Analysis of Waves Technical Documentation For WaveLab 3
Analysis of Waves Technical Documentation For WaveLab 3
Analysis of Waves
Technical documentation for WaveLab 3
Frigaard, Peter; Andersen, Thomas Lykke
Publication date:
2014
Document Version
Publisher's PDF, also known as Version of record
General rights
Copyright and moral rights for the publications made accessible in the public portal are retained by the authors and/or other copyright owners
and it is a condition of accessing publications that users recognise and abide by the legal requirements associated with these rights.
- Users may download and print one copy of any publication from the public portal for the purpose of private study or research.
- You may not further distribute the material or use it for any profit-making activity or commercial gain
- You may freely distribute the URL identifying the publication in the public portal -
Take down policy
If you believe that this document breaches copyright please contact us at [email protected] providing details, and we will remove access to
the work immediately and investigate your claim.
Peter Frigaard
Thomas Lykke Andersen
ISSN 1901-7286
DCE Lecture Notes No. 33 Department of Civil Engineering
Aalborg University
Department of Civil Engineering
Water and Soil
Analysis of Waves
Technical documentation for WaveLab 3
by
Peter Frigaard
Thomas Lykke Andersen
October 2014
c Aalborg University
Scientific Publications at the Department of Civil Engineering
Technical Reports are published for timely dissemination of research re-
sults and scientific work carried out at the Department of Civil Engineering
(DCE) at Aalborg University. This medium allows publication of more de-
tailed explanations and results than typically allowed in scientific journals.
Technical Memoranda are produced to enable the preliminary dissemi-
nation of scientific work by the personnel of the DCE where such release is
deemed to be appropriate. Documents of this kind may be incomplete or
temporary versions of papers—or part of continuing work. This should be
kept in mind when references are given to publications of this kind.
Contract Reports are produced to report scientific work carried out un-
der contract. Publications of this kind contain confidential matter and are
reserved for the sponsors and the DCE. Therefore, Contract Reports are
generally not available for public circulation.
Lecture Notes contain material produced by the lecturers at the DCE for
educational purposes. This may be scientific notes, lecture books, example
problems or manuals for laboratory work, or computer programs developed
at the DCE.
Theses are monograms or collections of papers published to report the sci-
entific work carried out at the DCE to obtain a degree as either PhD or
Doctor of Technology. The thesis is publicly available after the defence of
the degree.
Latest News is published to enable rapid communication of information
about scientific work carried out at the DCE. This includes the status of
research projects, developments in the laboratories, information about col-
laborative work and recent research results.
Published 2014 by
Aalborg University
Department of Civil Engineering
Sohngaardsholmsvej 57,
DK-9000 Aalborg, Denmark
1 Introduction 5
3 Frequency-Domain Analysis 21
4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
4.3 Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
1
4.4 Goda & Suzuki’s method . . . . . . . . . . . . . . . . . . . . . 40
5.1 Principle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
5.3 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69
7.5 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99
9 References 115
Chapter 1
Introduction
The present book describes the most important aspects of wave analysis tech-
niques applied to physical model tests. Moreover, the book serves as technical
documentation for the wave analysis software WaveLab 3, cf. Aalborg Uni-
versity (2012). In that respect it should be mentioned that supplementary to
the present technical documentation exists also the online help document de-
scribing the WaveLab software in detail including all the inputs and output
fields. In addition to the two main authors also Tue Hald, Jacob Helm-
Ptersen and Morten Møller Jakobsen have contributed to the note. Their
input is highly acknowledged.
5
Chapter 2
Time-Domain Analysis of
Waves
7
Figure 2.2: Application of zero-down crossing.
Table 4.1. Ranked individual wave heights and corresponding periods in Fig. 2.2.
rank i 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15
H (m) 5.5 4.8 4.2 3.9 3.8 3.4 2.9 2.8 2.7 2.3 2.2 1.9 1.8 1.1 0.23
T (s) 12.5 13.0 12.0 11.2 15.2 8.5 11.9 11.0 9.3 10.1 7.2 5.6 6.3 4.0 0.9
wave no.
in 2.2 7 12 15 3 5 4 2 11 6 1 10 8 13 14 9
Both wave height and wave period can be considered as random variables,
which have certain probability distributions.
Mean wave: H, T
H and T are the mean values of the heights and periods , respectively, of all
8
individual waves. Table 4.1 yields
1 1
15 15
H = Hi = 2.9 m T = Ti = 9.25 s
15 i=1 15 i=1
1 N
Hrms = H2
N i=1 i
The significant wave height is the average of the wave heights of the one-third
highest waves. The significant wave period is the average of the wave periods
of the one-third highest waves. From Table 4.1 one finds
1 1
5 5
Hs = Hi = 4.44 m Ts = Tj = 12.8 s i, j are the rank no.
5 i=1 5 j=1
The significant wave is very often used as the design wave. The reason might
be that in old days structures were designed an a basis of visually observed
waves. Experiences show that often the wave height and period reported
by visual observation correspond approximately to the measured significant
wave. Therefore the choice of significant wave as design wave can make use
of the existing engineering experience.
9
Maximum wave: Hmax , THmax
This is the wave, which has the maximum wave height. In Table 4.1,
Note, however, that Hmax is a random variable which depends on the number
of individual waves in the time series.
The maximum wave from a long time series corresponding to a storm with a
return period of e.g. 100 years is often chosen as the design wave for struc-
tures which are very important and very sensitive to wave loads.
H1/10 is the average of the wave heights of the one-tenth highest waves. TH1/10
is the average of the wave periods of the one-tenth highest wave.
Instead of showing all individual wave heights, it is easier to use the wave
height histogram which gives information about the number of waves in var-
ious wave height intervals. Fig. 2.3 is the histogram of wave heights corre-
sponding to Table 4.1.
Non-dimensionalized histogram
10
Figure 2.3: Wave height histogram.
the Rayleigh distribution in case of deep water waves. In other words, the
individual wave heights follow the Rayleigh distribution.
Rayleigh distribution
π π H
f (x) = x exp − x2 where x= (2.1)
2 4 H
11
The Rayleigh distribution function is
π
F (x) = Prob{X < x} = 1 − exp − x2 (2.2)
4
H1/10 = 2.03 H
H1/3 = 1.60 H
H2% = 2.23 H
H0.1% = 2.97 H
12
Individual wave height distribution in shallow water
1
2
− 13
ln100 Hm0
H1% = Hm0 1+ (2.5)
2 h
1
2
− 12
ln1000 Hm0
H0.1% = Hm0 1+ (2.6)
2 h
where h is the water depth, H1% means the 1% exceedence value of the wave
height determined by zero-down-crossing analysis, whereas the significant
wave height Hm0 is determined from the surface elevataion spectrum, cf.
chapter 3. The correction formulae are very useful for checking the wave
height distribution in small scale physical model tests, cf. Fig. 2.6.
Figure 2.6: Comparison of the expression by Stive, 1986, for shallow water
wave height distribution with model test results. Aalborg University Hy-
draulics Laboratory 1990.
13
Klopmann et al. (1989) proposed also a semi-empirical expression for the
individual wave height distribution. Worth mentioning is also the very much
used method of Battjes and Groenendijk (2000) which was calibrated against
a lot of physical model test data, but the numerical procedure is somewhat
more complicated.
Distribution of Hmax
π
FX (x) = Prob{X < x} = 1 − exp − x2 (2.7)
4
Note that FXmax (x) can be interpreted as the probability of the non-occurrence
of the event ( X > x ) in any of N independent trials. The probability density
function of Xmax is
dFXmax
fXmax (x) =
dx
π π π N −1
= N x exp − x2 1 − exp − x2 (2.9)
2 4 4
1
A storm usually lasts some days. The significant wave height is varying during a
storm. However we are more interested in the maximum significant wave height in a short
period of time. In practice, N is often assumed to be 1000.
14
The density function of X and the density function of Xmax are sketched in
Fig. 2.7.
Mean, median and mode are often used as the characteristic values of a
random variable. Their definitions are given in Fig. 2.8.
+∞
Mean xmean = E[X] = xfX (x)dx
−∞
Median xmedian = x
FX (x)=0.5
Mode xmode = x
fX (x)=max
ln N 0.577
(Hmax )mean ≈ + √ Hs (2.10)
2 8 ln N
ln N
(Hmax )mode ≈ Hs (2.11)
2
15
Furthermore, (Hmax )μ , defined as the maximum wave height with exceedence
probability of μ (cf. Fig. 2.9), is
⎛ ⎞
1
N
(Hmax )μ ≈ ln ⎝ ⎠ Hs (2.12)
2 ln 1
1−μ
We get for the commonly used N = 1000 the following maximum wave
heights:
16
Monte-Carlo simulation of Hmax distribution
− ln(1 − F (H))
H = F −1 ( F (H) ) = Hs (2.17)
2
4) Repeat steps 2) and 3), say, 10,000 times. Thus we get 10,000
values of Hmax .
17
2.5 Distribution of wave periods
It is summarized as
Non-linear waves do not have a gaussian distributed surface as the waves are
assymetric with higher and narrower crests and wider and less deep troughs.
This distortion of the distribution of the surface can be described by the
skewness (β1 ) and kurtosis (β2 ) defined as:
1
N
1
β1 = 3
· (ηi − η)3 (2.18)
ηrms N i=1
1
N
1
β2 = 4
· (ηi − η)4 (2.19)
ηrms N i=1
For a Gaussian surface the skewness is zero. Positive skewness indicate that
the median value is below the mean value and a heavy positive tail which
will be the case for non-linear waves. Huang and Long (1980) found that in
deep water the skewness is proportional to the wave skewness. The kurtosis
describes the peakedness of the statistical distribution. A gaussian process
has a kurtosis of 3 and higher values indicate high narrow peak and heavier
tails than given by the gaussian distribution.
Goda (1986) introduced the overall atiltness paramter (β3 ) which describes
18
the assymetry in the direction of propagation.
N
−1
1
N −1
(η̇i − η)3
i=1
β3 = 3/2
(2.20)
N
−1
1
N −1
(η̇i − η)2
i=1
Waves in the near shore zone as the waves tend to have a steep front slope
and a gentle rear slope. Such waves would have a positive atiltness and the
atiltness paramter can exceed unity within the surf zone.
