The Hilbert-Huang Transform in Engineering (Norden E. Huang, Nii O. Attoh-Okine)
The Hilbert-Huang Transform in Engineering (Norden E. Huang, Nii O. Attoh-Okine)
The
Hilbert-Huang
Transform
in
Engineering
The
Hilbert-Huang
Transform
in
Engineering
Edited by
Norden Huang
Nii O. Attoh-Okine
A CRC title, part of the Taylor & Francis imprint, a member of the
Taylor & Francis Group, the academic division of T&F Informa plc.
Published in 2005 by
CRC Press
Taylor & Francis Group
6000 Broken Sound Parkway NW, Suite 300
Boca Raton, FL 33487-2742
Preface
Data analysis serves two purposes: to determine the parameters needed to construct
a model and to confirm that the model constructed actually represents the phenom-
enon. Unfortunately, the data — whether from physical measurements or numerical
modeling — most likely will have one or more of the following problems:
These problems dictate how the data can be analyzed and interpreted.
This book describes the formulation and application of the Hilbert-Huang trans-
form (HHT) in various areas of engineering, including structural, seismic, and ocean
engineering.
The primary objective of this book is to present the theory of the Hilbert-Huang
Transform (HHT) and its application to engineering. The presentation of the book
is such that it can be used as both a reference and a teaching text. The authors of
the individual chapters provide a strong theoretical background and some new
developments before addressing their specific application. This approach demon-
strates the versatility of the HHT.
The book comprises 13 chapters, covering more than 300 pages. These chapters
were written by 30 invited experts from 6 different countries.
The book begins with an introduction and some recent developments in HHT.
Chapter 2 uses HHT to interpret nonlinear wave systems and provides a compre-
hensive analysis on the assessment of rogue waves. Chapter 3 discusses HHT
applications in oceanography and ocean-atmosphere remote sensing data and pre-
sents some examples of these applications. Chapter 4 presents a comparison of the
energy flux computation for shooting waves of HHT and wavelet analysis techniques.
In Chapter 5, HHT is applied to nearshore waves, and the results are compared to
field data. In Chapter 6, the author uses HHT to characterize the underwater elec-
tromagnetic environment and to identify transient manmade electromagnetic distur-
bances, where the HHT was able to act as a filter effectively discriminating different
dipole components. Chapter 7 presents a comparative analysis of HHT and wavelet
transforms applied to turbulent open channel flow data.
In Chapter 8, nonlinear soil amplification is quantified by using HHT. Chapter
9 extends the application of HHT to nonstationary random processes. Chapter 10
presents a comparative analysis of HHT wavelet and Fourier transforms in some
structural health-monitoring applications. In Chapter 11, HHT is applied to molec-
ular dynamics simulations. Chapter 12 presents a straightforward application of HHT
to decomposition of wave jumps. Chapter 13, the concluding chapter, presents
perspectives on the theory and practices of HHT. It attempts to review HHT appli-
cations in biomedical engineering, chemistry and chemical engineering, financial
engineering, meteorological and atmospheric studies, ocean engineering, seismic
studies, structural analysis, health monitoring, and system identification. It also
indicates some directions for future research.
One important feature of the book is the inclusion of a variety of modern topics.
The examples presented are real-life engineering problems, as well as problems that
can be useful for benchmarking new techniques.
The studies reported in this book clearly indicate an increasing interest in HHT
and analysis for diversified applications. These studies are expected to stimulate the
interest of other researchers around the world who are facing new challenges in new
theoretical studies and innovative applications.
Norden Huang
NASA
Nii O. Attoh-Okine
University of Delaware
Acknowledgments
The editors are grateful to the contributing authors. We also wish to express our
thanks to Tao Woolfe, B. J. Clark, and Michael Masiello of Taylor & Francis for
providing useful feedback and guiding the editors throughout the complex editorial
phase.
Norden E. Huang would like to thank Professors Theodore T. Y. Wu of the
California Institute of Technology, and Owen M. Phillips of the Johns Hopkins
University for their guidance and encouragement throughout the years, without
which the Hilbert-Huang Transform would not be what it is today.
Nii O. Attoh-Okine, the co-editor of the book, wishes to express his gratitude
to his parents, Madam Charkor Quaynor and Richard Attoh-Okine, for their support
and encouragement through the years.
