato, the average turbine efficiency is affected
as shown in Figure 43. These demonstrate that the average turbine efficiency can be significantly
improved by the use of a blow-off valve.
Opening/closing pressure (Pa)
Period
(secs) Amp (mm) 140 170 200 260 290
Turbine
only
1.1 37.5 68% 65% 63% 58% 57% 44%
1.6 37.5 67% 64% 62% 59% 57% 48%
2.1 37.5 55% 53% 51% 51% 47% 25%
1.1 50 59% 56% 53% 50% 49% 29%
1.6 50 59% 56% 53% 51% 50% 33%
2.1 50 54% 53% 53% 52% 51% 21%
Figure 43 Effect of blow-off valve on of LIMPET turbine efficiency
When the blow-off valve is operating not all of the energy will be available to the turbine; some of
the energy will go through the blow-off valve. The percentage of input wave energy to enter the
turbine duct is called the blow-off valve efficiency. Figure 44 details the blow-off valve efficiency
for the waves and valve opening/closing pressures tested. This clearly illustrates a significant
penalty of using a blow-off valve to improve turbine performance, with a large proportion of the
energy not reaching the turbine duct.
Opening/closing pressure (Pa)
Period
(secs) Amp (mm) 140 170 200 260 290
1.1 37.5 41% 45% 50% 58% 63%
1.6 37.5 52% 56% 59% 66% 69%
2.1 37.5 38% 40% 41% 44% 52%
1.1 50 39% 41% 45% 51% 53%
1.6 50 50% 53% 56% 58% 59%
2.1 50 36% 37% 38% 44% 45%
49
Figure 44 Effect of blow-off valve on input wave power entering turbine duct
To determine the overall effect of a blow-off valve these individual effects need to be combined
together into a single value. Figure 45 shows the percentage capture of input wave power, where the
input wave power is defined as the power available when the blow-off valve is not operating. This
shows that in the majority of cases the use of the blow-off valve has been detrimental to the overall
power capture. In only one case, where the initial efficiency of the turbine was particularly low does
operation of a blow-off valve increase the power captured.
Opening/closing pressure (Pa)
Period
(secs)
Amp
(mm) 140 170 200 260 290
Turbine
only
1.1 37.5 28% 30% 32% 34% 36% 44%
1.6 37.5 35% 36% 37% 39% 39% 48%
2.1 37.5 21% 21% 21% 22% 24% 25%
1.1 50 23% 23% 24% 25% 26% 29%
1.6 50 29% 29% 30% 30% 30% 33%
2.1 50 19% 20% 20% 23% 23% 21%
Figure 45 Overall change in power capture with the use of a blow-off valve
5.1.4 Discussion of Wave tank results
The use of a blow-off valve influences not only the performance of the turbine, by reducing stall,
but also changes the dynamics of the water column and the amount of captured energy going
through the turbine (the rest going through the blow-off valve). Although reduction of stall has
subsidiary benefits in the reduction of vibrations and noise, the primary objective is to improve
average energy capture, which means that any energy lost due to a change in water column
dynamics and through the valve must be more than compensated for by an increase in turbine
performance. Further consideration of these two factors clarifies the influence of a blow-off valve.
The use of a blow-off valve reduces the effective damping level applied to the water column when
compared to the turbine alone. To understand the influence of this on performance it is necessary to
understand how the turbine damping level has been determined. For any particular sea, there is an
optimal damping level that maximises the power capture; both higher and lower damping levels
reduce the power captured. To increase bandwidth the LIMPET turbine is typically described as
over-damped, i.e. applying a higher damping level than optimal, where the optimal damping is that
required to maximise power capture for an average monochromatic sea. Thus in mixed seas, the
actual level of damping applied is approximately optimal, with it being under-damped in some seas
and over-damped in other seas. Thus, the use of a blow-off valve, without any change to the turbine,
will result on average in the water column being under-damped and the water column capture-factor
reduced. To compensate for this the turbine damping coefficient when used with a blow-off valve
needs to be higher than when the turbine is used in isolation. Moreover, the optimal damping level
is generally higher for more energetic seas, which is when a blow-off valve would be used so that
the optimal design of the turbine requires a compromise between two distinct scenarios; a lower
damping level for when the turbine operates in isolation and a higher damping level for when the
blow-off valve is required.
50
The results of the wave tank experiments indicate that the level of turbine damping simulated,
which approximated the damping level of the LIMPET turbine, was too low for effective use with a
blow-off valve.
Optimal control of a proportional blow-off valve would result in a net improvement in turbine shaft
power for a given amount of pneumatic power. That is, the energy lost through the blow-off valve
is more than compensated by an improvement in turbine efficiency. However, an actual poppet-type
blow-off valve cannot guarantee this characteristic. The two-state nature of the valve means that in
general the optimal flow through the turbine cannot be achieved. Thus, the moment of valve
opening depends on a compromise between the flow coefficients being at times too high when the
valve is closed and possibly too low when the valve is open. This is further complicated by the
finite opening time of the valve so that the flow coefficient at the time the valve is fully open has to
be predicted.
The wave tank experiments have illustrated that for the valve used the loss in power going to the
turbine is generally not compensated for by a sufficient increase in turbine efficiency. In the one
case where the increase in turbine efficiency did produce a net increase in power captured the initial
turbine efficiency was approximately 20%. This suggests that overall the turbine has to be operating
for a significant amount of time in stall before the use of a blow-off valve is beneficial.
The wave tank experiments have only been able to examine one configuration of the blow-off
valve. Examination of other configurations, in particular different valve areas and control strategies,
have been performed using a numerical simulation.
51
5.1.5 Numerical simulation of the LIMPET blow-off valve
A time series simulation has been constructed using Labview. A linearised model of the LIMPET
OWC is used as shown where
B = hydrodynamic damping
K = water plane stiffness
f = wave force
M = water column mass (inc. added mass)
x = displacement of water column surface
y = time derivative of turbine and poppet valve flow
= air spring stiffness
= effective coefficient damping of turbine and poppet
valve
By using linear elements within the simulation, the equations of motion can be solved easily. A
more accurate model would use a convolution integral in place of the hydrodynamic damping and
added mass coefficients, together with a non-linear spring rate for the air compressibility and non-
linear coefficient for the poppet valve. Such a model would require significant development and
itself would rely on linear hydrodynamics. In shallow water, linear hydrodynamics has been found
to have poor accuracy and thus it is not clear that the additional complications would provide
significantly more accurate results.
The linearised system can be defined as a set of ordinary differential equations and then solved by
calculating the eigenvalues and eigenvectors of the coefficient matrix. Labview provides a built-in
function for this.
(
(
(
(

(
(
(
(


=
(
(
(
(

r
x
x
f
M M
k
M
B
M
dt
dr
dt
dx
dt
x d
dt
df
&
&

0 1 0
0 0 1 0
0 0 0 0
1
Where, for simplicity, y has been replaced with x y r = .
Exact coefficients of this linearised system are not available, however reasonable estimates can be
obtained by consideration of the underlying mechanics/hydrodynamics as contained in Appendix
5.A. These can then be further refined using the results of wave tank experiments.
1M
2
3
4 K
5 B
6
7
8 f
52
5.2 Results of the numerical simulation
It would be inappropriate to use such a crude numerical model for anything other than examining
general trends in performance as parameters are adjusted. However, these trends could vary
significantly over the design space and so the results of the wave tank experiments are used to
refine and verify that the initial parameters chosen are in the correct region. In particular, the
relationship between blow-off valve efficiency and turbine efficiency needs to be well matched, as
well as a tendency for the input wave power to reduce with operation of the blow-off valve. Finally,
the effect on chamber pressure should be similar in both the wave tank and numerical simulations.
This is achieved using the following coefficients
M = 15 kg
K = 800 N/m
= 960 N/m
B = 20 Ns/m
Turbine B
0
= 15 N
1/2
s
Valve B
0
= 50 N
1/2
s
Valve time delay = 0.06 secs
Nominal turbine speed = 47 (set to produce the correct flow coefficients)
Figure 46 shows energy accounting as the valve damping-coefficient is adjusted. Changing the
damping coefficient is equivalent to changing the size of the blow-off valve orifice. Figure 40
shows that above a flow coefficient of approximately 0.3 the turbine torque starts to decrease. In
this simulation, the valve is set to open and close at a flow coefficient of 0.15, allowing time for the
valve to open. Below a valve damping coefficient of 20 N
1/2
s, the valve experiences excessive
bounce and is used as the lower limit of damping. These results show that as the blow-off valve size
changes the combined effect is to keep the energy captured approximately constant.
0.0%
10.0%
20.0%
30.0%
40.0%
50.0%
60.0%
70.0%
80.0%
90.0%
100.0%
20 40 60 80 100 150 200 500 closed
Valve damping coefficient
P
e
r
c
e
n
t
a
g
e