19
Chapter 3
Frequency-Domain Analysis
Spectral analysis of irregular waves is very important for the design of struc-
tures. For example, in the oil-drilling platform design where wave forces
plays an important role, it is of importance to design the structure in such a
way that the natural frequency of the structure is rather far away from the
frequency band where the main part of wave energy concentrates. In this
way resonance phenomenon and the corresponding dynamic amplification of
force and deformation can be avoided.
Surface elevation
The surface elevation of a linear wave is:
H
η(x, t) = cos(ωt − kx + δ) = a cos(ωt − kx + δ) (3.1)
2
21
where H wave height
a amplitude, a = H/2
ω angular frequency, ω = 2π/T
T wave period.
k wave number, k = 2π/L
L wave length
δ initial phase
The relation between wave period and wave length (dispersion relationship)
is
g T2 2πh
L = tanh (3.3)
2π L
Wave energy
1 1
E = ρ g H 2 = ρ g a2 (Joule/m2 in SI unit) (3.4)
8 2
1
T
= T 0
η 2 (t) dt (T: wave period)
1
= 2
a2
22
Superposition of linear waves
Since the governing equation (Laplace equation) and boundary conditions are
linear in small amplitude wave theory, it is known from mathematics that
small amplitude waves are superposable. This means that the superposition
of a number of linear waves with different wave height and wave period will
be:
4
4
η(t) = ηi (t) = ai cos(ωi t + δi ) (3.5)
i=1 i=1
23
Figure 3.1: Simulation of irregular waves by superposition of linear waves.
Variance diagram
Instead of Fig. 3.1, we can use a variance diagram, shown in Fig. 3.2, to
describe the irregular wave.
In comparison with Fig. 3.1, the variance diagram keeps the information on
amplitude (ai ) and frequency (fi , hence Ti and Li ) of each component, while
the information on initial phase (δi ) is lost. This information loss does not
24
matter because the surface elevation of irregular wave is a random process.
We can simply assign a random initial phase to each component.
1 2
2
a
Sη (f ) = (m2 s) (3.6)
Δf
25
When Δf approaches zero, the variance spectrum becomes a continuous
curve. A variance spectrum is also called energy spectrum. But strictly
speaking, the energy spectral density should be defined as
1
ρ g a2
S(f ) = 2
( Jm2s ) (3.7)
Δf
We can also construct time series of surface elevation from variance spectrum.
In Fig. 3.4 the known variance spectral density Sη (f ) is divided into N parts
by the frequency band width Δf . This means that the irregular wave is
composed of N linear waves
N
N
η(t) = ηi (t) = ai cos(ωi t + δi ) (3.8)
i=1 i=1
1 2
Sη (fi ) Δf = a i = 1, 2, · · · , N (3.9)
2 i
ai = 2 Sη (fi ) Δf i = 1, 2, · · · , N (3.10)
2π
ωi = = 2πfi i = 1, 2, · · · , N (3.11)
Ti
The initial phase δi is assigned a random number between 0 and 2π. Hence
by use of Eq. (3.8) we can draw the time-series of the surface elevation of the
irregular wave which has the variance spectrum as shown in Fig. 3.4.
26
3.3 Fourier series
∞
2πi 2πi
x(t) = a0 + ai cos t + bi sin t
i=1
T0 T0
∞
= (ai cos ωi t + bi sin ωi t) (3.12)
i=0
2
Not all mathematicians agree that an arbitrary function can be represented by a
Fourier series. However, all agree that if x(t) is a periodic function of time t, with period
T0 then x(t) can be expressed as a Fourier series. In our case x(t) is the surface elevation
of irregular wave, which is a random process. if T0 is large enough, we can assume that
x(t) is a periodic function with period T0 .
27
where ai and bi are Fourier coefficients given by
1
T0
a0 = T0 0
x(t) dt and b0 = 0
2
T0 ⎫
ai = T0 0
x(t) cos ωi t dt ⎬
T0 i = 1, 2, 3, · · · , ∞ (3.13)
2 ⎭
bi = T0 0
x(t) sin ωi t dt
Physical interpretation
Now we say that the continuous function x(t) is the surface elevation of
irregular wave. η(t) can be expanded as a Fourier series.
∞
η(t) = (ai cos ωi t + bi sin ωi t)
i=0
∞
= ci sin δi cos ωi t + ci cos δi sin ωi t
i=0
∞
= ci (sin δi cos ωi t + cos δi sin ωi t)
i=0
∞
= ci sin(ωi t + δi )
i=0
∞
= ci cos(ωi t + δi ) (3.14)
i=0
where δi = δi − π
2
and sin(x + y) = sin(x) cos(y) + cos(x) sin(y) have been
used.
28
function, is composed of infinite number of linear waves with
amplitude ci = a2i + b2i
i = 0, 1, · · · , ∞ (3.15)
2π T0
period Ti = ωi
= i
If the sampling frequency is fs , then the time interval between two succeeding
points is Δt = 1/fs . Corresponding to the sample duration T0 the total
number of sample is N = T0 /Δt. Thus we obtain a discrete time series of
surface elevation
η0 , η1 , ··· , ηN −1
29
can be obtained by Fast Fourier Transforms (FFT)3 . That is to say, the
irregular wave surface elevation, expressed by digital time series, is composed
of N linear waves
N −1
N −1
η(t) = ηi (t) = a2i + b2i cos(ωi t + δi ) (3.16)
i=0 i=0
⎫
amplitude a2i + b2i ⎪
⎪
⎪
⎪
angular frequency ωi = 2πi ⎬
T0
i = 0, 1, · · · , N − 1 (3.17)
period Ti = 2π
= Ti0 ⎪
⎪
ωi ⎪
⎪
1 ⎭
frequency fi = Ti
= Ti0
1
frequency width Δf = fi+1 − fi =
T0
1 1 2
2
(amplitude)2 (a
2 i
+ b2i )
spectral density Sη (fi ) = = (3.18)
Δf Δf
3
FFT is a computer algorithm for calculating DFT. It offers an enormous reduction in
computer processing time. For details of DFT and FFT, refer to Newland (1975)
30
Nyquist frequency fnyquist
i
fi = i = 0, 1, · · · , N − 1 (3.19)
T0
N 1 T0
2 2 Δ fs
fnyquist = f N = = = (3.20)
2 T0 T0 2
Fig. 3.8 gives an example on aliasing after the Fourier analysis of discrete
time series of a linear wave.
31
Figure 3.8: Aliasing after Fourier analysis.
32
Taper data window
33
3.5 Characteristic wave height and period
The variance spectrum, illustrated in Fig. 3.11, says nothing about how
high the individual waves will be. Now we will see how to estimate the
characteristic wave height and period based on the variance spectrum.
n order moment mn
mn is defined as
! ∞
mn = f n Sη (f ) df (3.22)
0
which is actually the area under the curve, cf. Fig. 3.11.
From the definition of mn , it can be seen that the higher the order of moment,
the more weight is put on the higher frequency portion of the spectrum. With
the same m0 , a wider spectrum gives larger values of the higher order moment
(n ≥ 2). Longuet-Higgins has defined a broadness parameter:
"
m22
ε4 = 1 − (3.24)
m0 m4
34
An alternative parameter is the narrowness paramter:
m0 m2
ε2 = −1 (3.25)
m21
In reality ε lies in the range of 0.4-0.5. It has been found that Rayleigh dis-
tribution is a very good approximation and furthermore conservative, as the
Rayleigh distribution gives slightly larger wave height for any given proba-
bility level.
When wave height follows the Rayleigh distribution, i.e. ε = 0 , the signifi-
cant wave height Hs 4 can theoretically be expressed as
√
H s ≈ H m0 ≡ 4 m0 (3.26)
√
Hs = 3.7 m0 (3.27)
4
Hm0 denotes a wave height determined from spectrum while Hs or H1/3 is significant
wave height determined from time-domain analysis. They are equal to each other when
wave height follows the Rayleigh distribution.
35
36
Chapter 4
Reflection Analysis of
Long-Crested Waves
4.1 Introduction
Coastal and offshore engineering cover a wide spectrum in the field of civil
engineering. Some of the major subjects involved are the construction of
harbours and breakwaters.
Design of breakwaters and breakwater lay-outs are based on the need for an
acceptable wave climate in the harbour and in the harbour entrance. How-
ever, the complex behaviour of waves propagating into a more or less closed
basin makes the design of the quaywalls and breakwaters important.
The wave field in front of a structure consist of both incident waves and re-
flected waves. As one cannot first measure the incident waves and then the
reflected waves, methods are required to separate or distinguish between the
incident and reflected waves.
37
4.2 Wave Reflection
Wave reflection occurs when waves are propagating onto breakwaters and
quaywalls.
Separating the incident waves and the reflected waves from a recorded wave
elevation time series has several purposes. Either:
In order to apply physical models and numerical models the reflection char-
acteristics must be modeled correct.
In the following several different methods for separating incident waves and
reflected waves will be presented. After separation of the incident waves and
the reflected waves it is easy to calculate the reflection characteristics for the
structure.
The following methods for separating the waves will be divided into two main
groups:
There is a wide need for 2D methods due to the amount of experiments car-
ried out in wave flumes. In wave flumes reflection and re-reflection influence
upon the validity of measurements if it is not possible to extract the incident
38
waves. Further when operating in the near shore environment the wave-field
will often be approximately two-dimensional as the waves refract, i.e. they
bend towards orthogonals to the shore.
The 2D methods presented herein are all based on surface elevation measure-
ments at a number of positions.
Three of the 2D methods are derived from the same principles but for vari-
ous number of probes. The simplest method requires two probes. The other
methods need three or more probes. The three first methods work in fre-
quency domain and give the incident wave spectrum as well as the reflected
wave spectrum. The last 2D method works in time domain and gives the
incident wave timeserie.
4.3 Methods
In the following four methods for separation of incident and reflected waves
in a two-dimensional wave field will be presented.