Contributors
Nii O. Attoh-Okine Ping Gu
Civil Engineering Department Department of Civil Engineering
University of Delaware University of Illinois at Urbana-
Newark, Delaware Champaign
Urbana, Illinois
Brad Battista
Information Systems Laboratories, Inc. Norden E. Huang
San Diego, California Goddard Institute for Data Analysis
NASA Goddard Space Flight Center
Rodney R. Buntzen Greenbelt, Maryland
Information Systems Laboratories, Inc.
San Diego, California Paul A. Hwang
Oceanography Division
Marcus Dätig Naval Research Laboratory
Civil Engineering Department Stennis Space Center, Mississippi
Bergische University Wuppertal
Wuppertal, Germany Lide Jiang
Graduate College of Marine
Brian Dzwonkowski Studies
Graduate College of Marine Studies University of Delaware
University of Delaware Newark, Delaware
Newark, Delaware
Young-Heon Jo
Colin M. Edge Graduate College of Marine Studies
GlaxoSmithKline Pharmaceuticals University of Delaware
Harlow, United Kingdom Newark, Delaware
Contents
Chapter 1 Introduction to Hilbert-Huang Transform and Some
Recent Developments...........................................................................1
Norden E. Huang
1 Introduction to
Hilbert-Huang
Transform and Some
Recent Developments
Norden E. Huang
CONTENTS
1.1 INTRODUCTION
Hilbert-Huang transform (HHT) is the designated name for the result of empirical
mode decomposition (EMD) and the Hilbert spectral analysis (HSA) methods, which
were both introduced recently by Huang et al. (1996, 1998, 1999, and 2003),
specifically for analyzing data from nonlinear and nonstationary processes. Data
analysis is an indispensable step in understanding the physical processes, but tradi-
tionally the data analysis methods were dominated by Fourier-based analysis. The
problems of such an approach were discussed in detail by Huang et al. (1998). As
data analysis is important for both theoretical and experimental studies (for data is
the only real link between theory and reality), we desperately need new methods in
order to gain a deeper insight into the underlying processes that actually generate
the data. The method we really need should not be limited to linear and stationary
processes, and it should yield physically meaningful results.
The development of the HHT is motivated precisely by such needs: first, because
the natural physical processes are mostly nonlinear and nonstationary, there are very
limited options in data analysis methods that can correctly handle data from such
processes. The available methods are either for linear but nonstationary processes
(such as the wavelet analysis, Wagner-Ville, and various short-time Fourier spectro-
grams as summarized by Priestley [1988], Cohen [1995], Daubechies [1992], and
Flandrin [1999]) or for nonlinear but stationary and statistically deterministic pro-
cesses (such as the various phase plane representations and time-delayed imbedded
methods as summarized by Tong [1990], Diks [1997], and Kantz and Schreiber
[1997]). To examine data from real-world nonlinear, nonstationary, and stochastic
processes, we urgently need new approaches.
Second, the nonlinear processes need special treatment. Other than periodicity,
we want to learn the detailed dynamics in the processes from the data. One of the
typical characteristics of nonlinear processes, proposed by Huang et al. (1998), is
the intra-wave frequency modulation, which indicates that the instantaneous fre-
quency changes within one oscillation cycle. Let us examine a very simple nonlinear
system given by the nondissipative Duffing equation as
∂2 x
+ x + ε x 3 = γ cos ωt , (1.1)
∂t 2
∂2 x
+ x (1 + ε x 2 ) = γ cos ωt , (1.2)
∂t 2
where the symbols are defined as in Equation 1.1. Then the quantity within the
parentheses can be regarded as a variable spring constant or a variable pendulum
length. With this view, we can see that the frequency should be changing from
location to location and from time to time, even within one oscillation cycle.
As Huang et al. (1998) pointed out, this intra-frequency frequency variation is
the hallmark of nonlinear systems. In the past, there has been no clear way to depict
this intra-wave frequency variation by using Fourier-based analysis methods, except
to resort to the harmonics. Even by the classical Hamiltonian approach, in which
the frequency is defined as the rate of change of the Hamiltonian with respect to
the action, we still cannot gain any more insight, for the definition of the action is
an integration of generalized momentum along the generalized coordinates; there-
fore, there is no instantaneous value. Thus, the best we could do for any nonlinear
distorted waveform in the earlier approaches was to refer to harmonic distortions.