e
n
e
r
g
y
Power captured Turbine losses Blow-off valve losses Wave input losses
Figure 46 Effect of valve damping coefficient on power capture
In the above simulation, the amount of wave power input reduced as the valve damping-coefficient
increased due to an effective reduction in column damping. By increasing the effective turbine
damping it is possible to reverse this effect. In addition, starting with higher average flow
coefficients will increase the potential for larger improvements in turbine efficiency. Figure 47
53
shows the effect if increasing the turbine damping coefficient to 50 and reducing the average flow
coefficient without the blow-off valve so that the average turbine efficiency is 26%. The simulation
shows both the increase in wave power capture due to the improved hydrodynamics, together with
larger increases in turbine efficiency. However, though the net effect of this is to demonstrate a
scenario where a blow-off valve would be beneficial, the overall conversion of wave energy
remains poor, peaking at only 32% of the input wave power available when the turbine is running
without a blow-off valve.
Further ad-hoc attempts to increase performance using a blow-off valve have generally been
unsuccessful in identifying a turbine and blow-off valve combination that shows an improvement
over the turbine alone, whilst giving average efficiencies of over 50%. These have included changes
to the opening time delay and opening/closing flow coefficients. It would appear that whilst a blow-
off valve can be used to improve performance, the improvement is only significant if the turbine is
far beyond its optimum operating conditions, i.e. the turbine experiences very high flow
coefficients.
0%
20%
40%
60%
80%
100%
120%
140%
160%
180%
200%
20 40 60 80 100 150 200 500 Closed
Valve damping coefficient
P
e
r
c
e
n
t
a
g
e

e
n
e
r
g
y
Power captured Turbine losses Blow-off valve losses Wave energy gained/lost
Figure 47 Effect of valve damping coefficient on power capture for over-damped turbine at
high initial flow coefficients
5.3 Conclusions
The use of a blow-off valve to improve the performance of LIMPET using both wave tank and
numerical simulations has been investigated. In particular, a poppet-type valve has been
investigated because of the envisaged cost and complexity of a proportional valve. Both simulations
have shown that in some circumstances the operation of a blow-off valve can improve the overall
performance of LIMPET. However, it would appear that this potential is only available if the
turbine experiences an average flow coefficient of over 0.3 (Wells turbines would typically be
designed to have an average flow coefficient of approximately 0.1). This is only likely to occur
during extreme sea states and the economic case for including the blow-off valve for these
circumstances would need to be carefully investigated.
54
The simulations have also shown that the turbine needs to be over-damped, relative to its optimum
damping in mixed seas, for the full benefits of the blow-off valve to be achieved. A blow-off valve
reduces the effective turbine damping just when the circumstances dictate that the damping should
be increased; the increased damping used to limit water column motion to reduce turbulent losses
and/or water ingestion into the turbine.
Given these requirements a control valve in series with the turbine appear more appropriate.
Moreover, this would avoid additional noise problems that could arise with the use of a blow-off
valve. An in-line control valve is not without its own problems in that the valve would need to be
proportional with the associated complexity. In addition, it would be necessary to ensure that the
valve does not have a detrimental influence on the flow through the turbine, which could adversely
affect the turbine performance.
It is also worth noting that these results have been achieved in monochromatic waves with the
optimum control identified using heuristic methods. Although a better control
configuration\strategy may have been missed, it is reasonable to assume that they represent a
realistic limit on performance. It is likely that in mixed seas the performance will be reduced,
making the blow-off valve even less attractive.
5.4 References
Alcorn, R. (1999). Turbine Modelling and Analysis using Data obtained from the Islay Wave Power
Plant. in ISOPE, Brest
Falcao, A. F. D. O. and Justino, P. A. P. (1998). OWC wave energy devices with air flow control.
Ocean Engineering 26 (1999): 1275-1295.
5.5 Appendix 5A Estimations of coefficients for numerical simulation
5.5.1 K water plane stiffness
The water plane stiffness can be simply calculated using the water plane area multiplied by the
water density and gravitational acceleration.
w w
gA K =
Where
w
is the density of water, g is gravitation acceleration and A
w
is the water plane area.
5.5.2 f wave force
The wave force is approximated as equal to the Froude-Krylov force.

w w
gA f =
Where is the wave amplitude.
5.5.3 B hydrodynamic damping
The hydrodynamic damping presents the greatest difficulty in defining an appropriate coefficient.
Using linear wave theory, the damping coefficient can be obtained from the expression
55
B
X
Power
8
2
= , where X is the wave force amplitude.
In the narrow wave tank the LIMPET model can be considered as a shoreline terminator and thus
the power is equal to the incident wave power, P
i
.
)
) 2 sinh(
2
1 (
2
2
2
1
kh
kh
k
g P
w i
+ =

where ) tanh(
2
kh k
g
=

Where k is the wave number, h is the water depth and is the wave frequency. The value of k
needs to be solved iteratively.
5.5.4 M water column mass
The mass of the water column is assumed equal to the total volume of water contained within the
water column.
wc w
V M =
Where V
wc
is the volume of the water column.
5.5.5 - air spring stiffness
Assuming the change in volume of the plenum chamber is small in comparison to its total volume
then the spring stiffness can be linearised
1
2

pc
w a
V
A p
Where is the ratio of specific heat capacities of air, p
a
is the atmospheric pressure and V
pc
is the
volume of the plenum chamber.
5.5.6 - effective damping coefficient of the turbine and poppet valve
The damping coefficient for a Wells turbine is constant (for a particular turbine and speed of
rotation). The damping coefficient of an orifice can be linearised using the instantaneous pressure
within the coefficient.
p B
o 0
= and ) ( y x p =
Where B
0
is the orifice damping coefficient and p is the plenum chamber pressure.
The combined damping coefficient of the turbine,
t
, and the poppet valve,
v
, can be calculated by
combining their coefficients in parallel.
v t
v t
+

=
This coefficient is calculated at each time step using the plenum chamber pressure from the
precious time step where necessary. This technique introduces a minimal error provided the time
56
step is sufficiently short in comparison to the rate of change of pressure; it is expected to be the
same order of magnitude as other errors introduced to enable linearisation.
57
6 Task 5 Visual Investigation of Collector Flow Characteristics
6.1 Introduction
The analysis of wave energy converters, including oscillating water columns (OWCs), has
traditionally been achieved utilising linear wave theory (Evans 1980). The benefits of using linear
wave theory are numerous and significant. In particular, linear wave theory has allowed exact
solutions to be produced with relative ease for both monochromatic and mixed seas, where the
superposition of linear monochromatic waves can be used. Moreover, linear wave theory has been
generally successful in predicting the hydrodynamics and performance of a wide range of marine
structures, from ships to wave energy converters (Newman 1977). The relative simplicity and
efficacy of linear wave theory has meant that few other procedures have been used in the design of
wave energy converters.
An area where linear wave theory has been considered inaccurate for wave energy converters is the
localised flow where vortices are produced. However, in general the forces generated by these
localised flows are small relative to the other hydrodynamic forces and it has been possible to
linearise them without significantly affecting the accuracy of the model (Brendmo, Falnes et al.
1996).
Using linear wave theory oscillating water columns have typically been represented as shown in
Figure 48.
9M
10
11K
12B
13
14F
- applied damping
M mass of water column (inc.
added mass)
B radiation damping
K water column spring stiffness
F incident wave force at
frequency ()
58
Figure 48: Linear representation of an OWC
Thus, linear wave theory allows the system dynamics to be represented in the frequency domain,
with the associated powerful tools of linear superposition and Fourier analysis. In addition, using
the frequency dependent hydrodynamic coefficients (radiation damping, water column spring
stiffness, added mass and incident wave force) the dimensions of the OWC can be optimised to
maximise power capture. Having defined the broad device dimensions, local geometries can be
adjusted to provide appropriate solutions to localised flow situations, e.g. increasing the size of the
entrance lip radius to minimise the losses due to vortex shedding, though remaining within the
general region of design defined by the linear analysis (Muller and Whittaker 1994).
Linear wave theory has provided a powerful tool for the design of wave energy converters, however
it relies upon the sinusoidal nature of ocean waves. In shallow water the waves can no longer be
considered as sinusoidal but have a cnoidal surface profile (Weigel 1960; Kinsman 1965). Figure
49 shows the difference between a sinusoidal and a cnoidal wave. The Ursell number, Ur, which
dictates the extent that waves can be considered cnoidal, rather than sinusoidal, is given by
Ur = h.L
2
/d
3
Where h is the wave height, L is the wavelength and d is the water depth. It is clear from this
equation that the water depth has a very significant effect on the non-linearity of the waves. In
addition, if the wave height is more than approximately 70% of the water depth then the waves will
break, with the type of breaker depending on the seabed slope (US Army Corp of Engineers 1977).
These broken waves similarly produce problems for a linear analysis.
0 2 4 6 8 10 12 14 16 18 20 22
5
5.5
6
6.5
7
7.5
Time (secs)
W
a
t
e
r