The first three methods are all working in the frequency domain and have
the same basic principle. They assume the wave elevation to be a sum of
regular waves travelling with different frequency and phase. The first method
by Goda & Suzuki needs measurements of the wave elevation in two distinct
points. Hence by use of Fourier analysis the amplitude of the incident and
reflected waves for a given frequency can be estimated. This procedure does
not account for the noise which probably is contaminated in the measured
39
wave signals, i.e. the measured wave elevations.
The method presented by Mansard & Funke takes this into account, but
also requires the wave elevation to be measured in three distinct points. By
applying Fourier analysis this should result in the same waves except that the
measured noise varies from wave gauge to wave gauge. That is, 10 Fourier
coefficients are needed but only 6 are available. Instead the noise is expressed
in terms of the Fourier coefficients for the incident and reflected waves and
the measured coefficients. The best estimate of the Fourier coefficients is
then found by minimising the noise.
The method by Zelt & Skjelbreia extents the method to apply to an arbitrary
number (though larger than one) of wave gauges. Further a weighting of
each frequency component is introduced. This is to control the possible
influence of singularities which occur for some geometric relations between
the distances between the probes. Grønbech et. al (1996) presented a revised
version of the Zelt & Skjelbreia method to separate cross-modes from the
incident and reflected waves. The method is good to quantify the amount of
cross-mode activity and if neglectable the standard method can be applied.
However, due to the more degrees of freedom more gauges are needed to give
reliable results.
Other methods have been published to account for sloping bottoms and non-
linear waves, but they will not be presented herein. The methods by Goda
& Suzuki and Mansard & Funke are the methods being used most often
as they are relatively simple to apply and yield reliable results within most
applications.
The method presented by Goda & Suzuki (1976) is the most simple method.
It is based on the assumption that the wave elevation can be considered as a
sum of waves travelling with different frequency, amplitude and phase. Fur-
ther for each wave a reflected wave will travel in the opposite direction. The
method makes use of Fourier analysis and will due to singularities put con-
40
straint to the distance between the waveprobes. The method is very easy to
implement but has a lack of accuracy as no measuring errors are accounted
for.
N
η(x, t) = an cos(kn x − ωn t + Φn ) (4.1)
n=1
N
N
η(x, t) = aI,n cos(kn x − ωn t + ΦI,n ) + aR,n cos(kn x + ωn t + ΦR,n )(4.2)
n=1 n=1
In the following the index n will be omitted for simplicity, that is, only one
frequency is considered. Hence eq. (4.2) will consist of only two terms:
41
Figure 4.1: Definition sketch showing x1,2 .
(4.1), yields:
leads to
which is rearranged to
η1 = A1 cos(ωt) + B1 sin(ωt)
(4.8)
η2 = A2 cos(ωt) + B2 sin(ωt)
42
where
A1 = aI cos(kx1 + ΦI ) + aR cos(kx1 + ΦR )
B1 = aI sin(kx1 + ΦI ) − aR sin(kx1 + ΦR )
(4.9)
A2 = aI cos(kx2 + ΦI ) + aR cos(kx2 + ΦR )
B2 = aI sin(kx2 + ΦI ) − aR sin(kx2 + ΦR )
In eq. (4.8) the elevation is seen to be a composite signal of a sine and cosine
signal having different time-constant amplitude, i.e. the LHS of eq. (4.9) must
correspond to the Fourier coefficients which can be obtained from Fourier
analysis of the recorded time series.
Thus eq. (4.9) contains four equations with four unknowns, i.e. aI , aR , ΦI
and ΦR . The solution giving the amplitudes is given by Goda & Suzuki
(1976) as:
aI = 1
2| sin(kx1,2 )|
(A2 − A1 cos(kx1,2 ) − B1 sin(kx1,2 ))2
where x1,2 = x2 − x1 .
It is seen from eq. (4.10) that singularities will occur for sin(kx1,2 ) = 0.
Hence it should be avoided that
x1,2 n
= where n = 0, 1, 2, ...
L 2
Further Goda & Suzuki (1976) suggest to avoid values in the range ±0.05 xL1,2
at the singularity points. For a wide wave spectrum this will be impossible
for all frequencies, but of course values applying for the peak frequencies
should be weighted highest.
For regular waves the method is quite accurate but for irregular waves the
confidence of the FFT analysis plays a significant role. However in both cases
noise may be the dominant error as it cannot be accounted for.
43
4.5 Mansard & Funke’s method
As a natural extension to the method by Goda & Suzuki, Mansard & Funke
(1980) presented a three-points method. Here an additional probe is taken
into use which makes it possible to add an error to the measurements and
hence minimise it in a least squares sense.
The general equation for a progressive wave field is obtained as in the previous
method, i.e. eq. (4.1):
N
η(x, t) = an cos(kn x − ωn t + Φn ) (4.11)
n=1
The wave elevation given by eq. (4.1) and eq. (4.11) is separated into incident
waves and reflected waves, also as previously, but now a noise function is
added. This leads to:
N
N
η(x, t) = aI,n cos(kn x − ωn t + ΦI,n ) + aR,n cos(kn x + ωn t + ΦR,n )
n=1 n=1
or
N
η(x, t) = aI,n cos(kn x − ωn t + Φn )
n=1
N
+ aR,n cos(kn (x + 2xR ) + ωn t + Φn + θs ) + Ω(t) (4.12)
n=1
Here in contrary to Goda & Suzuki (1976) the distance from the point of
observation to the reflecting structure is assumed known. However as the
distance may be impossible to determine (e.g. for a slope) a phase θs is
introduced to compensate herefore. This has the consequence that the phase
Φn remains the same for the incident and reflected wave. Ω(t) is the noise
function and expresses all kinds of errors. xR is the distance from the wave
probe to the reflecting structure.
44
Figure 4.2: Definition sketch.
each probe.
Hence
ηp = η(xp , t)
N
= aI,n cos(kn xp − ωn t + Φn )
n=1
N
+ aR,n cos(kn (xp + 2xR,p ) + ωn t + Φn + θs ) + Ωp (t) (4.13)
n=1
Inserting x1,p , which is the distance from the 1st probe to the p’th probe,
eq. (4.13) can be modified to
N
ηp = aI,n cos(kn (x1 + x1,p ) − ωn t + Φn )
n=1
N
+ aR,n cos(kn (x1 + 2xR,1 − x1,p ) + ωn t + Φn + θs ) + Ω(4.14)
p (t)
n=1
In the following until eq. (4.25) only one frequency will be considered, i.e. in-
45
dex n will be omitted. Thus
ηp = aI cos(k(x1 + x1,p ) − ωt + Φ)
+aR cos(k(x1 + 2xR,1 − x1,p ) + ωt + Φ + θs ) + Ωp (t) (4.15)
Let
A1 + iB1 = ZI + ZR + ZN,1
A2 + iB2 = ZI exp(ikx1,2 ) + ZR exp(−ikx1,2 ) + ZN,2
A3 + iB3 = ZI exp(ikx1,3 ) + ZR exp(−ikx1,3 ) + ZN,3
or in general form
46
Now eq. (4.20) can be rearranged to
ε1 = ZI + ZR − (A1 + iB1 )
ε2 = ZI exp(ikx1,2 ) + ZR exp(−ikx1,2 ) − (A2 + iB2 ) (4.21)
ε3 = ZI exp(ikx1,3 ) + ZR exp(−ikx1,3 ) − (A3 + iB3 )
where
εp = −ZN,p + fe (ZI , ZR )
3
3
2
(εp ) = (ZI exp(ikx1,p ) + ZR exp(−ikx1,p ) − (Ap + iBp ))2 (4.22)
p=1 p=1
should be minimised.
Assuming that the minimum of eq. (4.22) is achieved when both partial
derivatives are zero, i.e.
3 3
∂ p=1 ε2p ∂ p=1 ε2p
= =0 (4.23)
∂ZI ∂ZR
(4.24)
3
0 = 2 p=1 (ZI exp(ikx1,p ) + ZR exp(−ikx1,p ) − (Ap + iBp )) exp(ikx1,p )
3
0 = 2 p=1 (ZI exp(ikx1,p ) + ZR exp(−ikx1,p ) − (Ap + iBp )) exp(−ikx1,p )
47
When written out and again including index n eq. (4.24) leads to
(A1,n + iB1,n ) + (A2,n + iB2,n ) exp(ikn x1,2 ) + (A3,n + iB3,n ) exp(ikn x1,3 )
(A1,n + iB1,n ) + (A2,n + iB2,n ) exp(−ikn x1,2 ) + (A3,n + iB3,n ) exp(−ikn x1,3 )
(4.25)
1
ZI,n = Dn ((A1,n + iB1,n )(R1,n + iQ1,n )
1
ZR,n = Dn ((A1,n + iB1,n )(R1,n − iQ1,n )
where
Dn = 2(sin2 (kn x1,2 ) + sin2 (kn x1,3 ) + sin2 ((kn x1,3 ) − kn x1,2 ))
R1,n = sin2 (kn x1,2 ) + sin2 (kn x1,3 )
Q1,n = sin(kn x1,2 ) cos(kn x1,2 ) + sin(kn x1,3 ) cos(kn x1,3 )
R2,n = sin(kn x1,3 ) sin(kn x1,3 − kn x1,2 )
Q2,n = sin(kn x1,3 ) cos(kn x1,3 − kn x1,2 ) − 2 sin(kn x1,2 )
R3,n = − sin(kn x1,2 ) sin(kn x1,3 − kn x1,2 )
Q3,n = sin(kn x1,2 ) cos(kn x1,3 − kn x1,2 ) − 2 sin(kn x1,3 )
The only unknowns in eq. (4.26) are the Fourier coefficients of the measured
wave elevations. These are obtained by use of FFT analysis of the measure-
ments.
Compared to the previously discussed method this method has the advan-
tage that it minimises the noise contaminated to the elevation measurements.
48
Thus instead of obtaining an exact solution a fitted solution is obtained.
but if sin(kn x1,2 ) = sin(kn x1,3 ) = 0 then sin(kn x1,3 − kn x1,2 ) = 0 whereas the
solution is reduced to
as by definition 0 < x1,2 < x1,3 . Mansard & Funke (1980) suggest that
Ln
x1,2 =
10
Ln Ln
< x1,3 <
6 3
Ln
x1,3 =
5
3Ln
x1,3 =
10
Zelt and Skjelbreia (1992) introduced a method based on the same principles
as the previously described methods.