Harmonic distortion is, in fact, a rather poor alternative, for it is the result obtained
by imposing a linear structure on a nonlinear system. Consequently, the results may
make perfect mathematical sense, but at the same time they have absolutely no
physical meaning. The physically meaningful way to describe the system should be
in terms of the instantaneous frequency.
The easiest way to compute the instantaneous frequency is by the Hilbert trans-
form, through which we can find the complex conjugate, y(t), of any real valued
function x(t) of Lp class,
∞
x (τ)
∫ t − τ dτ ,
1
y(t ) = P (1.3)
π
−∞
in which the P indicates the principal value of the singular integral. With the Hilbert
transform, we have
where
( ) y
1/ 2
a(t ) = x 2 + y 2 ; θ(t ) = tan −1 . (1.5)
x
Here a is the instantaneous amplitude and θ is the phase function; thus the
instantaneous frequency is simply
dθ
ω= − . (1.6)
dt
FIGURE 1.1 (See color insert following page 20). Data of Length-of-Day measure the
deviation from the mean 24-hour-day.
FIGURE 1.2 (See color insert following page 20). Analytic function in complex phase
plane formed by the real data and it Hilbert Transform. It shows not apparent order. After the
EMD, the annual cycle is extract and plotted also.
FIGURE 1.3 The phase function of the analytic function based on the Length-of-Day data.
the data span. If we follow through the definition of the instantaneous frequency as
given in Equation 1.6 literally, we would have a totally nonsensical result, as given
in Figure 1.4, where the instantaneous frequency is equally likely to be positive or
negative. Unfortunately, this is exactly the procedure recommended by Hahn (1996).
To show how this should not be the case, let us consider the three curves given
by the following three expressions
x1 = sin ωt ;
x3 = 1.5 + sin ωt ;
shown in Figure 1.5. All three curves are perfect sine functions, but with the mean
displaced: for x1, its mean is exactly zero; for x2, its mean is moved up by half of
its amplitude; for x3, its mean is moved up by 1.5 times its amplitude. As a result,
the curve represented by x3 is totally above the zero reference axis. If we perform
the Hilbert transform to all three functions given in Equation 1.7, we would get
three different circles in the phase plane, with the centers of two of the circles
displaced by the amount of the added constants, as shown in Figure 1.6. Conse-
quently, the phase functions from the three circles will be different, as shown in
Figure 1.7: for x1, the phase function is a straight line; for x2, the phase function is
a wavy line, but the general trend still agrees with the straight line; for x3, the phase
function is also wavy, but the variation is always within ± π.
FIGURE 1.4 Instantaneous frequency obtained form derivative of the phase function without
decomposition first. The values are equally like to be positive as negative.
FIGURE 1.5 (See color insert following page 20). Model data to illustrate the fallacy of
the instantaneous frequency without decomposition.
FIGURE 1.6 (See color insert following page 20). The analytic function in complex
phase plane of the data given in Figure 1.5.
sin x
0.5 + sin x
1.5 + sin x
15
Phase Angle: radians
10
−5
0 100 200 300 400 500 600 700 800
Time: (100) second
FIGURE 1.7 (See color insert following page 20). Phase function of the model function
given in Figure 1.5.
FIGURE 1.8 (See color insert following page 20). Instantaneous frequency computed
from the cosine model functions consist of the identical cosine function with different dis-
placements.
Based on these phase functions, the instantaneous frequency values are very
different for the three expressions, as shown in Figure 1.8: for x1, the instantaneous
frequency is a constant value, which is exactly what we expected; for x2, the
instantaneous frequency is a variable curve with all positive values; for x3, the
instantaneous frequency is a highly variable curve fluctuating from positive to
negative values. Even if we are prepared to accept negative frequency, the result of
three different values for the same sine wave with only a displaced mean is very
unsettling: some of the results are, of course, nonsensical. The only meaningful
result is from the sine curve with a zero mean.
What went wrong was the fact that two of the curves do not have a zero mean,
or the envelopes of the curves are not symmetric with respect to the zero axis. Thus,
before performing the Hilbert transform, we have to preprocess the data. In the past,
any preprocessing usually consisted of band-pass filtering. For some of the data
from linear and stationary processes, this band-pass filtering method will give the
correct results. For data from nonlinear and nonstationary processes, however, the
band-pass filter will alter the characteristics of the filtered curve. The problem with
the filtering approach is that all the frequency domain filters are Fourier-based, which
means they are established under linear and stationary assumptions. When the data
are from nonlinear and nonstationary processes, such Fourier-based analysis will
surely generate spurious harmonics, which are mathematically necessary but phys-
ically meaningless, as discussed by Huang et al. (1998, 1999).