d
e
p
t
h

(
m
)
cnoidal wave
sinusoidal wave
Figure 49 Cnoidal and sinusoidal wave profiles
The significance of water depth to the performance of the LIMPET was first identified during wave
tank tests examining its power capture at a range of tidal positions (McStay 1995, QUB 1996).
Unfortunately, these tests did not provide a sufficient amount of information to explain the change
in performance and so a new set of experiments was devised to investigate this aspect.
6.1.1 Experimental set-up
A set of experiments was devised, using a 1/40
th
scale 2-D model of the LIMPET shoreline OWC,
to provide appropriate data to analyse the effect of shallow water on its hydrodynamics. A cross-
section of the LIMPET shoreline OWC is shown in
59
Figure 50: LIMPET shoreline OWC
The air volume of the plenum chamber was increased by the scale factor to provide a more accurate
representation of the effects of air compressibility (Alcorn 1999). The turbine was simulated using a
simple orifice with a cross-sectional area set to provide approximately the same amount of applied
damping to the water column motion. This has previously been shown to be a reasonable
approximation for the simulation of the plenum chamber and turbine (Folley 2001). To isolate the
influence of wave height and period the tests were performed using monochromatic waves using a
matrix of three wave periods and four wave amplitudes.
60
Figure 50: LIMPET shoreline OWC
Data was collected using a digital video camera, providing a spatial resolution of 720x596 pixels at
a frame rate of 25 frames per second. The physical frame size was 300mm x 200mm, providing a
resolution of 0.5mm (20mm FS). Whilst the narrow tank largely constrained transverse motion a
small amount of transverse oscillations was evident at times. This could reduce the resolution to
approximately 2mm (80mm FS), however these motions generally occurred with the larger wave
amplitudes, consequently the relative resolution remained adequate, and generally very high. The
frame rate equates to 31 frames per wave for the shortest wave period, or approximately 12 degrees
phase angle. This was considered an acceptable temporal resolution. The longer wave periods
clearly have a higher temporal resolution. To provide additional contrast the water was dyed using a
non-toxic red colorant.
The relative simplicity in the use of a digital video camera in comparison to PIV (particle image
velocimetry) or LDA (laser Doppler anemometry) makes this procedure well suited to ad-hoc and
exploratory experiments, where the objective is to determine the general characteristics of the flow,
as opposed to quantitative values of fluid velocity.
6.2 Identification and discussion of non-linear flow characteristics
For discussion of non-linear flow characteristics it is useful to consider the flow assumed by the
linear theory exemplified in
Figure 48. This linear analysis would expect a phase and amplitude difference between the motion
of the water surface inside and outside of the water column. Implicit in the theory and this
61
description is that the water contained within the column will move largely as a coherent body, with
all motions at the fundamental frequency of the incident wave. Thus, the flow characteristics are
discussed with reference to their divergence from this ideal. Examination of the video sequences
identified three major characteristics of the flow that would not be expected from linear wave
theory.
1. Front wall swash
2. Front wall down-wash
3. Water column slosh
This separation into three distinct flow characteristics is to some extent arbitrary, however the
separation facilitates the discussion of the characteristics and is used for this purpose only, as
opposed to any experimental or theoretical purpose.
6.2.1 Characteristic 1: Front wall swash
The front wall swash refers to the water that goes up the front wall on the outside of the device.
This is clearly illustrated in the video sequence shown in Figure 51 for a full-scale wave amplitude
of 1.5 metres and a wave period of 11 seconds (this wave is used for all of the video sequences).
Figure 51 a shows the crest of the wave arriving at the front wall of LIMPET. The water starts to
run-up the front wall of LIMPET (Figure 51 b), with the thickness of the water film on the front
wall deceasing with the distance it has progressed up the front wall (Figure 51 c). The final
maximum height of the run-up is greater than the incident wave amplitude (Figure 51 d). The
maximum height of the run-up and the time that it takes to reach this peak increases with both the
wave period and amplitude. The height and time to peak both appear to be related to the amplitude
of horizontal water particle velocity immediately in front of the front wall due to the incident wave;
the horizontal motion being converted into vertical motion by the slope of the front wall. Because of
this characteristic, the lower part of the front wall of LIMPET is steeper in order to reduce this run-
up.
(a) (b)
(c) (d)
Figure 51 Front Wall Slosh
6.2.2 Characteristic 2: Front wall down-wash
The front wall down-wash occurs as the water that has run-up the front wall falls back down. The
front wall down-wash starts after the finish of the swash as shown in Figure 52a. The water on the
front wall drops away (Figure 52b) and on re-entry into the sea pulls the water surface down (Figure
52c). The vigorousness of this down-wash entrains air into the water to form bubbles (Figure
52d,e). The combined effect of the down-wash waters velocity and the modified water surface
profile is to form a vortex in front of the device lip (Figure 52f) that then propagates away from the
device (Figure 52g). In addition, the down-wash causes an increase in localised hydraulic pressure
as some of its dynamic head is converted into static head. The effect of this can be seen as a small
increase in the water level next to the inside of the front wall (Figure 52g).
62
(a)
(b) (c)
(d)
(e)
(f)
(g)
Figure 52 Front wall swash
6.2.3 Characteristic 3: Water column slosh
The water column slosh can be seen in the video sequence shown in Figure 53. The run-up on the
back wall of the water column is similar to the water motion on the outside of the front wall and can
be seen as a gradient on the water column surface (Figure 53a). The steepness of this increases with
both the wave's amplitude and period and is related to the amplitude of the horizontal velocity. As
the water level in the column begins to fall (Figure 53b), the pressure pulse due to the front wall
down-wash occurs next to the inside of the front wall (Figure 53c). This small pulse can be seen as
a wave that propagates from the front to the back of the water column (Figure 53d). Simultaneously,
the gradient of the water column surface is reversed as the water beside the back wall pulls the
water surface down (Figure 53d). Following this run-up the water column pitches at approximately
its resonant frequency with the amplitude decaying as a result of damping and losses (Figure 53e,
6f). The variable geometry of the water column, which decreases the water plane area close to its
mouth, is designed to prevent the water column exiting the chamber and thus avoid water hammer
when the next wave rapidly closes the chamber. This has clearly been effective, as the water column
appears to pivot about a point on the front wall.
(a) (b)
(c)
(d)
63
(e) (f)
Figure 53 Front wall swash
Whilst these characteristics have been identified using monochromatic waves in a narrow 2-D wave
tank, observations from the LIMPET wave power plant sited on the Isle of Islay have generally
been in accord with them. An example of the front wall swash and down-wash is shown in Figure
54. This implies that the characteristics are neither dependent on the two dimensionality of the tests,
nor on scaling effects.
Figure 54 Front Wall Slosh
64
Figure 55:Front wall down-wash
65
The conservation of energy is a universal principle that could form the basis of an
additional tool in the analysis of wave energy converters. This principle can be
applied both temporally and spatially and is equally valid whether the motion is linear
or non-linear. To make use of this extremely powerful principle an understanding of
what influences energy flux and conversion is required. That is, what makes the
energy move from the waves into the front wall swash, water column motion, etc.
For the purposes of analysing the primary energy conversion of wave energy
converters, the energy can be considered as being one of two types, kinetic or
gravitational potential. Thus, chemical energy, electrical energy, pressure energy, and
thermal energy are all ignored. Sensibly, there is only a single source of energy, i.e.
the incoming waves, and we assume that the only possible energy sinks are the power
take-off mechanism, outgoing waves and energy lost due to viscosity. Energy
conversion and flux will depend on the localised pressures and velocities within the
fluid. In addition to these energy sources/sinks there will be one or more energy
stores.
As an example consider an idealised simple symmetrical deep-water 2-D heaving
terminator, where there is a single energy source of the incoming waves, the energy
sinks are outgoing waves, the power take-off mechanism and viscous losses in the
form of turbulence due to vortex shedding. In addition, there is an energy store in the
motion and excursion of the terminator, together with the surrounding water. Its
energy model is illustrated in Figure 56. If necessary this model can be further refined
by separating, where appropriate, the energy into kinetic and potential energies,
and/or different regions of the surrounding water. Though this model is accurate in its
own terms it is not definitive and alternative models could be drawn where this
boundaries are defined differently; the choice of model depending on the purpose for
which it is designed.
Figure 56: Energy model for 2D heaving terminator
The tendency for energy to move from the incoming waves to the energy stores or the
energy stores to the energy sinks depends on the hydrodynamics/dynamics of the
device. Using linear theory these relationships can be solved relatively easily. In
general, a well-known condition for optimum power capture is that the terminators
resonant frequency is the same as the incident wave frequency and that the applied
damping is equal to the effective damping due to the outgoing waves and viscous
losses (Evans 1985). With reference to energy these conditions ensure that energy
Incoming
waves
Outgoing
waves
Power
take-off
Viscous
losses
Terminator motion
Motion of
surrounding water
66
flows easily from the incoming waves into the terminator motion and that a minimum
amount of this energy then flows back out into the surrounding water.
In this idealised case there appear to be no significant advantages to the use of an
energy model, however consider a case where the terminator experiences induced roll
due to the operation of its power take-off mechanism. In the energy model the effect
of roll would be represented as an additional energy sink. The significance of motion
would then depend upon the amount of energy required to produce and sustain the
roll, as opposed to the amplitude of the motion itself (provided of course that the
motion is sufficiently small to not affect the primary interaction of the hull with the
waves).
Development of an energy model for an actual device requires identification of the
potential energy stores and sinks and mechanisms that define the propensity of energy
to flow from one element to another. The production of such a model is not
necessarily an easy exercise and may require a number of refinements before a
suitable model is identified. The video analysis of LIMPET, together with the linear
model has been used to produce the energy model for LIMPET shown in Figure 57.
This model has three energy stores into which the incoming wave energy can go,
together with the potential for the incoming wave energy to be simply reflected back
away from the device, which is represented as a direct path between the incoming and
outgoing waves.
Figure 57: Energy model for LIMPET shoreline OWC
LIMPET is sited in relatively shallow water, which causes the incoming wave energy
flux to have two particular characteristics. Firstly, the energy flux is not uniform, with
the majority of the energy arriving with the crest of the wave and a lower energy flux
during the trough. Secondly, a large proportion of the energy in the wave is kinetic
energy associated with the horizontal motion of the water. These two characteristics
are valid for both cnoidal and broken waves, with the broken waves representing a
more extreme situation.
From Figure 57, the energy in the incoming wave can have four destinations; into
water column heave, water column slosh, front wall swash and as a reflected outgoing
Incoming
waves
Outgoing
waves
Power
take-off
Viscous
losses
Water column
h
Front wall
swash/down-rush
Water column
l h
67
wave. For LIMPET to work effectively the majority of this energy needs to go into
the water column heave since this is the only destination that is coupled directly to the
power take-off mechanism. However, the tendency for the incoming wave energy to
go into water column heave depends on its relative impedance to the wave energy, in
comparison to the other energy stores/sinks. This can be determined using an analysis
of the hydrodynamics/dynamics of LIMPET. Though a detailed CFD model would be
able to provide this information, the video data provides sufficient information to
produce a reasonable analysis.
The front wall swash is evidence that a large amount of the incoming wave energy is
being diverted into this energy store. The slope on the front wall provides a relatively