This method applies for an arbitrary number p of probes and further intro-
duces a weighting of the measurements from the gauges. The latter facility
takes into account the spacing between each pair of wave gauges. However
49
determining the weighting coefficients is a rather subjective process and re-
quires some additional work.
For p = 2 the method corresponds to the method by Goda & Suzuki. For
p = 3 and a uniform weighting, the method corresponds to the method by
Mansard & Funke.
Once again the wave elevation is considered as a sum of incident and reflected
waves, i.e. for the pth probe at position xp utilising eq. (4.2)
N
ηp = η(xp , t) = n=1 aI,n cos(kn xp − ωn t + ΦI,n ) +
N
n=1 aR,n cos(kn xp + ωn t + ΦR,n )
N
ηp = aI,n cos(kn (x1 + x1,p ) − ωn t + Φn )
n=1
N
+ aR,n cos(kn (x1 + 2xR,1 − x1,p ) + ωn t + Φn + θs ) + Ωp (t)
n=1
As in the method of Mansard & Funke (1980) this is by use of the Fourier
coefficients rearranged to eq. (4.21) which has the general form
εp,n = ZI,n exp(ikn x1,p ) + ZR,n exp(−ikn x1,p ) − (Ap,n + iBp,n ) (4.27)
That is, if the estimated coefficients are correct there will be no error,
i.e. εp,n = 0. A function is chosen to be a weighted sum of the squares
of εp,n , i.e.
P
En = Wp,n εp,n ε∗p,n (4.28)
p=1
It is assumed that the minimum of eq. (4.28) occurs where the partial deriva-
50
tives with regard to ZI,n and ZR,n are zero. Thus
P
Wp,n εp,n exp(±ikn x1,p ) = 0 (4.29)
p=1
P P
p=1 Wp,n ZI,n exp(i2kn x1,p ) + p=1 Wp,n ZR,n =
P
p=1 Wp,n (Ap,n + iBp,n ) exp(ikn x1,p )
(4.30)
P P
p=1 Wp,n ZI,n + p=1 Wp,n ZR,n exp(−i2kn x1,p ) =
P
p=1 Wp,n (Ap,n + iBp,n ) exp(−ikn x1,p )
This, eq. (4.30) can be solved with regard to ZI,n and ZR,n as a linear system
of equations. Simply isolate aI,n and aR,n in both equations in eq. (4.30) and
substitute into the other equation respectively. Hence the following solution
is obtained
ZI,n = 1
Sn Pp=1 Wp,n (Ap,n + iBp,n ) exp(−ikn x1,p )
Dn
− Pp=1 Wp,n (Ap,n + iBp,n ) exp(ikn x1,p ) Pq=1 Wq,n exp(−i2kn x1,q )
1 P
ZR,n = Dn Sn p=1 Wp,n (Ap,n + iBp,n ) exp(ikn x1,p )
P
− p=1 Wp,n (Ap,n + iBp,n ) exp(−ikn x1,p ) Pq=1 Wq,n exp(i2kn x1,q )
(4.31)
where
P
P
Dn = Sn2 − Wp,n exp(i2kn x1,p ) Wq,n exp(−i2kn x1,q )
p=1 q=1
P
Sn = Wp,n
p=1
51
4.7 Robustness to possible errors
The following section describes a numerical test, where the elevations are
calculated for monochromatic waves according to the shown equation:
52
Goda & Suzuki
β ω Δt aI α α̃12 α̃13 α̃23 α̃average
% rad/sec sec metre % % % % %
10 2π 0.0625 0.10 50.00 50.88 50.92 50.24 50.68
10 2π 0.0625 0.10 10.00 11.03 10.98 10.14 10.72
All tests were performed with water depth d = 0.50 metre with a 160 sec.
long time series. Distances between wave gauges were x1,2 = 0.25 metre and
x1,3 = 0.60 metre .
The tests show good restistance to random noise. Though, the method pro-
posed by Mansard & Funke gives sligthly better results than the method
proposed by Goda & Suzuki.
53
The tests were repeated for 15 different wave periods and all tests showed
the same tendency as described in the examples above.
All the described methods assume linear waves. In order to examine the
effect from non-linearites a 2-order wave were generated. First with a free
2-harmonic eq. (4.32), and then with a bounded 2-harmonics eq. (4.33):
54
4.7.3 Numerical problems
Numerical problems due to the discretisation of the signal and the succeeding
FFT-analysis.
All tests were performed with water depth d = 0.50 metre with a 80 sec.
long time serie. The time serie were divided into 10 sub-series, which were
cosine tapered. Distances between wave gauges were x12 = 0.25 metre and
x13 = 0.60 metre .
55
Timeseries were calculated from:
56
Chapter 5
Goda and Suzuki (1976) presented a frequency domain method for estimation
of irregular incident and reflected waves in random waves. Mansard and
Funke (1980) improved this method using a least squares technique.
5.1 Principle
To illustrate the principle of the SIRW-method the set-up shown in Fig. 5.1
will be considered. The surface elevation η(x, t) at a distance x from the wave
generator may be written as the sum of the incident and reflected waves: the
incident wave propagating away from the wave generator, and the reflected
wave propagating towards the wave generator. Even though the method
works for irregular waves it will be demonstrated in the following pages for
57
Figure 5.1: Wave channel with piston-type wave generator.
where
f : frequency
a = a(f ) : wave amplitude
k = k(f ) : wave number
φ = φ(f ) : phase
It is seen that the incident wave is phase shifted Δφ = kΔx from signal
η(x1 , t) to signal η(x2 , t), and the reflected wave is phase shifted Δφ = −kΔx
due to opposite travel directions. These phase shifts are called the physical
phase shifts and are denoted φphys I and φphys
R , respectively.
58
The idea in the following manipulations of the elevation signals is to phase-
shift the signals from the two wave gauges in such ways that the incident
parts of the wave signals are in phase while the reflected parts of the signals
are in mutual opposite phase. In this case the sum of the two manipulated
signals is proportional to and in phase with the incident wave signal.
The sum of η ∗ (x1 , t) and η ∗ (x2 , t), which is denoted η calc (t), gives:
59
It is seen that η calc (t) and ηI (x1 , t) = aI cos(2πf t − kx1 + φI ) are identical
signals when the following three conditions are met:
2C cos(0.5(−kΔx − φtheo
1 + φtheo
2 )) = 1 (5.8)
0.5(−kΔx + φ1 + φ2 ) = n · 2π
theo theo
n ∈ (0, ±1, ±2, ..) (5.9)
π
0.5(−kΔx + φtheo
1 − φtheo
2 ) = +m·π m ∈ (0, ±1, ±2, ..) (5.10)
2
φtheo
1 = kΔx + π/2 + mπ + n2π (5.11)
theo
φ2 = −π/2 − mπ + n2π (5.12)
1
C = (5.13)
2 cos(−kΔx − π/2 − mπ)
All the previous considerations and calculations were done in order to find
an amplification and a phase shift for each of the two elevation signals η1
and η2 . Eqs. 5.11 - 5.13 give the result of our efforts, i.e. ηI (x1 , t) = η calc (t).
Remembering that φtheo 1 = φtheo
1 (f ), φtheo
2 = φtheo
2 (f ) and C = C(f ), it is
seen that the goal is already reached in the frequency domain. However, the
implementation of the principle will be done in the time domain using digital
filters. It is seen that singularities may occur. The consequences and the
handling of the singularities will be treated later on in the chapter. Here
it should just be mentioned that one way to bypass the singularities is to
use a velocity meter instead of one of the two wave gauges. Such system is
dealt with in the following chapter while the present chapter deals with two
surface elevation signals only.
η(x1, t) −→ FILTER 1
+ −→ ηI (x1, t)
η(x2, t) −→ FILTER 2
60
The purposes of the filters shown in Fig. 5.2 are exactly a frequency depen-
dent amplification and a frequency dependent phase shift on each of the two
elevation signals.
1
Re{H1 (f )} = · cos(kΔx + π/2)
2 cos(−kΔx − π/2)
1
Im{H1 (f )} = · sin(kΔx + π/2) (5.14)
2 cos(−kΔx − π/2)
1
Re{H2 (f )} = · cos(−π/2)
2 cos(−kΔx − π/2)
1
Im{H2 (f )} = · sin(−π/2) (5.15)
2 cos(−kΔx − π/2)
Based on Eqs. 5.14 and 5.15 it is straightforward to design the time domain
filters. The design of the filters will be given on the next pages.
1 r
N −1
2πr j
h = h(j · Δtf ilter ) =
j
H · exp i (5.16)
N r=0 N
where
r = 0,. . . , N -1
j = 0,. . . , N -1
and H r is the complex frequency response given by Eqs. 5.14 and 5.15 at the
frequency f = r · Δf . Note that H r should be hermitian, i.e. H(fn + fi ) =
H ∗ (fn − fi ).
61
|H(f)|
10.0
8.0
6.0
4.0
2.0
0.0
0.0 1.0 2.0 3.0 4.0 5.0 6.0 7.0
frequency (Hz)
Figure 5.3:
Magnitude (gain) of the frequency responses of a discrete filter.
N = 64 , h = 0.5 m, Δf =0.10 Hz, Δtf ilter = 0.16 sec.,Δx =
0.2m.
1
Δf = (5.17)
N · Δtf ilter
62
h(t), Filter 1
2.0
1.0
0.0
-1.0
-2.0
h(t), Filter 2
2.0
1.0
0.0
-1.0
-2.0
0.0 2.0 4.0 6.0 8.0 10.0
time (sec.)
Figure 5.4:
Filter coefficients corresponding to Filter 1 and Filter 2. N = 64,
water depth h = 0.5 m, Δf =0.10 Hz, Δtf ilter = 0.16 sec.,Δx =
0.2m.