Considering these points leads us to this conclusion: The correct preprocessing
for data from nonlinear and nonstationary processes will have to be adaptive and
FIGURE 1.9 Test data to illustrate the procedures of Empirical Mode Decomposition also
known as sifting.
implemented in the time domain. The only method presently known to achieve this
is based on the Hilbert-Huang transform proposed by Huang et al. (1996, 1998,
1999, and 2003). This method is the subject of the next section.
x (t ) − m1 = h1 . (1.8)
FIGURE 1.10 (See color insert following page 20). The cubic spline upper and the lower
envelopes and their mean, m1.
This result is shown in Figure 1.11. The procedure of extracting an IMF is called
sifting. By construction, this PIMF, h1, should satisfy the definition of an IMF, but
the change of its reference frame from rectangular coordinate to a curvilinear one
can cause anomalies, as shown in Figure 1.11, where multi-extrema between suc-
cessive zero-crossings still existed. To eliminate such anomalies, the sifting process
has to be repeated as many times as necessary to eliminate all the riding waves. In
the subsequent sifting process steps, h1 is treated as the data. Then
where m11 is the mean of the upper and lower envelopes of h1. This process can be
repeated up to k times; then, h1k is given by
Each time the procedure is repeated, the mean moves closer to zero, as shown
in Figures 1.12a, b, and c. Theoretically, this step can go on for many iterations, but
each time, as the effects of the iterations make the mean approach zero, they also
make amplitude variations of the individual waves more even. Yet the variation of
the amplitude should represent the physical meaning of the processes. Thus this
iteration procedure, though serving the useful purpose of making the mean to be
zero, also drains the physical meaning out of the resulting components if carried
too far. Theoretically, if one insists on achieving a strictly zero mean, one would
Data and h1
10
h1
data
8
2
Amplitude
−2
−4
−6
−8
−10
200 250 300 350 400 450 500 550 600
Time: second
FIGURE 1.11 (See color insert following page 20). Comparison between data and h1 , as
given by Equation (1.9). Note most, but not all, riding waves are eliminated in h1.
2
Amplitude
−2
−4
−6
−8
−10
200 250 300 350 400 450 500 550 600
Time: second
FIGURE 1.12 (A) (See color insert following page 20). Repeat the sifting using h1 as
data.
2
Amplitude
−2
−4
−6
−8
−10
200 250 300 350 400 450 500 550 600
Time: second
FIGURE 1.12 (B) (See color insert following page 20). Repeat the sifting using h2 as data.
FIGURE 1.12 (C) (See color insert following page 20). After 12 iterations, the first
Intrinsic Mode Function is found.
∑h
2
k −1 (t ) − hk (t )
SD = t =0
T
; (1.11)
∑h
t =0
2
k −1 (t )
then the sifting will stop when SD is smaller than a preassigned value. This definition
is a slight modification from the original one proposed by Huang et al. (1998), where
the SD was defined simply as
2
hk −1 (t ) − hk (t )
T
SD = ∑
t =0
hk2−1 (t )
. (1.12)
The shortcoming of this old definition as given in Equation 1.12 is that the value
of SD can be dominated by local small values of hk–1, while the definition given in
Equation 1.11 sums up all the contributions over the whole duration of the data.
Even with this modification, there is still a problem with this seemingly mathemat-
ically sound approach: in this definition, the important criterion that the number of
extrema has to equal the number of zero-crossings has not been checked. To over-
come this practical difficulty, Huang et al. (1999, 2003) proposed an alternative in
a second stoppage criterion.
The second stoppage criterion is based on a number called the S-number, which
is defined as the number of consecutive siftings when the numbers of zero-crossings
and extrema are equal or at most differing by one; it requires that number shall
remain unchanged. Through exhaustive testing, Huang et al. (2003) used this S-num-
ber method of defining a stoppage criterion to establish a confidence limit for the
EMD, to be discussed later.