low resistance path to convert the kinetic energy in the horizontal water velocity into
the swash. Incoming wave energy that passes the front wall enters the water column.
The back wall slope helps to convert the waters kinetic energy into water column
heave, together with some run-up the back wall that becomes the slosh in the water
column. The exact proportions of energy in heave and slosh of the water column
depend on the back wall slope and percentage the incoming waves kinetic energy
associated with the waters horizontal motion (the gravitational potential energy being
more easily converted into heave). Depending on the combined resistance of these
three energy stores, some incoming wave energy will be reflected directly away from
the device.
The energy stored in the front wall swash is transmitted to other energy stores/sinks
with the front wall down-rush. Much of this energy is converted into viscous losses,
via the production of a significant vortex at the bottom of the front wall. An outgoing
wave is also produced, though it does not have the same frequency as the incoming
wave. Finally, some energy is transmitted to the water column, where it may induces
some heave and slosh. Unfortunately, the pressure pulse from the front wall down-
rush generally opposes the water column heave so that energy is rarely transmitted to
this energy store.
Energy in the water column slosh is transmitted either to outgoing waves, generally at
the resonant frequency of the slosh, or dissipated as a viscous loss. The energy in the
water column heave can be converted into useful power via the power take-off
mechanism, though some will also be radiated in outgoing waves due to its motion
and some dissipated as viscous losses. In many cases, nearly all of the energy in the
water column motion had dissipated prior to the subsequent wave crest that carries the
majority of the energy for each wave. This creates difficulties in producing large
water column motions.
These general mechanisms were common to all of the waves tested; however, their
relative significance varied depending on the incident wave period and amplitude. The
amount of energy that goes into the front wall swash increased with both wave period
and amplitude. This is due to the higher non-linearity of the incident wave defined by
the Ursell number. This higher non-linearity results in both a larger proportion of the
energy arriving in a shorter proportion of the wave, and an increase in the proportion
of the incoming wave energy associated with horizontal motion of the water particles.
The amount of water column slosh also increased with wave period and amplitude for
the same reason.
68
It would be expected that the amount of water column heave would increase as the
incoming wave frequency approached the water columns natural frequency in heave.
This would be due to the energy stored in the water column heave cycling between
kinetic and potential energy, so that the motion of the water column would provide a
low resistance to the acceptance of energy. Unfortunately, the other wave induced
motions in LIMPET obscure this effect, indicating that even at the water column
heave resonant frequency the incoming wave energy does not convert easily into
water column heave.
6.3 Conclusions and Implications for Shoreline OWCs
The results of this video analysis can be usefully compared to data from a previous set
of flow visualisation tests for a shoreline OWC (Muller and Whittaker 1994). These
tests showed a surging breaking wave forming when the back wall of the chamber is
vertical. The use of the sloped back wall in LIMPET eliminated this destructive wave.
Previous wave tank tests have also shown that a sloped back wall can be beneficial to
the performance of a shoreline OWC (McStay 1995), though these tests were
performed in slightly deeper water. Optimisation of the back wall slope of a shoreline
OWC requires a compromise between a gentle slope that eliminates destructive
internal waves and helps to convert horizontal water particle motion into vertical
motion and a steeper slope that reduces the induced pitching of the water column due
to the run-up of the water on the back wall.
The above analysis of the energy system for LIMPET has enabled some conclusions
to be drawn regarding the design of shoreline OWCs. Of key significance is that in
shallow water much of the incoming wave energy is in the form of kinetic energy
associated with the horizontal motion of the fluid particles. An effective design of an
OWC requires that this kinetic energy be converted into a change in the water column
heave with a minimum amount of energy being diverted into other motions. It would
seem that this requirement is more difficult to satisfy in shallow water than deeper
water, so that the design of an OWC in shallow water depends critically on the exact
water depth and the non-linearity of the waves. Some additional studies into the
performance of LIMPET have indicated that the performance is influenced by the
seabed profile, in addition to the water depth (Folley 2001b). It has been suggested
that the Ursell number provides a good indication of a waves non-linearity due to
water depth. The Ursell number is inversely proportional to the water depth cubed
indicating that only a small variation in water depth could have a significant effect on
the device performance.
The above analysis illustrates that productive interpretations can be obtained using the
relatively simple procedure of digital video analysis. The low cost of digital videos,
together with their portability and the absence of health and safety restrictions that
apply to laser-based PIV and LDA systems, makes them ideally suited to ad-hoc and
exploratory analysis.
The data obtained from the video sequences has been used in conjunction with an
energy model of LIMPET to provide useful insight into its performance. The use of
an energy model has enabled the non-linear effects of shallow water to be clearly
identified. This would have been difficult, if not impossible, to achieve using standard
69
linear analysis. Of course, a detailed non-linear CFD model may have been able to
provide the same results, though it would not have provided the insight that the
energy model was able to achieve. The relative simplicity of an energy model enables
it to be used for interpretative as well as predictive purposes. The efficacy of using an
energy model for LIMPET implies that it may also be productive to produce models
for other wave energy converters, especially in shallow water or where there are
significant non-linearities.
6.4 References
Alcorn, R. (1999) Turbine Modelling and Analysis using Data obtained from the Islay
Wave Power Plant. in ISOPE 1999, Brest
Brendmo, A., Falnes, J., et al. (1996) Linear Modelling of Oscillating Water
Columns including Viscous Losses. Applied Ocean Research 18: 65-75.
US Army Corp of Engineers (1977) Shore Protection Manual. Coastal Engineering
Research Centre.
Evans, D. V. (1980) Some Analytic Results for Two and Three Dimensional Wave-
energy Absorbers. in Power from Sea Waves, Ed. B. Count .
Evans, D. V. (1985) The Hydrodynamic Efficiency of Wave-Energy Devices. in
Hydrodynamics of Ocean Wave-Energy Utilization, Lisbon, Ed. D. V. Evans and A.
F. de O. Falcao Springer-Verlag.
Folley, M. S. (2001) Modelling the Wells turbine and plenum chamber,
QUB/MSF002, Internal Report, Queens University Belfast
Folley, M. S. (2001b) Initial LIMPET trials examining the effect of bathymetry,
QUB/MSF001, Internal Report, Queens University Belfast
Kinsman, B. (1965) Wind waves. Englewood Cliffs, NJ, Prentice-Hall.
QUB (1996) European Wave Energy Pilot Plant on Islay (UK)
McStay (1995) An Experimental Study of Cylindrical Wave Power Devices, MSc.
Thesis, Queens University Belfast
Muller, G. and Whittaker, T. J. T. (1994) Visualisation of Flow Conditions inside a
Shoreline Wave Power-Station. Ocean Engineering 22(6): 629-641.
Newman, J. N. (1977) The Interaction of Stationary Vessels with Regular Waves. in
11th Symp. Naval Hydrodynamics, London
Weigel, R. L. (1960) A presentation of cnoidal wave theory for practical
application. Journal of Fluid Mechanics 7: 273-286.
70
7 Task 6 Investigate Model Loads
7.1 Introduction
A key driver in the cost of wave generated power is the capital cost of the equipment.
For a shoreline OWC a large proportion of this cost, typically 60-75%, is associated
with the structure and the structure cost is to a large extent determined by the
environmental loads to which it is designed. In the design of LIMPET the major
application of environmental load derived from two sources; external wave loads and
internal pressure forces resulting from the water column action. These were
considered independently and the design made proof against such independent
application. That the design is successful is evidenced by its survival in extreme
conditions.
Observations made during LIMPET model tests led to the conjecture that peak
external wave forces on pushing the outer wall of LIMPET inwards were not
independent of the internal pressure forces and that if in fact these forces were
concurrent then the design of LIMPET might well be overly conservative to the point
where the structure was unnecessarily expensive. The object of this work task was
therefore to perform a brief series of tests to investigate the net loads on the front wall
of the collector and to flag the implications for future designs.
7.2 Model testing
The model use in load testing was equivalent to a single cell LIMPET with a 10m
cell width. A straight roof was used and this is shown superimposed on the original
LIMPET profile in Figure 56.
Figure 56 Load Model Roof Quartiles
For convenience the roof was split into quartiles and a direct measurement was made
of the loads on the lower three sections. This was achieved by the use of duct tape to
loosely bond the roof sections to the model structure whilst providing an adequate air
71
seal against the low pressure air in the water column. Motion of the roof sections was
then restrained by a frame mounted load cell fixed normal to the wall surface, (Figure
57).
Figure 57
The load on the upper section was derived from other readings and may be subject to
significant error. A typical set of load traces taken in extreme wave conditions with
the device in a working condition is shown in
Roof loads wit h lower collector split into 3 sections
-2.5
-2
-1.5
-1
-0.5
0
0.5
1
1.5
2
2.5
3
49 54 59 64 69 74 79
T ime
K
g
-600
-100
400
900
1400
1900
2400
cell 8 kg
cell 9 kg
cell 10 kg
ai r pressure Pa
water level mm
Figure 58.
72
Roof loads wit h lower collector split into 3 sections
-2.5
-2
-1.5
-1
-0.5
0
0.5
1
1.5
2
2.5
3
49 54 59 64 69 74 79
T ime
K
g
-600
-100
400
900
1400
1900
2400
cell 8 kg
cell 9 kg
cell 10 kg
ai r pressure Pa
water level mm
Figure 58
In
Roof loads wit h lower collector split into 3 sections
-2.5
-2
-1.5
-1
-0.5
0
0.5
1
1.5
2
2.5
3
49 54 59 64 69 74 79
T ime
K
g
-600
-100
400
900
1400
1900
2400
cell 8 kg
cell 9 kg
cell 10 kg
ai r pressure Pa
water level mm
Figure 58 traces labelled cells 8,9 & 10 refer to the loads on quartiles 1,2,&3 of the
roof. These traces show a number of significant features.
As is typical in an OWC the peaks in internal air pressure are 90
o
out of phase
with the water column position.
The peak loads on the three lowest quartiles do not occur at the same time.
The maximum outwards load on sections 2 and 3 are generally in phase with the
internal air pressure.
The maximum outward load on the lowest section (1) is generally in phase with
the minimum internal air pressure and 180
o
out of phase with the load on the other
sections.
73
The results indicate that when the water column is rising, so that the internal air is
pressurised, the internal air pressure dominates the load on the roof in sections where
there is limited external water coverage. In this case the roof load and internal
pressure are in phase. Towards the bottom however where the incoming wave creates
a large hydrostatic pressure on the roof the internal pressure is lower than the external
head and as such the net force is inwards. As the wave recedes the external head will
vanish whilst the water inside the column will now generate an internal hydrostatic
head which is greatest at section 1 and is acting outwards. At the top of the roof
(sections 3 & 4) there is no internal hydrostatic head and so no outwards force. The
air in the water column is however rarefied by the falling water so that with
atmospheric pressure acting outside the roof there is a net inward force. It is thus seen
that the combination of internal pressure and external load serves to produce a load
cycle which is quite different at the top and bottom of the water column. This is in
marked contrast to the independent treatment of internal and external loading in the
original basis of design for LIMPET.
Figure 59. Equivalent Max/Min Roof Pressures
The peak internal and external forces measured for each quartile have been converted
to an equivalent full scale roof pressure by dividing the measured load by the area of
the roof quartile. It is seen that whilst the difference between the maximum and
minimum pressures remains substantially constant at around 30kPa, the mid-point
varies significantly up the roof. Thus at the top of the water column the maximum
outward load is equivalent to 26kPa whilst at the bottom the maximum equivalent
pressure is only 13kPa. For comparison the peak internal pressure measured during
the tests was in excess of 90kPa.
Similarly the maximum equivalent external pressure was 18kPa at the bottom of the
roof but on 6kPa at the top section. This compares to a maximum wide area pressure
assumed in the basis of design of 400kPa.
-30
-20
-10
0
10
20
30
1 2 3 4
Roof Quartile
M
a
x