N −1
∗p
η = hj · η p−j (5.18)
j=0
where
j,p = 0,. . . , N -1
η p−j : elevation at time t = (p − j) · Δtf ilter
η ∗p : output from filter at time t = p · Δtf ilter
hj : the filter coefficient corresponding to time t = j · Δtf ilter
Fig. 5.3 indicates that in the present example, singularities are present at
frequencies of about 2.0 Hz and 2.8 Hz. The figure also shows that due to
63
the fact that the frequency response is calculated only at discrete frequencies
in the filters, the singularities will not destroy the calculations. However,
it is recommended to cut off the frequency responses whenever the value is
larger than around 5. For practical use this means that, if |H(f )| ≥ 5 when
calculated, then |H(f )| should be valued 5. Furthermore, it is recommended
to place the singularities in a frequency range where the wave spectrum is
without significant energy, for example 3 times the peak frequency of the
spectrum. This can always be done by choosing appropriate values of Δx
and Δtf ilter , i.e. Δx smaller than a quarter of the shortest wave lengths.
In the examples the total elevation due to two superimposed sine waves is
described by Eq. 5.19, corresponding to 50 % reflection of the incident waves.
The signals are sampled with a frequency of 6.4 Hz. Fig. 5.5 illustrates the
functionality of the method, when f1 and f2 are both coinciding with some
frequencies of the discrete filter, i.e. n · Δf .
64
η (cm) ηI
η calc
3.0 ηx1
1.5
0.0
-1.5
-3.0
3.0
Error
1.5
0.0
-1.5
-3.0
0.0 5.0 10.0 15.0 20.0
time (sec.)
As expected the method is exact for signals consisting only of energy placed
at the discrete frequencies (Fig. 5.5), though it is seen that errors are present
during warm up of the filters.
The second example (Fig. 5.6) is identical to the first example except that f1
and f2 are not coinciding with frequencies in the digital filter i.e. f1 = 4.2Δf ,
f2 = 7.5Δf . It must be stressed that the output signal shown in Fig. 5.6
corresponds to the worst case situation, where the wave frequencies are placed
midway between filter frequencies.
65
η (cm) ηI
η calc
3.0 ηx1
1.5
0.0
-1.5
-3.0
3.0
Error
1.5
0.0
-1.5
-3.0
0.0 5.0 10.0 15.0 20.0
time (sec.)
One way to improve the results is to apply a tapering of the filter coefficients,
because the output from a digital filter is more stable in case the absolute
values of the filter coefficients are almost zero in both ends of the filter, Karl
(1989). Cosine tapering of the filter coefficients improves the accuracy of the
SIRW method as demonstrated in Fig. 5.7.
66
η (cm) ηI
η calc
3.0 ηx1
1.5
0.0
-1.5
-3.0
3.0
Error
1.5
0.0
-1.5
-3.0
0.0 5.0 10.0 15.0 20.0
time (sec.)
Figure 5.7: A comparison between ηI , η calc and ηx1 . The filters have been
cosine tapered. f1 = 4.2Δf , f2 = 7.5Δf .
First, the waves (incident part of the timeseries) were generated and sent
towards a spending permeable beach (slope 1:8) with low reflection (app. 5
%) in order to obtain a good estimate of the incident waves. Next, a reflecting
wall was mounted in the flume giving a fairly high reflection (app. 50 %) and
the same incident waves were reproduced by play back of the same digital
steering signal to the wave maker. Notice, that the incident wave fields are
identical only until re-reflection occurs. In Fig. 5.9 the output from the
SIRW-filters is compared with the incident waves measured in the case of
very low reflection.
67
Figure 5.8: Set-up for physical model tests.
η (cm) ηImeasured
η calc
3.0 ηxmeasured
1
1.5
0.0
-1.5
-3.0
Error (cm)
3.0
η calc − ηImeasured
1.5 η calc − ηxmeasured
1
0.0
-1.5
-3.0
10.0 15.0 20.0 25.0 30.0
time (sec.)
The specific part of the signals, where reflection is present but re-reflection
68
from the wave paddle is still not present, is shown. Two different estimates
of the incident waves are used, namely the measured elevation at gauge no
1 (ηxmeasured
1
) and the calculated elevation at gauge no 1 (η calc ). In the the
specific example the SIRW-method reduces the error (variance of the differ-
ence η calc −ηx1 ) from 30 % of the incident energy to 3 % of the incident energy.
5.3 Conclusions
69
Chapter 6
In the case of oblique waves, i.e. 2D-waves travelling along a line not perpen-
dicular to the reflecting structure, the previous derived methods can all can
be applied by assuming that the waves can be decomposed into two vectorial
components, being respectively ⊥ and || to the structure.
N
η(x, t) = n=1 aI,n cos(kn x − ωn t + ΦI,n ) +
N
n=1 aR,n cos(kn x + ωn t + ΦR,n ) (6.1)
71
Omitting index n and rewriting eq. (6.1) in cartesian coordinates:
The easiest way to solve the problem is to place the wave gauges on a line
perpendicular to the reflecting structure (y-coordinate ≡ 0 for all gauges) as
shown in fig 6. Eq. (6.2) will then simplify to:
72
Which shows the with this positioning of the gauges the normal methods
presented in the previous two chapters can be used when the geometrical
gauge distances are modified by the factor cos(θ).
Please notice that near the structure edge-waves will exist. These waves
travel along the structure and can destroy the calculations because they are
not included in the modelling of the waves.
73
Chapter 7
For a two dimensional sea state the elevation was assumed to be a summation
of a number of wavelets as stated in e.g. eq. (4.1). In a three dimensional
case the corresponding expression is
N
M
η(x, t) = amn cos(kmn x − ωn t + Φmn ) (7.1)
n=1 m=1
where k is the wavenumber vector. Eq. (7.1) implies that for each pair of
discrete values of frequency and direction of travel there exist a long crested
wave propagating with these properties.
75
as the variance of a sinusoidal wave is half the amplitude squared. Hence
1 2
a = S(ωm , θn )ΔθΔω
2 mn
amn = 2S(ωm , θn )ΔθΔω (7.2)
N
M
η(x, t) = 2S(ωm , θn )ΔθΔω cos(kmn x − ωn t + Φmn ) (7.3)
n=1 m=1
! ∞ ! π
η(x, t) = cos(kx − ωt + Φ(ω, θ)) 2S(ω, θ)dθdω (7.4)
0 −π
Two wave gauges are considered. These are positioned as shown in fig. 7.
6 Gauge 2
e
r12
Direction of travel
β12
*
θ
`
e - x
Gauge 1
76
The elevations at gauge 1 and gauge 2, respectively, are
! ∞ ! π
η1 = η(x1 , t) = cos(kx1 − ωt + Φ(ω, θ)) 2S(ω, θ)dθdω
−π
! 0
∞! π
η2 = η(x2 , t) = cos(kx2 − ωt + Φ(ω, θ)) 2S(ω, θ)dθdω
−π
! 0
∞! π
= cos(kx1 − ωt + kr12 cos(θ − β12 ) + Φ(ω, θ)) ·
0 −π
2S(ω, θ)dθdω
Here it is necessary to specify the time argument, thus η1 (t) = η1 and η2 (t) =
η2 . Inserting into eq. (7.6) yields
! ! !
1 T ∞ π
Rη1 η2 (τ ) = cos(kx1 − ωt + Φ(ω, θ)) ·
T 0 0 −π
cos(kx1 − ω(t + τ ) + kr12 cos(θ − β12 ) + Φ(ω, θ)) ·
2S(ω, θ)dθdωdt (7.7)
leads to
! ! !
1 T ∞ π
Rη1 η2 (τ ) = [cos(kr12 cos(θ − β12 ) − ωτ )
T 0 0 −π
+ cos(2kx1 + kr12 cos(θ − β12 ) − 2ωt − ωτ + 2Φ(ω, θ))] ·
S(ω, θ)dθdωdt
! ∞! π
= cos(kr12 cos(θ − β12 ) − ωτ )S(ω, θ)dθdω
0 −π
77
which can be written as
! ∞ ! π
Rη1 η2 (τ ) = [cos(ωτ ) cos(kr12 cos(θ − β12 ))
0 −π
+ sin(ωτ ) sin(kr12 cos(θ − β12 ))] S(ω, θ)dθdω (7.8)
where c12 is the co-spectrum and q12 the quad-spectrum in the cross-spectrum
! π
c12 (ω) = S(ω, θ) cos(kr12 cos(θ − β12 ))dθ
!−π
π
q12 (ω) = S(ω, θ) sin(kr12 cos(θ − β12 ))dθ
−π
Inserting these expressions for the co- and quad-spectrum into eq. (7.10)
yields
! π
Sη1 η2 (ω) = S(ω, θ) exp(−ikr12 cos(θ − β12 ))dθ
!−π
π
Sη1 η2 (ω)
= H(ω, θ) exp(−ikr12 cos(θ − β12 ))dθ (7.11)
Sηη (ω) −π
Eq. (7.11) relates data, which can be computed from wave elevation mea-
surements, to the unknown directional spreading function. The geometry is
represented by r12 and β12 .
Eq. (7.11) is the basic relation that was in demand. It cannot be solved
analytically. Therefore methods have been proposed in order to make a
reliable estimate.
Common for all methods is that they make assumptions about the shape
of the directional spreading function. That may either be a parameterized
78
analytic expression or a number of discrete values making a step curve. This
initial assumption is very important to the ability of the particular method.
The next step is to fit the unknown parameters to the measured data. The
reliability of the fit depends on the reliability of the measured data but also on
the number of coefficients to fit, i.e. the more coefficients the less reliability.
This is a problem as a high resolution requires many coefficients.
In it most simple form MLM applies for simultaneous wave elevations, but
by use of transfer functions it can readily be extended to apply for various
measurements. The MLM was originally presented by Capon (1969), but has
since been modified by several authors. Especially the papers by Davis and
Regier (1977) and Isobe et al. (1984) are of importance.
k = 2π/L
L = L0 tanh(kd)
N
N
S̃(ω, θ) = αmn (ω, θ)Smn (ω) (7.12)
m=1 n=1
Then using
! 2π
Smn (ω) = exp(−ik(xn − xm ))S(ω, θ)dθ (7.13)
0
79
eq. (7.12) leads to
N
N ! 2π
S̃(ω, θ) = αmn (ω, θ) exp(−ik (xn − xm ))S(ω, θ )dθ(7.14)
m=1 n=1 0
$ %
k cos θ
where k = .
k sin θ
! 2π
S̃(ω, θ) = w(ω, θ, θ )S(ω, θ )dθ (7.15)
0
where
N
N
w(ω, θ, θ ) = αmn (ω, θ) exp(−ik (xn − xm )) (7.16)
m=1 n=1
From eq. (7.15) it is seen that the estimated directional spectrum is a convo-
lution of the true directional spectrum and the window function w(ω, θ, θ )
given by eq. (7.16).