When the resulting function satisfies either of the criteria given above, this
component is designated as the first IMF, c1, as shown in Figure 1.12c. We can then
separate c1 from the rest of the data by
X (t ) − c1 = r1 . (1.13)
2
Amplitude
−2
−4
−6
−8
−10
200 250 300 350 400 450 500 550 600
Time: second
FIGURE 1.13 (See color insert following page 20). Comparison between data and the
residue, r1, after the first IMF, c1, is removed. Notice the residue behaves like a moving mean
to the data that bisect all the waves.
This resulting residue is shown in Figure 1.13. Since the residue, r1, still contains
information with longer periods, it is treated as the new data and subjected to the
same process as described above. This procedure can be repeated to all the subse-
quent rj’s, and the result is
r1 − c2 = r2 ,
... . (1.14)
rn−1 − cn = rn
X (t ) = ∑c
j =1
j + rn . (1.15)
Thus, we achieve a decomposition of the data into n IMF modes, and a residue,
rn, which can be either a constant, a monotonic mean trend, or a curve having only
one extremum. Recent studies by Flandrin et al. (2004) and Wu and Huang (2004)
established that the EMD is a dyadic filter, and it is equivalent to an adaptive wavelet.
Since it is adaptive, we avoid the shortcomings of using an a priori–defined wavelet
basis, and we also avoid the spurious harmonics that would have resulted. The
components of the EMD are usually physically meaningful, for the characteristic
scales are defined by the physical data. The sifting process is, in fact, a Reynolds-type
decomposition: separating variations from the mean, except that the mean is a local
instantaneous mean, so that the different modes are almost orthogonal to each other,
except for the nonlinearity in the data.
Having obtained the intrinsic mode function components, we can apply the
Hilbert transform to each IMF component and compute the instantaneous frequency
as the derivative of the phase function. After performing the Hilbert transform to
each IMF component, we can express the original data as the real part, RP, in the
following form:
n
X(t) = RP ∑ a (t) e
j=1
j
i ∫ ω j (t) dt
. (1.16)
h( ω ) =
∫ H(ω ,t)dt.
0
(1.17)
The marginal spectrum offers a measure of total amplitude (or energy) contri-
bution from each frequency value. It represents the cumulated amplitude over the
entire data span in a probabilistic sense.
The combination of the EMD and the HSA is known as the Hilbert-Huang
transform for short. Empirically, all tests indicate that HHT is a superior tool for
time–frequency analysis of nonlinear and nonstationary data. It has an adaptive basis,
and the frequency is defined through the Hilbert transform. Consequently, there is
no need for the spurious harmonics to represent nonlinear waveform deformations
as in any of the a priori basis methods, and there is no uncertainty principle limitation
on time or frequency resolution resulting from the convolution pairs possessing a
priori bases. Table 1.1 compares Fourier, wavelet, and HHT analyses.
From this table, we can see that the HHT approach is indeed a powerful method
for the analysis of data from nonlinear and nonstationary processes: it has an adaptive
basis; the frequency is derived by differentiation rather than convolution — therefore,
it is not limited by the uncertainty principle; it is applicable to nonlinear and
nonstationary data; and it presents the results in time–frequency–energy space for
feature extraction. This basic development of the HHT method has been followed
TABLE 1.1
Comparisons between Fourier, Wavelet, and Hilbert-Huang Transform
in Data Analysis
Fourier Wavelet Hilbert
by recent developments that have either added insight to the results or enhanced
their statistical significance. Some of the recent developments are summarized in
the following section.
only if the Fourier spectra for f(t) and h(t) are totally disjoint in frequency space,
and if the frequency content of the spectrum for h(t) is higher than that of f(t). This
limitation is critical, for we need to have
otherwise, we cannot use Equation 1.5 to define the phase function. According to
the Bedrosian theorem, Equation 1.19 is true only if the amplitude is varying so
slowly that the frequency spectra of the envelope and the carrier waves are disjoint.
This has made the application of the Hilbert transform even to IMFs problematic.
To satisfy this requirement, Huang and Long (2003) have proposed the normalization
of the IMFs in the following steps: starting from an IMF, we first find all the maxima
of the IMFs, defining the envelope by spline through all the maxima and designating
the envelope as E(t). Now, we normalize the IMF by dividing the IMF by E(t). Thus,
we have the normalized function with amplitude always equal to unity.