e
q
u
i
v
a
l
e
n
t

p
r
e
s
s
u
r
e
(
k
P
a
)
74
It is thus seen that if the external wave forces and internal pressure forces are
considered in combination rather than independently then the design loads fall
significantly. In relation to the LIMPET basis of design internal pressure loads reduce
by a factor of three whilst external wave forces reduce by over 20. This offers the
potential for a dramatic reduction in design complexity and cost.
7.3 Internal Wave Slam
In addition to the measurement of global front wall loads a series of 1/40
th
scale
model experiments to determine the amplitude of wave loads on the internal back wall
of LIMPET were performed using the narrow wave tank at Queens University
Belfast. Wave loading was measured using three submersible absolute pressure
transducers installed flush with the surface of the back wall. All three transducers
were installed along the vertical centreline of the back wall, at the still water level and
50mm (equivalent to 2m at full scale) both above and below this level. The
transducers were installed around the still water level because for sloped seawalls the
maximum wave loading is typically found at this position. The clear Perspex model of
LIMPET also allowed to the motion of the water column to be observed, permitting
the location of any wave breaking to be identified.
The pressures measured by the transducers were recorded for a range of different
monochromatic wave periods and amplitudes typical for the LIMPET site. In
addition, a number of seas with Pierson-Moskowitz spectra were used for testing.
Within the limitations of 2D wave tank testing, these produced excitation of the water
column over a representation range of conditions. Data from the pressure transducers
was collected at 40 kHz. This enabled any particularly transient event to be fully
captured. This is necessary in the measurement of wave pressures, where the
maximum pressure can occur in a very short spike due to compression of air bubbles
contained or entrained by the water.
In all of the model tests, no wave breaking was observed that impacted onto the back
wall of the water column. Similarly, the video of the water column motion obtained
from the full scale LIMPET has shown no waves breaking onto the back wall. In
accord with this observation, no pressures that would be associated with wave
breaking were recorded by the pressure transducers. The maximum pressures
measured were of a similar magnitude to the maximum air chamber pressures, i.e.
approximately 500 Pa (equivalent to 20 kPa at full scale).
The occurrence of wave breaking onto the back wall of the water column had been a
particular problem in the 75 kW OWC prototype built by Queens University Belfast.
It would appear that the use of an inclined back wall has successfully eliminated this
potentially significant problem.
75
8 Task 7 Conceptual Evaluation of New Device
During the construction of Limpet and the subsequent operation of the plant it became
clear that there remain conceptual problems to be solved before the technology can be
generally applied. Amongst the key findings were that:
The construction in an excavation behind a rock wall did not permit free working
in all weather conditions.
Any reduction in work at the waters edge and in particular sub-surface will greatly
enhance the economic prospects of the collector.
In shallow water the setting back of the collector in a straight gully is detrimental
to the performance of the device relative to the same unit at the cliff edge.
The design loads used for LIMPET significantly overestimate the environmental
forces, (see section 8)..
Wash back of rock into the collector mouth can significantly affect the potential
performance of the device.
Consideration has been given to these factors and an alternative approach to the
construction of shoreline collectors has been developed. To appreciate this
development it is first necessary to review the development of the LIMPET form.
8.1 Development of the LIMPET Cross Section
Figure 60. LIMPET Cross Section
The width of the LIMPET collector was determined by estimating the width of wave
front necessary to give the plant a rated output of 500kw, this value being assessed as
76
a suitable size of device to make a significant contribution to the Islay grid and being
a worthwhile step forward from the 75kW prototype. Having established this width at
21m it was, for two reasons, necessary to divide the water plane into three separate
columns:
As the width of the column increases there is an increasing risk of transverse wave
excitation within the water column. This reduces the energy capture performance
of the column. Whilst the 6m width of the prototype device is known to perform
satisfactorily there was concern that a significant increase above this might cross
the limit of acceptability.
The depth of roof required to span the 21m width of the column without additional
support was considered too large to be economically efficient as an in situ casting,
the originally preferred method of construction.
The three chamber design requires two dividing walls. It then becomes necessary to
determine how the walls are to be held to the base rock against the internal chamber
pressure. There were two clear choices; either the walls could be fixed directly to the
excavated slope with rock anchors, or a rear wall could be cast on the excavated slope
so that the cast structure formed a closed circuit in terms of load containment. For a
number of reasons the latter option was selected. The roof is subject to downward
loads from external wave action and upward forces from the internal pressure
generated by the OWC action. Model tests have indicated an internal design pressure
of 1bar which translates into a linear load on the walls of approximately 450kN/m.
Whilst this figure does not take account of the weight of the structure there remains a
substantial anchor requirement. This coupled with the fact that the quality of the
surface to which the walls would be anchored would not be known until after the
excavation was complete and that a rough surface on the rear water column could
detract from the column performance reduced the attraction of direct fixing. The role
of the LIMPET as a research tool again weighed heavily in the thinking and despite
the likely cost penalty the closed option as shown in Figure 60 was selected.
If the design philosophy is revisited then it is possible to draw a different conclusion.
For example let us assume that the design pressures described in section 9 are
applicable and that a 10m wide device will not suffer significant loss from transverse
resonance.
The immediate impact of the first assumption is that a concrete beam of
approximately 1.2m thick on the lower sections of the collector will be heavy enough
to stay in place against the peak uplift forces so that in principle there is no need to fix
the roof in place beneath the water surface. Above the water surface where the uplift
forces are higher there is a remote danger of uplift and it is prudent to assume a small
degree of rock anchoring in this area. Critically however there is no requirement in
either case for there to be any vertical walls to tie the roof to the rear wall of the
collector chamber. This means than not only can the vertical walls be eliminated from
the design but also the rear wall can be cast against the sloping rock profile. If the
base rock is used to form the sides and rear of the collector section then, firstly the
volume of concrete required from the collector section is reduced to approximately
one third of that used in LIMPET, and secondly sub surface working is almost totally
eliminated. A methodology for forming this type of collector is currently being
developed.
77
8.2 Capture Performance of 10m wide Cliff Mounted Device
-0.20
0.00
0.20
0.40
0.60
0.80
1.00
1.20
1.40
1.60
1.80
2.00
5 6 7 8 9
1
0
1
1
1
2
1
3
1
4
1
5
1
6
1
7
1
8
1BD
2BD
3BD
4BD
5BD
6BD
7BD
0BD
Figure 61 Performance curves for 10m wide ROLIM
To assess the likely performance of a 10m wide device mounted at the cliff edge a
brief series of tests were made to measure capture performance. The results are shown
in Figure 61 . The different curves relate to different numbers of test damper units
(BD) applied in series. It is seen that the model, which is not optimised for the site
conditions, exhibits a high capture over a broad bandwidth. No significant internal
transverse waves were observed.
8.3 Discussion and Conclusions
By reviewing the experience of constructing LIMPET in the light of experience and
model tests for both performance and structural load it has been demonstrated that in
principle it is more effective both in energy capture and in efficacy of construction to
reduce the unit width from that of LIMPET. Whilst attention has focussed on a 10m
wide collector this is not necessarily the optimum.
It has also been shown that the 10m wide unit mounted at the cliff edge gives a
significantly higher capture performance than the LIMPET set in the gully.
For the proposed construction technique to be available the site needs to have deep
enough water to prevent any blast debris washing into the mouth of the gully whilst
having a suitable cliff height and rock quality to allow the formation of the rock
walled collector. A number of sites have been examined in the UK and elsewhere.