80
Inserting this into eq. (7.12) and eq. (7.16) yields
N
N
S̃(ω, θ) = γm (ω, θ)γn∗ (ω, θ)Smn (ω) (7.18)
m=1 n=1
N N
w(ω, θ, θ ) = γm (ω, θ)γn∗ (ω, θ) exp(−ik (xn − xm ))
m=1 n=1
N N
= γm (ω, θ)γn∗ (ω, θ) exp(ik xm ) exp(−ik xn )
m=1 n=1
N
N
= γm (ω, θ) exp(ik xm ) γn∗ (ω, θ) exp(−ik xn )
m=1 n=1
N N
∗
= γm (ω, θ) exp(ik xm ) (γn (ω, θ) exp(ik xn ))
m=1 n=1
N
= | γm (ω, θ) exp(ik xm )|2 (7.19)
m=1
The maximum value of eq. (7.20) is equal to the maximum eigenvalue of the
matrix S −1 T . Eq. (7.20) is then the Rayleigh quotient of γ T (T − λS)γ ∗ = 0
81
whereas
γT T γ∗
maximum = λmax (7.23)
γ T Sγ ∗
−1
where λmax is the maximum eigenvalue. Thus by multiplication of γ T and
S −1 and substitution of T
γT T γ∗ = λmax γ T Sγ ∗
T γ∗ = λmax Sγ ∗
S −1 T γ ∗ = λmax γ ∗
S −1 γ ∗o γ To γ ∗ = λmax γ ∗
γ To S −1 γ ∗o γ To γ ∗ = λmax γ To γ ∗ (7.24)
Thus
γ To S −1 γ ∗o = λmax (7.25)
S̃(ω, θ) ∝ 1/λmax
& '−1
S̃(ω, θ) = κ γ To S −1 γ ∗o (7.26)
The terms in the RHS of eq. (7.26) are all known when a sample has been
carried out. S is the cross-spectrum matrix and is calculated from the time
series. γo is dependent only on the wavenumber vector and the geometry.
The factor κ is used in order to achieve the correct variance (i.e. the mea-
sured variance). While the MLM (and BDM) provides reasonable results for
directional wave spectra containing incident waves only the methods becomes
less reliable when reflections occur. This problem happen because there is a
corelation of the phases of the incident and reflected waves. This problem
82
arises in most methods.
N
ηp (x, t) = (Ap,l cos ωl t + Bp,l sin ωl t) , p = 1, 2, ..., M (7.27)
l=1
where ωl = l 2π
T
and T is the length of the time series. The coefficients Ap,l
83
and Bp,l are given as the stochastic integrals
! T
2 2
Ap,l = ηp (xp , t) cos ωl t dt l = 1, 2, . . . , N p = 1, 2, . . .(7.28)
,M
T − T2
! T
2 2
Bp,l = ηp (xp , t) sin ωl t dt l = 1, 2, . . . N p = 1, 2, . . . (7.29)
,M
T − T2
From eq. (7.28) and eq. (7.29) it is seen that all the coefficients Ap,l and Bp,l
are joint Gaussian distributed stochastic variables.
In the following only the l’th components are considered. The primary
aim is to determine the joint distribution of the coefficients Ap,l and Bp,l
p = 1, 2, . . . , M , at the frequency wl .
The reason for limiting the analysis to a single frequency at a time is that the
unknown parameters in S(ω, θ) to be estimated may be frequency dependent.
AT = [A1 A2 . . . AM B1 B2 . . . BM ]
= [A1 . . . AM AM +1 . . . A2M ] (7.30)
B E
κ A AT = (7.31)
ET D
84
From eq. (7.28) and eq. (7.29) the following results are found
! 2
2 T
E [Ap,l ] = E [ηp (xp , t)] cos ωl t dt = 0 , p = 1, 2, . . . , M (7.32)
T − T2
and
! 2
2 T
E [Bp,l ] = E [ηp (xp , t)] sin ωl t dt = 0 , p = 1, 2, . . . , M (7.33)
T − T2
i.e. E [A] = 0.
In the following the cross covariance function matrix E A AT will be de-
termined. As an example the submatrix B in eq. (7.31) is considered.
( ! 2 ! 2
)
2 T 2 T
κAm,l An,l (t1 , t2 ) = E ηm (t1 ) cos (ωl t1 ) dt1 · ηn (t2 ) cos (ωl t2 ) dt2
T − T2 T − T2
! 2 ! 2
4 T T
= E [ηm (t1 ) ηn (t2 )] cos (ωl t1 ) cos (ωl t2 ) dt1 dt2
T2 − T2 − T2
! 2 ! 2
4 T T
= κmn (t2 − t1 ) cos (ωl t1 ) cos (ωl t2 ) dt1 dt2 (7.34)
T2 − T2 − T2
It is seen, that κAm n depends on the cross covariance between the elevation
cj Acj
processes ηm (t1 ) and ηn (t2 ). The cross covariance matrix κmn (τ ) is related
to the cross spectral density matrix Smn (ω) in terms of the Wiener-Khinchine
relation
! ∞
Gmn (ω) = 2Smn (ω) = 2 κmn (τ ) e−iωτ dτ
−∞
! ∞ ! ∞
= 2 κmn (τ ) cos ωτ dτ − i · 2 κmn (τ ) sin ωτ dτ
−∞ −∞
85
where Gmn (ω) is a one-sided spectrum.
! ∞
Cmn (ω) = 2 κmn (τ ) cos ωτ dτ
−∞
and
! ∞
Qmn (ω) = 2 κmn (τ ) sin ωτ dτ
−∞
1 Δω
κAm,l An,l (t1 , t2 ) = κAm,l An,l (τ ) = Cmn (ωl ) = Cmn (ωl ) = Bmn
(7.36)
T 2π
Using the same procedure the elements E and D of eq. (7.31) can be found
and after some calculation the following result is obtained
Δω C Q Δω
κAAT (ωl ) = = · Ω (ωl ) (7.37)
2π −Q C 2π
So far expressions have been established for the mean value vector (eq. (7.32)
and eq. (7.33)) and the cross covariance matrix eq. (7.37) of the vector A.
These relations were established at a given frequency ωl . Besides the fre-
quency the wave pattern is also characterised by a direction of travel. The
dependence of the frequency and the direction of travel is expressed in terms
of the directional spectrum S(ω, θ). The following relation exists between
86
the one-sided autospectral density G(ω) and S(ω, θ)
! 2π
G(ω) = S(ω, θ) dθ (7.38)
0
2π
Gmn (ω) = 0 S(ω, θ) { exp(ik (xm − xn ) + r(ω, θ) exp(ik (xm − xn T ))
+ r(ω, θ) exp(ik (xm T − xn )) + r2 (ω, θ) exp(ik (xm − xn ) T )}
(7.39)
dθ
If the directional spectrum S(ω, θ) was known eq. (7.39) may be used to cal-
culate Gmn (ω), e.g. by numerical integration.
Based on eq. (7.35) C and Q are identified as the real and the imaginary
parts of G. Finally κA AT (ωl ) can be determined from eq. (7.37). However,
S(ω, θ) is generally unknown. In the following a method will be presented
which can be used to determine a directional spectrum expressed in standard
form in terms of some unknown parameters. The method is known as the
Maximum Likelihood (ML) method. As expressed earlier the elements in
A have a joint normal distribution. The general expression for a density
function of a vector A with 2M joint Gaussian elements having a mean
value vector E(A) = 0 is
1 1 T −1
pA (a) = &√ '2M # # 1 exp − 2 a κAAT (τ )a (7.41)
2π #κ T (τ )#
2
AA
# #
where #κAAT # and κ−1 T are the determinant and the inverse matrix of
AA
87
κAAT respectively.
If S(ω, θ) was given eq. (7.41) could be used to calculate the probability of
the observed realisation a of A, where a represents the actual Fourier coef-
ficients obtained from a given time-series. Since S(ω, θ), or some parameters
in S(ω, θ), are unknown, a Likelihood function, L(·), will be formulated.
Expressed in terms of the Likelihood function the unknown parameters in
Sη (ω, θ) are determined as the values corresponding to the maximum value
of L.
where
2π
L
Ω̃lh = aml amh (7.43)
Δω · L m=1
Eq. (7.42) represents the probability of obtaining exactly the estimates a(l) , l =
1, 2, . . . , L. The unknown quantity in eq. (7.42) is G (or S(ω, θ)).
88
7.3 The Bayesian Directional spectrum esti-
mation Method
K $
1 (l − 1)Δθ ≤ θ < kΔθ
H(ω, θ) exp(xl (ω))Il (θ) Il = (7.45)
0 otherwise
l=1
! 2π
K
Smn (ω)
exp(xl (ω))Il (θ) exp(−ikrmn cos(θ − βmn ))dθ
S(ω) 0 l=1
K
exp(xl (ω)) exp(−ikrmn cos(θl − βmn ))Δθ (7.47)
l=1
89
eq. (7.47) can be rearranged to
K
Sj (ω) = exp(xl (ω))αjl (ω) + εj (ω) j = 1, 2, . . . , M (7.48)
l=1
where an error εmn (ω) of the cross-spectrum Smn (ω) has been added, and
K
ε j = Sj − exp(xl )αjl : N (0; σ 2 )
l=1
1 ε2j
p(εj ) = √ exp − 2
2πσ 2σ
90
The likelihood function is then
,
2N
1 ε2j
L(ε1 , ε2 , ..., ε2N ; σ) = √ exp − 2
j=1
2πσ 2σ
2N
1 ε2j
= exp −
(2πσ 2 )N j=1
2σ 2
⎛ 2
⎞
1 −1 2N K
= 2 N
exp ⎝ 2 Sj − ⎠
exp(xl )αjl(7.52)
(2πσ ) 2σ j=1 l=1
K
(xl − 2xl−1 + xl−2 )2 (7.54)
l=1
Maximising the likelihood eq. (7.52) and minimising eq. (7.54) lead to the
estimate which also maximises
u2
K
ln L(x1 , x2 , ..., xK ; σ) − (xl − 2xl−1 + xl−2 )2
2σ 2 l=1
u2
K
L(x1 , x2 , ..., xK ; σ) exp − 2 (xl − 2xl−1 + 2xl−2 )2 (7.55)
2σ l=1
91
distribution of x = (x1 , x2 , ..., xK )
u2
K K
u
2
p(x|u , σ ) =2
√ exp − 2 (xl − 2xl−1 + xl−2 )2 (7.56)
2πσ 2σ l=1
2
1
2N K
u 2
K
Sj − exp(xl )αjl + 2 (xl − 2xl−1 + xl−2 )2
2σ 2 j=1 l=1
2σ l=1
2
2N
K
K
Sj − exp(xl )αjl +u 2
(xl − 2xl−1 + xl−2 )2 (7.58)
j=1 l=1 l=1
The BDM turns out to be a very useful and reliable method for estimation of
directional wave spectra. But some inaccuracy arises when the wave samples
contain reflections, which in worst cases results in non-converging results.