Even with this normalization, we have not resolved all the limitations on the
Hilbert transform. The new restriction is given by the Nuttall theorem (1966). This
theorem states that the Hilbert transform of cosine is not necessarily the sine with
the same phase function for a cosine with an arbitrary phase function. Nuttall gave
an error bound, ∆E, defined as the difference between y(t), the Hilbert transform of
the data, and Q(t), the quadrature (with phase shift of exactly 90°) of the function:
T 0
where Sq is Fourier spectrum of the quadrature function. The proof of this theorem
is rigorous, but the result is hardly useful, for it gives a constant error bound over
the whole data range. For a nonstationary time series, such a constant bound will
not reveal the location of the error on the time axis.
With the normalized IMF, Huang and Long (2003) have proposed a variable
error bound based on a simple argument, which goes as follows: let us compute the
difference between the squared amplitude of the normalized IMF and unity. If the
Hilbert transform is exactly the quadrature, then the squared amplitude of the
normalized IMF should be unity; therefore, the difference between it and unity
should be zero. If the squared amplitude is not exactly unity, then the Hilbert
transform cannot be exactly the quadrature. Consequently, the error can be measured
simply by the difference between the squared normalized IMF and unity, which is
a function of time. Huang and Long (2003) and Huang et al. (2005) have conducted
detailed comparisons and found the result quite satisfactory.
Even with the error indicator, we can only know that the Hilbert transform is
not exactly the quadrature; we still do not have the correct answer. This prompts
the suggestion of a drastic alternative, eschewing the Hilbert transform totally. To
this end, Huang et al. (2005) suggest that the phase function can be found by
computing the arc-cosine of the normalized function. A checking of the results so
obtained has also proved to be satisfactory. The only problem is that the imperfect
normalization will give some values greater than unity. Under that condition, the
arc-cosine will break down.
An example of the normalized and regular Hilbert transforms is given in Figure
1.14, from the data given in Figure 1.12c. There are three different instantaneous
frequency values: the instantaneous frequency from regular Hilbert transform, the
normalized Hilbert transform, and the generalized zero-crossing, which can serve
as the standard in the mean. It is easy to see that the normalized instantaneous
Data
IF-H
IF-NH
IF-Z
0.15
0.1
Amplitude
0.05
–0.05
0 50 100 150 200 250 300 350 400
Time : second
FIGURE 1.14 (See color insert following page 20). Comparison of the instantaneous fre-
quency values derived from different methods. Note that the Instantaneous from IMF is still
not correct when the amplitude fluctuates too much. The normalized Hilbert Transform,
however, gives a much better instantaneous frequency when compared with the values derived
from the generalized-zero-crossing method.
frequency is very close to the zero-crossing values, while the regular Hilbert trans-
form result gives large undulations that will never result in the mean as given by
the zero-crossing method. The high undulation results from the large changes of
amplitude and some nonlinear distortions of the waveform, both of which will cause
the envelope to fluctuate as shown in Figure 1.15. In the normalization scheme, the
smooth spline helps to eliminate many of the undulations in the resulting instanta-
neous frequency. One can also see that the problem of the regular Hilbert transform
occurs always at the location where either the amplitudes change drastically or the
amplitude is very low, as predicted by the Nuttall theorem. The normalized Hilbert
transform alleviates the problems substantially.
Finally, the error index is given in Figure 1.16; here we can see that the error
is also small in general, unless the waveform is locally distorted. Even over the large
error location, the index values are smaller than 10%, except for the end region,
where the end effect of the Hilbert transform causes additional problems. Thus the
normalized Hilbert transform has helped to overcome many of the difficulties of the
regular Hilbert transform, and it should be used all the time.
6
Amplitude : cm
0
0 50 100 150 200 250 300 350 400
Time : second
FIGURE 1.15 (See color insert following page 20). The instantaneous amplitude (or the
envelope) of the test data. Note the improvement in adopting the spline envelope over the
simple analytic function.
0.08
0.06
0.04
0.02
Amplitude
–0.02
–0.04
–0.06
–0.08
–0.1
0 50 100 150 200 250 300 350 400
Time : second
FIGURE 1.16 (See color insert following page 20). The Error Index of the normalized
Hilbert transform; it has large value whenever the wave form deviated from a smooth sinu-
soidal form. But their values are, in general, small except near the ends.