The sites identified in the UK were remote from useful grid connections and for that
reason the current focus of development of the next generation of shoreline collector
is outwith the UK with a detailed feasibility study currently in process for the
favoured location.
[new here end]
78
9 Task 8. Installation of Power Dumping Equipment.
After commissioning the plant and observing the lower than expected generation it
became clear that the power dumping equipment would not be necessary and on those
occasions when excess power was generated it could be controlled by modulating the
in line valves. In consequence the supply and installation of the power dumping
equipment was deleted from the scope of work for this project and the budget released
reallocated to new tasks 13-15.
10 Task 9. Develop Additional Control Strategies
10.1 Introduction
Whilst LIMPET is a grid connected power producer it is also a research tool. A major
consideration in the selection of inverter drives for the generators was the facility it
gave for changing device operating parameters. By adjusting the inverter controls the
speed of the turbines can be adjusted over a wide range with the effect of altering both
the turbine efficiency, the water column damping and the system losses. The interplay
between these items is complex and in the time domain, not well understood.
Operational data is gathered from the plant for 15 minutes every hour and statistical
analysis of this data is being used to assess the impact of changes to the control
strategy. Because of the irregularity of the input power to the device it is often
difficult to assess whether minor changes to the strategy are beneficial and for this
reason a cautious approach has been adopted in adjusting the control software.
Nevertheless the changes made since the plant commissioning in late 2000 are bearing
fruit.
10.2 Baseline Control Strategy
LIMPET is controlled via an industrial computer programmed in C to fulfil the
following core functions:
1) Determine whether it is safe and desirable to operate the plant.
2) Start up the equipment.
3) Control the generation and monitor the operation of the plant, instituting an
appropriate shutdown in the event of problem.
10.2.1 Pre-start Check
Before allowing the plant to start the software performs a number of checks including:
Emergency stop circuit closed.
No warnings from any monitoring equipment
Adequate energy entering the water column
If these pre-start checks are positive then the plant will commence its start routine.
79
10.2.2 Start up
The start routine includes the following:
Operation of the vane valve to check function
Operation of the butterfly valve to check function
Start generator 1 and motor to a set speed. Generator 1 then enters production.
Start generator 2 and motor to a set speed. Generator 2 then enters production.
Enter Power Production Mode
10.2.3 Power production
Once in power production the control system enters a monitoring and checking phase.
Valves are periodically cycled to check function and the unit will be shut down either
in the event of a fault condition or if the input wave energy falls below a minimum
level. Fault conditions include:
Excessive turbine speed
High bearing temperature
Excessive bearing vibration
G59 Fault on the grid (over/under voltage/frequency, phase imbalance etc)
The controller software also provides control signals to the two valves and to the two
generators. The generators are inverter driven via a torque demand signal from the
control unit. The control algorithms are written in C by Wavegen. There are two
key algorithms, one controlling the generator speed and the second the valve
positions. The baseline function of these algorithms may be explained with regard to
Figure 62.
0
20
40
60
80
100
500 700 900 1100
Speed (rpm)
Demand Torque Valve Position
Figure 62
80
10.2.4 Generator Speed Control
There is a large rotating mass associated with each of the two turbines (1250kg.m
2
per
unit) and as the energy input varies through the wave cycle power is either fed into or
is drawn from the inertia in order to smooth the power supply to the grid. This is
achieved by varying the torque reference signal to the inverters. The demand torque is
determined as follows:
If the turbine speed falls below a set minimum the demand torque is zero. This
prevents the turbine falling to a low speed from which it cannot absorb
sufficient power to recover.
When the turbine speed is above the minimum speed but below a second set
speed (action speed) the demand torque varies linearly from zero to the
maximum available (the maximum available being determined either from
grid or generator limitations).
When the turbine speed is above the action speed the maximum available
torque is drawn.
A separate but identical algorithm controls each generator.
10.2.5 Valve Position Control
The function of the two in line valves (vane and butterfly) is to reduce airflow to the
turbines in storm conditions and to close in an emergency. The butterfly valve is held
fully open in normal operation but is fully closed in the event of a shutdown. The
vane valve modulates during operation but closes in the event of a shutdown. The
position of the vane valve is determined as follows:
If the turbine is running beneath a first set speed the valve is fully open.
Between the first and second set speeds the valve closes linearly to zero.
10.3 Limitation of Baseline Controls
Figure 63. Contra-rotating Turbine Efficiency (after Curran)
Contra-rotating Biplane
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4
Flow Coefficient
E
f
f
i
c
i
e
n
c
y
81
Turbine efficiency is dependent upon flow coefficient (F), which is defined as the
ratio of axial flow velocity to turbine tip speed (Figure 63). Throughout the wave
cycle the axial flow velocity is varying between zero and maximum and the flow
coefficient tracks in sympathy. There is relatively little energy involved when the
flow coefficient is low so that inefficiency in this area does not dramatically affect the
overall performance. Since the energy associated with the flow increases with F
2
,
poor efficiency at high values of F are by contrast very damaging. It is important
therefore that the turbine speed is matched to the axial flow so that the flow
coefficient is typically reaching a maximum of little more than that required for peak
efficiency; i.e. around 0.2 for the curve of Figure 63. The problem arises that as the
wave conditions change to supply more power to the turbine, the required turbine
speed to limit flow coefficient also increases. The relationship between demand
torque and speed in the baseline control strategy addresses this problem but there is no
immediate way of knowing whether the parameter values chosen for the control
speeds are correct. The situation is further complicated by the manner in which
parasitic losses increase with speed. The research facilities built into LIMPET also
have a major influence on its effectiveness as a power generator. The choice of the
contra rotating turbine, whilst nominally offering an improved aerodynamic
efficiency, also carries with it two sets of electrical losses in inverter and generator
and two sets of loss in the flywheel. Whilst the dominant loss in the electrical system
at the production rates achieved to date is a fixed system loss, the flywheel windage
losses rise with the cube of speed so that any attempt to improve turbine efficiency by
increasing turbine speed carries an immediate penalty of higher losses. The problem
is therefore to determine a method of determining the operational parameters which
takes into account all of these factors. This has been achieved using a verified
mathematical model relating chamber pressure to electrical output and which takes
account of all system losses.
10.4 Revised Control Strategy
It is clear that the single variable that determines the turbine efficiency is the turbine
speed and that at any particular moment there is a turbine speed which is optimal for
the available pneumatic power. Because of the high inertia of the generator and
turbine set it is not possible to match the turbine speed instantaneously, but this needs
to be matched over a period of at least one wave cycle. Given that quasi-steady state
assumptions are used for the turbine performance then the optimum turbine speed is
related to the average pneumatic power over the wave cycle.
Generally, it is assumed that the wave climate is stationary at least for the 15 minute
samples obtained and thus the optimum turbine speed can be related to the average
pneumatic power over this period. Simulations of the LIMPET conversion from
pneumatic to delivered power have been performed on the 53 sea states originally
specified as representative for the site but scaled in respect of the measured power
reaching the collector. This simulation included both the site measured turbine
efficiency and the windage and bearing power losses in the system as derived from
run down tests. The parameters of the simulation were then adjusted to establish the
optimal operational speed for each sea state and the optimal speed plotted as a
function of pneumatic power input inFigure 64
82
0 100 200 300 400 500 600 700
700
800
900
1000
1100
1200
1300
1400
1500
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44 45 46
47
48
49
50
51
52 53
Pneumatic power (kW)
Optimum speed (rpm) =19.016*Pneumatic
P
1 / 3
(W)
T
u
r
b
i
n
e