Similarly to the MLM it is possible to create additional constraints to the
shape of the directional spreading function or make some assumptions about
location and direction of the reflector.
92
7.4 Tests of the presented methods
In principle all methods for estimating reflected waves should be tested upon:
In this section the performance of the methods are evaluated with numerical
data in situations where the reflection is known, and no possible noise (errors)
on the data are included.Moreover, the application of the BDM method to
data from physical model tests is demonstrated.
In the numerical tests performed, the generated data were simultaneous re-
alizations of surface elevation time series recorded in a CERC5 wave gauge
array with a radius of 1.0 m positioned at a water depth of 4.0 m, see fig. 7.2.
Two tests were performed: One in which no reflection occurred and one in
which a wall with a reflection coefficient of r=0.5 was positioned at x=0.0
m, see fig. 7.2.
93
Figure 7.2: Wave gauge arrangement for numerical tests.
In both tests, the incident wave fields were irregular waves corresponding to
a Pierson-Moskowitz spectrum (fp = 0.5Hz, Hs = 0.5m) with a Mitsuyasu-
type spreading function (θ0 = π/3, s = 6). A sample frequency of fs = 4Hz
was applied.
The estimated value of the auto spectral density Sη (ω) and the maximum
likelihood estimates of the parameters θ0 , s and r are given in fig. 7.3 and
fig. 7.4. For comparison, the target values are plotted.
In both tests, the estimated values of the parameters are in good agree-
ment with the target values. However, in the test involving reflection the
method has some problems separating incident and reflected waves at some
frequencies resulting in larger estimated values of the auto spectral density,
see fig. 7.4 compared to the results obtained without reflection, see fig. 7.3.
94
Autospectral density Peak wave direction
60m 6.3
50m 5.2
40m 4.2
teta [rad]
S [m*m]
30m 3.1
20m 2.1
10m 1.0
0m 0.0
0.3 0.4 0.6 0.7 0.8 1.0 1.1 0.3 0.4 0.6 0.7 0.8 1.0 1.1
frequency [Hz] frequency [Hz]
4.0 0.40
3.0 0.30
2.0 0.20
1.0 0.10
0.0 0.00
0.3 0.4 0.6 0.7 0.8 1.0 1.1 0.3 0.4 0.6 0.7 0.8 1.0 1.1
frequency [Hz] frequency [Hz]
95
Autospectral density Peak wave direction
60m 6.3
50m 5.2
40m 4.2
teta [rad]
S [m*m]
30m 3.1
20m 2.1
10m 1.0
0m 0.0
0.3 0.4 0.6 0.7 0.8 1.0 1.1 0.3 0.4 0.6 0.7 0.8 1.0 1.1
frequency [Hz] frequency [Hz]
4.0 0.40
3.0 0.30
2.0 0.20
1.0 0.10
0.0 0.00
0.3 0.4 0.6 0.7 0.8 1.0 1.1 0.3 0.4 0.6 0.7 0.8 1.0 1.1
frequency [Hz] frequency [Hz]
96
7.4.2 BDM on numerical data
The BDM method has been tested utilising numerical simulations of an uni-
modal wave field and a bi-modal wave field.
Only the bi-directional wave field test is shown, see fig. 7.5. In this case
it is seen that both peaks are estimated well. Though the narrow peak is
estimated most accurate.
0.45
0.4 target
BDM
0.35
0.3
0.25
H(ω, θ)
0.2
0.15
0.1
0.05
0
0 50 100 150 200 250 300 350
θ
Data. Test 1.
Date of sample 070893
Depth of water 200m
Radius of array 26m
Sample frequency 4Hz
Duration 720sec.
Quantity Elevation
Autospectrum P–M,
Tp = 10s, Hs = 4.0m
Directional spreading function Mits.,
s1 = 5, s2 = 8, θ1 = −90◦ , θ2 = 90◦
97
7.4.3 BDM on Model Test Data
In fig. 7.6 a test from Aalborg University’s 3D wave basin is shown. The
main direction of the incident waves is 290 degrees and the mean direction
of the reflected waves should be 70 degrees.
0.8
generated
0.7 BDM
0.6
0.5
H(ω, θ) 0.4
0.3
0.2
0.1
0
0 50 100 150 200 250 300 350
θ
Data. Test 2.
Date of sample 090193
Depth of water 0.6m
Radius of array 0.25m
Sample frequency 10Hz
Duration 720sec.
Quantity Elevation
Autospectrum JONSWAP,
Tp = 0.8s, Hs = 0.1m
Directional spreading function Mits.,
truncated s1 = 3, θ1 = −90◦
The BDM method estimates the incident wave energy to be spread over
a wider range. This is reasonable considering the conditions in the basin,
i.e. limitations due to boundary conditions. The mean direction of the re-
flected waves though deviate about 20 degrees.
98
7.5 Conclusion
Several different methods for estimating reflection have been presented and
tested in the previous chapters.
The frequency domain methods give the incident wave spectrum and the
reflected wave spectrum. The time domain methods give the incident waves
as function of time.
Some of the methods take into account the 3-dimensionality of the waves. If
waves on location are 3-dimensional it is necessary to use a method for 3-
dimensional waves, though because the methods for estimating reflection in
3-dimensional waves introduce the directional spreading function with many
degrees of freedom, they become less statistically reliable.
99
Chapter 8
Wave Groups
In sea wave recordings, group formations of high waves occur from time to
time. This phenomenon corresponds to a non-zero correlation between suc-
cessive waves. Information concerning this correlation is of importance when
reproducing waves in the laboratory in order to determine the response of
the modelled structure. Normally, irregular waves are reproduced in accor-
dance with a specific energy spectrum solely defining the distribution of the
variances. The grouping of waves is determined by the distribution of the
phases. Hitherto, independence between successive waves have been applied
and the phases are treated as independent random variables, each with a
uniform probability density on the interval [0;2π] leading to a sea surface
that is Gaussian distributed. However, if the waves during wave propagation
become more non-linear there will be some coupling and thus dependence of
the phases of the component waves at different frequencies, which eventually
will modify the wave grouping.
To illustrate the effect of randomly assigned phases two wave trains are gener-
ated from the same energy spectrum. These two wave conditions are depicted
in Figure 8.1.
Burcharth (1979) and Johnson et al. (1978) found that the wave grouping
101
5.0
2.5
Elevation (m)
0.0
0 100 200 300 400
30.0
-2.5
25.0
S(f)/m0 (m2*sec/m2)
-5.0
20.0
time (sec.)
15.0
10.0
5.0
5.0
2.5
0.0
Elevation (m)
0.00 0.05 0.10 0.15 0.20 0.25 0.30
frequency (Hz) 0.0
0 100 200 300 400
-2.5
-5.0
time (sec.)
Figure 8.1: Wave energy spectrum and generated grouped and non-grouped
wave trains.
In irregular seas, model tests by Spangenberg (1980) showed that the wave
grouping has a significant influence on the slow drift motion of moored plat-
forms and vessels. This influence might be explained by the fact that the
period of the slow drift oscillations practically corresponds to the wave group
period where the wave grouping is pronounced.
102
preted as the square envelope.
N
η(t) = cn cos(ωn t + εn ) (8.1)
n=1
N
N
2
η (t) = cn cm cos(ωn t + εn ) cos(ωm t + εm ) (8.2)
n=1 m=1
N
N
1
= {cn cm ( cos((ωn + ωm )t + (εn + εm )) +
n=1 m=1
2
1
cos((ωn − ωm )t + (εn − εm )))} (8.3)
2
Equation (8.3) represents a splitting of η 2 (t) into a slowly varying part (rep-
resented by the difference-frequencies) and a more rapid oscillating part (rep-
resented by the summation-frequencies).
1 2 1 2
N N
2
η (t) = c + c cos(2ωn t + 2εn )
2 n=1 n 2 n=1 n
N
N
+ cn cm cos((ωn + ωm )t + (εn + εm ))
n=1 m=n+1
N
N
+ cn cm cos((ωn − ωm )t + (εn − εm ))) (8.4)
n=1 m=n+1
The four terms on the right-hand side of equation (8.4) are identified as fol-
lows: The first term consists of a constant off-set component. The second and
third term constitutes the superharmonic components, i.e. the summation-
frequency terms, and the fourth term constitutes the subharmonic compo-
nents, i.e. the difference-frequency terms. It is the latter that describes the
103
slowly varying part of the squared time signal and the term which may be
interpreted as the square envelope. By means of Bartlett filtering the su-
perharmonic components on the right-hand side of equation (8.4) may be
filtered out after subtraction of the constant off-set as done by Funke and
Mansard (1979).
Funke and Mansard denoted the filtered square of the time signal the SIWEH
(Smoothed Instantaneous Wave Energy History) function as the function
provides a measure of the instantaneous wave energy in the time signal.