Additionally, Huang et al. (2003) invoked the intermittence criterion and forced
the number of IMFs to be the same for different S-numbers. As a result, they were
able to find the mean for specific IMFs. Figure 1.17 shows the IMF representing
variations of the annual cycle of the length of day. The peak and valley of the
envelope represent the El Niño events. Of particular interest are the periods of high
standard deviations, from 1965 to 1970 and from 1990 to 1995. These periods turn
out to be the anomaly periods of the El Niño phenomena, when the sea surface
temperature readings in the equatorial region were consistently high based on obser-
vations, indicating a prolonged heating of the ocean, rather than the changes from
warm to cool during the El Niño to La Niña changes.
Finally, from the confidence limit study, an unexpected result was the determi-
nation of the optimal S-number. Huang et al. (2003) computed the difference between
the individual cases and the overall mean and found that there is always a range
where the differences reach a local minimum. Based on their limited experience
from different data sets, they concluded that an S-number in the range of 4 to 8
performed well. Logic also dictates that the S-number should not be too high (which
would drain all the physical meaning out of the IMF) nor too low (which would
leave some riding waves remaining in the resulting IMFs).
0.04
0.02
Amplitude
–0.02
–0.04
–0.06
–0.08
–0.1
1965 1970 1975 1980 1985 1990 1995 2000
Time : second
FIGURE 1.17 (See color insert following page 20). The mean envelope of the annual
cycle IMF component from LOD data. The peaks of the envelope are all aligned with El Nio
events, when the additional angular momentum imparted to the atmosphere from the over
heated Equatorial ocean water. The large scatter of the envelope periods in 1065-70 and 1990-
95 represent periods of El Nio anomalies.
Instead of fractal Gaussian noise, Wu and Huang (2004) studied the Gaussian
white noise only. They also found the relationship between the mean period and
RMS values of the IMFs. Additionally, they have also studied the statistical prop-
erties of the scattering of the data and found the bounds of the data distribution
analytically. From the scattering, they deduced a 95% bound for the white noise.
Therefore, they concluded that when a data set is analyzed with EMD, if the mean
period-RMS values exist within the noise bounds, the components most likely
represent noise. On the other hand, if the mean period-RMS values exceed the noise
bounds, then those IMFs must represent statistically significant information.
1.4 CONCLUSION
HHT is a relatively new method in data analysis. Its power is in the totally adaptive
approach that it takes, which results in the adaptive basis, the IMFs, from which the
instantaneous frequency can be defined. This offers a totally new and valuable view
of nonstationary and nonlinear data analysis methods. With the recent developments
on the normalized Hilbert transform, the confidence limit, and the statistical signif-
icance test for the IMFs, the HHT has become a more robust tool for data analysis,
and it is now ready for a wide variety of applications. The development of HHT,
however, is not over yet. We still need a more rigorous mathematical foundation for
the general adaptive methods for data analysis, and the end effects must be improved
as well.
REFERENCES
Bedrosian, E. (1963). On the quadrature approximation to the Hilbert transform of modulated
signals. Proc. IEEE, 51, 868–869.
Cohen, L. (1995). Time-Frequency Analysis. Prentice Hall, Englewood Cliffs, NJ.
Diks, C. (1997). Nonlinear Time Series Analysis. World Scientific Press, Singapore.
Daubechies, I. (1992). Ten Lectures on Wavelets. SIAM, Philadelphia.
Flandrin, P. (1999). Time-Frequency/Time-Scale Analysis. Academic Press, San Diego, CA.
Flandrin, P., Rilling, G., and Gonçalves, P. (2004). Empirical mode decomposition as a
filterbank. IEEE Signal Proc. Lett. 11 (2): 112–114.
Hahn, S. L. (1996). Hilbert Transforms in Signal Processing. Artech House, Boston.
Huang, N. E., and Long, S. R. (2003). A generalized zero-crossing for local frequency
determination. U.S. Patent pending.
Huang N. E., Long, S. R., and Shen, Z. (1996). Frequency downshift in nonlinear water wave
evolution. Advances in Appl. Mech. 32, 59–117.
Huang, N. E., Shen, Z., Long, S. R. (1999). A new view of nonlinear water waves — the
Hilbert spectrum. Ann. Rev. Fluid Mech. 31, 417–457.
Huang, N. E., Shen, Z., Long, S. R., Wu, M. C., Shih, S. H., Zheng, Q., Tung, C. C., and
Liu, H. H. (1998). The empirical mode decomposition method and the Hilbert spec-
trum for non-stationary time series analysis. Proc. Roy. Soc. London, A454, 903–995.