S
p
e
e
d

(
r
p
m
)

Figure 64
The clear relationship between optimum running speed and pneumatic power allows
an informed decision to be made on the setting of the speed parameters for the control
strategy. The control software has thus been modified so that instead of there being
fixed speeds for ramping the demand generator torque there is a central speed
determined by the relationship of Figure 64 based on a 10 minute moving average of
pneumatic power. The speeds for zero and full torque are set at 50rpm and +50 rpm
relative to this. This revision to the strategy has been implemented and is operational.
Initial results are promising but statistical data over a full season is required before
any firm conclusions can be drawn.
10.5 Stall Control
Until the implementation of the revised control strategy a relatively small increase in
input power could, in the absence of a manual speed change, lead to significant
turbine stall. In storm conditions the turbines could be in gross stall for prolonged
periods resulting both in poor energy conversion, high noise generation and high
levels of plant vibration. An immediate impact of the revised strategy was to
dramatically reduce the incidence of such stall. Given that the centre speed is being
set on the basis of a 10 minute moving average of input power an atypically large
wave within the averaging period can still lead to stall. To reduce the incidence of
such occurrence a further control modification is in preparation whereby the butterfly
valve will close in response to a 30 second average in chamber pressure in the
expectation that stall can to a large extent be anticipated and prevented by restricted
the air flow to the turbine. The modification is currently being tested in simulation
and is likely to implemented in late 2002.
83
10.6 Discussion and Conclusion
Whilst the baseline control strategy has operated reliably and safely it is clearly sub-
optimal in respect of power conversion efficiency. Analysis of site data using a
chamber pressure-wire simulation has allowed the development of a revised strategy
which has been implemented on the plant and is currently being monitored. It is
expected that the modifications will improve conversion efficiency and reduce the
incidence of stall. Further improvements are planned.
11 Task 10 Installation of the Blow Off Valve
The model investigation of the potential of a blow off valve (Task 4) showed that
fitment to LIMPET would be counterproductive. In consequence the task was deleted
from the scope of work for the project.
12 Task 11- Structural Load Measurement at the Full Scale
To examine the wave loading on LIMPET, two beams containing multiple pressure
transducers have been installed. One beam containing five pressure transducers has
been installed on the external face of the front wall of LIMPET; the other beam
containing five pressure transducers has been installed on the back wall of LIMPETs
water column. Both beams have been installed along the vertical centreline of the
southern water column section. Observation of the water motion at LIMPET has
indicated that this is generally the area with the largest wave activity.
Wave forces on structures can be separated into two distinct types, forces associated
with the change in momentum of water and forces due to compression of air
entrapped or entrained by a breaking wave. Of these two types of wave force, the
forces due to air entrapment or entrainment are significantly larger than the forces due
to changes in water momentum and thus for determining maximum wave loading are
more important. However, the forces from air entrapment only occur for very short
periods of time and thus to capture these events the signal from the transducers needs
to be collected at a relatively high frequency, typically at about 10 kHz.
To capture time series data at this speed the data loggers have to operate in burst
mode, using a trigger to start the collection of data. Unfortunately, the integration of
operating the data loggers in burst mode and data acquisition has not been fully
successful, thus there is currently no high frequency time series data available for the
wave loading on LIMPET. Some low frequency (5 Hz) data from the pressure
transducers has been collected, which shows pressures that would typically be
associated with changes in water momentum, however higher forces due to air
entrapment have not been observed.
The data loggers are also able to operate in statistical mode where only statistical
data for the signal is maintained. Of interest in this data set are the maximum values
obtained from the pressure transducers. In these data sets, recordings that could be
associated with air entrapment were obtained, however it is quite possible that these
84
were due to isolated impacts of water droplets, or the bursting of a small bubble over
the pressure transducers. Although it would not be possible to observe the wave
impacts on the front of the LIMPET during a severe storm the wave activity has been
observed with waves breaking in front of the device. As the sea state increases, the
waves tend increasingly to break further offshore. During the time of observation, no
wave has ever been seen to break onto LIMPET. This suggests that the high wave
loads due to air entrapment may not occur on the LIMPET structure due to the
particular bathymetry in which it is situated. Work continues on obtaining the
necessary time series wave loading data to determine whether wave loads due to air
entrapment occur and thus determine the appropriateness of the current LIMPETs
structural design.
85
13 Task 13 Fitment of Acoustic Attenuation
Figure 65
The operational experience of QUB on the 75kW prototype unit had highlighted noise
as a factor to be considered in the design of the turbo generation equipment. The
approach taken with LIMPET was to construct the unit in a secure housing (Figure
65), but without any specific acoustic treatment, to measure the noise output of the
plant and then to retrofit attenuation. On starting the plant the overall noise level was
measured at a point some 5m to the south of the exit. A peak value of 92dB(A) was
recorded. This level was highly variable dependent upon the wave conditions and the
turbine speed. At low turbine speeds (600-800rpm) the resistance of the turbine to
flow is low and the duct velocities are high so that the turbine is fully stalled in
significant waves. Under these artificial conditions the noise generated by eddy
turbulence of the turbine blades is high and dominates the overall noise. As the
turbine speed increases and with it the turbine damping flow rates fall and the turbine
is less prone to stall. At the same time the direct noise from the turbine rotation
(primarily at blade passing frequencies) increases so that a turbine whine becomes
evident. Except in storm conditions where occasional stall still occurs this turbine
whine predominates the perceived noise at speeds above 1100rpm. Whilst the plant
was operated to permit the performance of specific commissioning tasks it was
considered inappropriate to run on a continuous basis until attenuation had been fitted.
86
Figure 66
Acoustic specialists EMTEC were employed to design and fit the attenuation system.
The main difficulty was in the reduction of the low frequency noise components (80-
200Hz) which dominate the noise output under stall conditions. The approach adopted
was as shown in Figure 66 . The louvred side walls of the original enclosure were
removed and replaced with solid blockwork lined on the interior with acoustic panels.
A diagonal plate designed to provide multiple reflection of the noise was fitted across
the chamber and exit silencers fitted for the exhaust at the top of the chamber. The
treatment was extremely effective reducing the external noise level by 27dB to
65dB(A). This meant that under normal running conditions the plant was inaudible at
the nearest habitation. The specification put on the acoustic treatment required that the
pressure drop from the turbine exhaust to atmosphere should be no more than 200Pa.
This was achieved but it was subsequently established that not only was it important
to limit the pressure drop but it was also necessary to ensure that the acoustic
treatment did not cause a mal-distribution of flow in the turbine. This was not
achieved with the result that on the inlet flow to the turbine the turbine efficiency is
significantly reduced by the influence of the treatment. See section 8.3.3 of the report
for DTI Sustainable Energy Programmes project V/06/00180.
14 Task 14 Removal of Rock from the Gully
On completion of construction, the bund wall, which had been preventing the sea
entering the excavated gully, was removed by blasting leaving some 2,500m
3
of
shattered rock in front of the device. The contractors mobilised a long reach excavator
to site and used this to remove all the accessible rock. There were problems however
in that excavator delivered to site had a lesser reach than that ordered and could not
access all areas of the gully. Furthermore the blast was performed later in the year
than originally planned which meant that there was no opportunity for a diver survey
Resonator
Plate
Exit Silencer
Wall Absorber
87
to confirm that all rock had been cleared prior to demobilisation of the equipment.
There was concern therefore that blast debris had been left in the gully and that this
would wash into the entrance of the collector blocking access to the water column.
This proved to be the case. After the initial commissioning of the plant the
performance was seen to decline and this was taken as an indication that the collector
entrance was progressively blocking. When a diver survey was performed this
assumption was confirmed with the entrance more than 2/3 blocked. In consequence
a long reach excavator was mobilised together with a rock breaking attachment to
permit large rock segments from the blast to be split before removal.
Figure 67
With a 22m boom (Figure 67) the excavator could reach past the centreline of the
gully and as such had a much greater access than the previous machine which had a
14m boom. During the re-excavation approximately 350m
3
of debris was removed,
the majority from directly in front of the collector entrance. At the completion of this
work a diver survey showed that with the exception of a small number of relatively
large rocks under the lip of the collector (and therefore outwith the reach of the
excavator) the gully was clear. These rocks were removed by tying a rope around
them and towing them away from the gully. It was encouraging to note that the diver
surveys both prior to and after the re-excavation showed that outwith the gully there
was no significant quantity of loose rock. This gives the expectation that unless rock
is washed from the land that the entrance to the collector, once cleared, should stay
clear. The effects of the rock blockage on performance are further discussed in the
report for DTI Sustainable Energy Programme V/06/00180.
88
15 Task 15 Fitment of Variable Orifice to the Second Turbine
Outlet
A second outlet was included in the LIMPET collector design to offer flexibility in
the use of the device for research purposes. A number of possibilities were considered
at the design stage including:
The fitment of an alternative power take off.
The isolation of the north water column.
The fitment of a variable resistance valve to allow the water column damping to
be measured.
The first two items on this list remain possibilities for future development and
research. In the first instance however, a butterfly valve, which can service as a
variable orifice in order to give a controlled exit flow resistance to the chambers has
been fitted as shown in Figure 68.
Figure 68
By progressively opening the valve it was possible to vary the damping on the water
column from a relatively high level, where the actual damping was determined by
leakage at the circumference of the two butterfly valves, to a low level when the north
valve was fully open. In principle the same effect could have been achieved by
opening the valve in the main turbine duct but this could have led to an unacceptable
pressurisation of the turbine hall and acoustic enclosure. With variable water column
damping the water column motion and chamber pressure were measured and the
pneumatic power capture and water column damping calculated from this data.
89
Orifice Opening Vs Pneumatic power capture
0.00
50,000.00
100,000.00
150,000.00
200,000.00
250,000.00
300,000.00
0 100 200 300 400 500 600 700
Cylinder extension (mm)
E
s
t
i
m
a
t
e
d