The effect of the Bartlett filtering corresponds to a digital low pass filtering
and the efficiency of the SIWEH analysis can best be interpreted by examina-
tion of the energy spectrum of the stochastic process in (8.1) and the energy
spectrum of the squared process.
30.0 30.0
25.0 25.0
S(f)/m0 (m2*sec/m2)
S(f)/m0 (m2*sec/m2)
20.0 20.0
15.0 15.0
10.0 10.0
5.0 5.0
0.0 0.0
0.00 0.10 0.20 0.30 0.40 0.50 0.60 0.00 0.10 0.20 0.30 0.40 0.50 0.60
frequency (Hz) frequency (Hz)
Figure 8.2: a) JONSWAP energy spectrum for a linear stochastic process and
b) energy spectrum for the squared process.
From figure 8.2 it is understood that the SIWEH analysis does not exactly
isolate the slowly varying part; also contributions from the superharmonic
components occur and not the complete amount of energy from the subhar-
monic components is included. Only when the process is narrow-banded does
the SIWEH analysis perform well but as the process becomes more and more
broad-banded the SIWEH function is a poor estimator of the wave envelope,
see Hupspeth and Medina (1988).
104
8.2 Hilbert Transform Technique
From the sea surface elevation η(t) a conjugate signal η̂(t) is uniquely ob-
tained by shifting the phase of each elementary harmonic component of η(t)
by ± π2 . When the phase angles of all components of a given signal are shifted
± π2 , the resulting function η̂(t) is known as the Hilbert transform of the orig-
inal signal η(t). The Hilbert transform is defined by
! ∞
1 η(t)
η̂(t) = dτ (8.5)
π −∞ t−τ
From the definition of the Hilbert transform it is noted that η̂(t) is sim-
ply the convolution of η(t) with a linear filter with the impulse response
1 1
function h(t) = πt . Since a convolution of two functions in the time do-
main are transformed into a multiplication of their Fourier transforms in
the frequency domain 2 a frequency response function H(f ) is related to
the impulse response function. The frequency response function provides an
equally characterization of the linear time-invariant input and output system
in (8.5) and does furthermore visualize the effect of the Hilbert transform op-
eration. Through the Fourier transform the frequency response of the Hilbert
transformer becomes
⎧
1 ⎨ −i f >0
H(f ) = F[ ] = −i sgn(f ) = 0 f =0 (8.6)
πt ⎩
i f <0
The gain of this frequency response function is re2 (H(f )) + im2 (H(f ))
resulting in unity in magnitude, and thus, the amplitudes of the signal does
not change. The phase angle is arctan( im
re
(H(f ))
(Hf ))
) resulting in a phase angle of
− 2 for f > 0 and + 2 for f < 0. Such a system is denoted an ideal 90-degree
π π
phase shifter.
N
η̂(t) = cn sin(ωn t + εn ) (8.7)
n=1
Associated with the Hilbert transform is the complex analytical signal defined
1
∞
The convolution of the functions g(t) ∗ h(t) is: g(t) ∗ h(t) ≡ −∞
g(τ )h(t − τ )dτ
2
The convolution theorem: g(t) ∗ h(t) ⇔ G(f )H(f )
105
from the original signal η(t) and the Hilbert transform η̂(t)
where the envelope or the modulation | η̃(t) | = η 2 (t) + η̂ 2 (t) and the
η̂(t)
associated phase ψ(t) = arctan( η(t) ). The properties of the Hilbert transform
operation entail that the slowly varying difference-frequency terms in the
second order expression η 2 (t) are separated mathematically by the expression
where η̃ ∗ (t) = the complex conjugate and E(t) = the square wave envelope
function.
In order to visualize the effect of the defined envelope function the Hilbert
transform of the sea surface elevation is squared and rewritten by use of
trigonometry and symmetry of the double summation similar to η 2 (t)
N
N
2
η̂ (t) = cn cm sin(ωn t + εn ) sin(ωm t + εm ) (8.10)
n=1 m=1
N N
1
= {cn cm ( cos((ωn − ωm )t + (εn − εm )) −
n=1 m=1
2
1
cos((ωn + ωm )t + (εn + εm )))} (8.11)
2
1 2 1 2
N N
= cn − c cos(2ωn t + 2εn )
2 n=1 2 n=1 n
N
N
− cn cm cos((ωn + ωm )t + (εn + εm ))
n=1 m=n+1
N
N
+ cn cm cos((ωn − ωm )t + (εn − εm ))) (8.12)
n=1 m=n+1
Remembering that the squared time signal is given by (8.4), the square wave
envelope function, according to (8.9), then becomes
N
N
N
E(t) = c2n +2 cn cm cos((ωn − ωm )t + (εn − εm )) (8.13)
n=1 n=1 m=n+1
106
Introducing √12 in the complex analytical signal η̃(t) = √12 (η(t) + iη̂(t)) leads
to the definition of an envelope function which may be interpreted as half
the square envelope.
1
E(t) =| η̃(t) |2 = (η 2 (t) + η̂ 2 (t)) (8.14)
2
This envelope function isolates exactly the slowly varying part of the squared
time signal plus the constant off-set similar to what approximately is achieved
by the SIWEH analysis.
The present method seems to be more convenient than the SIWEH analysis
and it does not require the narrow-band spectrum assumption. The disad-
vantage of this method is however that the sea surface must be described by
a linear model.
Nc /2−1
N c −1
107
to obtain a least-square fit of the coefficients. Opposite the centered format
definition of the Fourier transformation, the FFT is based on a one-sided
format
1 1
Nc −1 Nc −1
2πjk
ck = Xj exp(iωj kΔt) = Xj exp(i ) (8.16)
Nc j=0 Nc j=0 Nc
Nc
τ= Δt (8.17)
2
2πj
ψτ = τ ωj = τ = πj (8.18)
Nc Δt
To compensate for the phase delay the original frequency response function
given by (8.6) only needs to be multiplied by a linear phase shift operator
exp(−iπk) and Xj might be interpreted as
where G(fj ) is the gain of the input amplitude to equal the output amplitude
and ψj − πj is the phase difference between the input and the output signal.
108
1.5 2.0
1.0
1.0
0.5
Phase
Gain
0.0 0.0
0.0 0.2 0.4 0.6 0.8 1.0 0.0 0.2 0.4 0.6 0.8 1.0
-0.5
-1.0
-1.0
-1.5 -2.0
frequency (Hz) frequency (Hz)
Figure 8.3: Gain and phase characteristic of linear FIR Hilbert filter with a
filter length Nc = 64 and fs = 1.0 Hz.
Original signal
Theoretical Hilbert transform
10.0 Approximated Hilbert transform
7.5
5.0
Signal
2.5
0.0
0 50 100 150 200
-2.5
-5.0
time (sec.)
To illustrate the envelope function, E(t) is plotted together with half the
squared elevation in figure 8.5 for a time signal generated from the JONSWAP
spectrum.
109
Original signal
Half the square of signal
10.0 Half the square envelope
7.5
5.0
Signal
2.5
0.0
0 50 100 150 200
-2.5
-5.0
time (sec.)
Figure 8.5: Comparison of half the square envelope E(t) and 12 η 2 (t) for signal
generated from a JONSWAP spectrum, fp = 0.1 Hz, γ = 3.3, and Nc = 64.
σ[E(t)]
GF = (8.21)
σ 2 [η(t)]
Instead of computing one value of the groupiness factor over the complete
length of the time signal, the groupiness factor can be evaluated as instanta-
neous values by computing an average groupiness factor over a time moving
window. The length of the window in time is dependent on the desired degree
of smoothing of the computed groupiness factor function.
In figure 8.6 to figure 8.9 the groupiness factor function is plotted for both
a narrow-banded and a broad-banded JONSWAP spectrum for two different
window sizes.
110
Original signal
Half the square envelope
10.0 Groupiness factor
7.5
5.0
Signal
2.5
0.0
0 100 200 300 400
-2.5
-5.0
time (sec.)
Figure 8.6: Groupiness factor function GF(t) for signal generated from JON-
SWAP spectrum, fp = 0.1 Hz, fs = 1.0 Hz, γ = 10.0, Nc = 64, and window
size = Tm .
Original signal
Half the square envelope
10.0 Groupiness factor
7.5
5.0
Signal
2.5
0.0
0 100 200 300 400
-2.5
-5.0
time (sec.)
Figure 8.7: Groupiness factor function GF(t) for signal generated from JON-
SWAP spectrum, fp = 0.1 Hz, fs = 1.0 Hz, γ = 10.0, Nc = 64, and window
size = 3 Tm .
111
Original signal
Half the square envelope
10.0 Groupiness factor
7.5
5.0
Signal
2.5
0.0
0 100 200 300 400
-2.5
-5.0
time (sec.)
Figure 8.8: Groupiness factor function GF(t) for signal generated from JON-
SWAP spectrum, fp = 0.1 Hz, fs = 1.0 Hz, γ = 1.0, Nc = 64, and window
size = Tm .
Original signal
Half the square envelope
10.0 Groupiness factor
7.5
5.0
Signal
2.5
0.0
0 100 200 300 400
-2.5
-5.0
time (sec.)
Figure 8.9: Groupiness factor function GF(t) for signal generated from JON-
SWAP spectrum, fp = 0.1 Hz, fs = 1.0 Hz, γ = 1.0, Nc = 64, and window
size = 3 Tm .
112
The method can easily be extended to a three-dimensional motion but a
physical interpretation of the more slowly varying part must then be revised.
113
Chapter 9
References
115
Frigaard, P., Helm–Petersen, J., Klopman, G., Stansberg, C.T., Benoit, M.,
Briggs, J., Miles, M., Santas, J., Schäffer, H.A. and Hawkes, P.J., 1997.
IAHR List of Sea State Parameters – an update for multidirectional waves.
IAHR Seminar Multidirectional Waves and their Interaction with Structures,
IAHR Congress, San Francisco.
116
”Estimation of Directional Spectrum Expressed in Standard Form”
Proc. 22nd Int. Conf on Coastal Eng., pp. 647-660.
117
Ochi, Michel K. 1998
”Ocean Waves - The Stochastic Approach”
Cambridge University Press, Cambridge. ISBN: 0-521-56378
118
ISSN 1901-7286
DCE Lecture Notes No. 33