p
n
e
u
m
a
t
i
c

p
o
w
e
r

c
a
p
t
u
r
e

(
W
)
Test series 07/09/01
Test series 08/09/01
shows the water column damping and Figure 70 the pneumatic power capture. The x
axis in each case is an arbitrary scale. Two data sets representing wave conditions on
different days are plotted.
As expected, the equivalent turbine damping increases as the orifice is closed.
However, the maximum value, limited by leakage, was only approximately 100
Ns/m
5
. This value corresponds to optimum estimated from model tests.
Because the incident power is continually changing, due to both the change in climate
and tide level, the interpretation of the estimated pneumatic power was difficult and
the results are inconclusive. There are indications however that the optimal damping
for peak pneumatic power is in the range of 80-100Ms/m
5
.
90
Orifice Opening Vs Turbine damping coefficient
0.00
20.00
40.00
60.00
80.00
100.00
120.00
0 100 200 300 400 500 600 700
Cylinder extension (mm)
E
q
u
i
v
a
l
e
n
t

t
u
r
b
i
n
e

d
a
m
p
i
n
g

c
o
e
f
f
i
c
i
e
n
t

(
N
s
/
m
^
5
)
Test series 07/09/01
Test series 08/09/01
Figure 69
Orifice Opening Vs Pneumatic power capture
0.00
50,000.00
100,000.00
150,000.00
200,000.00
250,000.00
300,000.00
0 100 200 300 400 500 600 700
Cylinder extension (mm)
E
s
t
i
m
a
t
e
d

p
n
e
u
m
a
t
i
c

p
o
w
e
r

c
a
p
t
u
r
e

(
W
)
Test series 07/09/01
Test series 08/09/01
Figure 70
91
16 Results and Conclusions
The project has delivered both tangible and intangible results. The experience of
building and operating LIMPET will be of significant benefit in the development of
future shoreline and off-shore plant. The main results and conclusions were as
follows:
The initial operation of the LIMPET plant under the baseline control strategy
demonstrated that wave energy plant can be operated reliably and remotely under
automatic control. In furthering the development of the control strategy it has been
established that there is a clear relationship between the power incident on the
collector and the optimal turbine speed. The control strategy, which is based upon
the continuous variation of desired turbine speed in relation to a moving average of
the input pneumatic power, has been modified to take into account this
relationship.
It has been demonstrated that, contrary to previous reports, pressure relief
valves in the collector chamber are unlikely to improve the performance of OWC
devices. The high energy cost of operating an effective valve added to the
difficulty in achieving satisfactory valve operation in irregular seas outweighs the
potential benefits.
It has been shown that considering internal pressure forces and the external
wave loads on an OWC structure as independent loads can lead to a gross over
design of the structure. An single evaluation of the total system forces enables a
major reduction in design loads which will allow for a simpler and lower cost
structure.
Scale model tests have confirmed that water depth has a critical effect on the
performance of a shoreline OWC. They have also shown that the benefits of the
harbour wall effect previously reported for OWCs in deep and intermediate waters
do not occur in shallow water and that the parallel gully is detrimental to device
performance.
Scale model measurements of structural loads on the LIMPET front wall have
shown that the two most challenging load conditions applied to the design, external
wave slam and peak internal pressure are not independent but are generally in
opposition. If the peak net load is used as the basis of design for the front wall
rather than the peak of the individual loads then the structural requirement of the
front wall is significantly reduced.
The demonstration of extreme loads very much less than those assumed in the
design of LIMPET has permitted a review of the design philosophy and led to a
new concept of collector construction which has reduced the concrete volumes per
unit capture width to approximately one third of that used in LIMPET.
92
A correlation has been achieved between freely available offshore buoy data
and the incident power at the LIMPET site. In the future it is anticipated that this
relationship can be developed to the point where it becomes possible to predict
wave energy generation at least 24 hours in advance.
Acoustic attenuation fitted to the exit of the turbine system has proved
extremely effective in preventing any noise disturbance from the LIMPET plant.
The particular design adopted has however caused some flow disturbance and has
affected turbine efficiency.
Pressure transducers fitted to the inner rear wall of the collector have shown
minimal internal wave loads showing that the internal design of the collector has
been successful in eliminating the incidence of internal wave slam.
A variable valve was fitted to the second turbine outlet. This enabled the
pressure flow characteristic of the collector to be measured independent of the
turbine operation.
After the first winter of operation it was found that debris from the original
construction remained in the gully. This was successfully removed in the summer
of 2001 using a long reach excavator.
17 Future Developments
The LIMPET wave energy device will continue operate as a grid connected generator
for the foreseeable future. In this short term a number of options are being considered
for changing the plant hardware to improve the power capture performance of the
device. These include:
Changing the turbine hardware to improve pneumatic to shaft efficiency and
reduce parasitic losses.
Introducing a tapered gully
Modifying the acoustic attenuation to reduce flow disturbance.
At the same time the programme of data capture and analysis with continue in order
to better understand the complex factors affecting the overall plant performance and
to facilitate further improvements in performance via changes to the control strategy.
The successful construction and commissioning of LIMPET has led to wide spread
interest in the adoption of the basic technology and it is likely that this will lead to the
construction of LIMPET derivatives within a relatively short timescale.

Das könnte Ihnen auch gefallen