Huang, N. E., Wu, Z., Long, S. R., Arnold, K. C., Blank, K., Liu, T. W. (2005). On instan-
taneous frequency. Proc. Roy. Soc. London (submitted).
Huang, N. E., Wu, M. L., Long, S. R., Shen, S. S. P., Qu, W. D., Gloersen, P., and Fan, K.
L. (2003). A confidence limit for the empirical mode decomposition and the Hilbert
spectral analysis. Proc. Roy. Soc. London, A459, 2317–2345.
Kantz, H., and Schreiber, T. (1997). Nonlinear Time Series Analysis. Cambridge University
Press, Cambridge.
Nuttall, A. H. (1966). On the quadrature approximation to the Hilbert transform of modulated
signals. Proc. IEEE, 54, 1458–1459.
Priestley, M. B. (1988). Nonlinear and nonstationary time series analysis. Academic Press,
London.
Tong, H. (1990). Nonlinear Time Series Analysis. Oxford University Press, Oxford.
Wu, Z., and Huang, N. E. (2004). A study of the characteristics of white noise using the
empirical mode decomposition method. Proc. Roy. Soc. London, A460, 1597–1611.
CONTENTS
25
ABSTRACT
2.1 INTRODUCTION
A significant number of reported damages to ships and offshore structures suggest
the existence of rogue waves (or freak waves), which are characterized by single
exceptional large wave heights [1]. Rogue waves are defined as transient waves,
existing only in one specific location in one particular instant in time. The assumption
that these waves occur is preliminary, based on superposition of an infinite number
of linear (free) wave components that coincide in phase during storm events (see
e.g.: [2], [3], or [4]). Other attempts to explain them focus on wave–current inter-
action processes as a possible mechanism to generate or to enlarge the possibility
of those abnormal events [5].
Extensive work has been carried out to study these phenomena, either to prove
their existence by attempting to record rogue waves or to examine primary driving
mechanics and perform numerical simulations (see e.g.: [3], [4], [25], [7], and [8]).
Other researchers reproduce transient waves under laboratory conditions and derive
useful kinematic insights in order to establish advanced design criteria for offshore
structures and deep-sea vessels (see e.g.: [1], [9], [10], or [11]). However, simple
superposition of free wave components appears not to be an adequate tool to simulate
rogue waves, since linear wave theories fail to describe crest and trough heights; they
possess no Gaussian or even near-Gaussian statistics [12]. Taking the mathematical
approach of the perturbation expansion theory to higher order degrees and, yet, allo-
cating wave–wave interaction between nonlinear components often makes little physical
sense, although numerical models [10] and analytical approaches [13] match data
measured from laboratory experiments exceptionally well. Whether those approaches
are appropriate to fit recorded extreme wave events from real sea states is still uncertain.
Recent research results propose different, more complex underlying hydrody-
namic mechanisms [14, 15] that are based on phenomena defined as wave breakdown
and focusing and self-organization. However, it is not in the scope of the present
paper to deepen those nonlinear theories; rather, it is our purpose to identify the
embedded structure of rogue waves to determine physical insights about their occur-
rence in accordance with those newly developed theoretical explanations.
Therefore, the present paper relates to the newly developed Hilbert-Huang trans-
formation (HHT) pioneered by Huang et al. [16, 17]. The numerical procedure of
the key element, the so-called empirical mode decomposition (EMD), is made clear
in principle, although we direct the interested reader to study the elementary publi-
cations by Huang et al. In this context, we first examine the generation of validated
Fourier-based spectra and phase representations, since these are understood to be
most familiar to the reader associated or working in ocean engineering or other
related fields. The performance and limitations of those time-invariant techniques
provide an introduction to the second section. Here, we apply the EMD and empha-
size an appropriate calculation of the embedded modes from the original signals.
Next, we quantify a certain essential parameter within the intrinsic mode function
(IMF) and attempt to discover internal relationships between associated energy
densities and averaged periods. We then compare our data with proposed classifica-
tions on this decomposition technique performed for synthetically generated noisy
data by Flandrin et al. [18] and Wu and Huang [19]. We demonstrate that our data
accord, tentatively, with the above-mentioned recently established, more complex
generation mechanisms of rogue waves. Finally, we summarize the main outcomes
of the present investigation and offer conclusions for future research.
20
15
10
η(t) (m)
−5
−10
0 200 400 600 800 1000 1200
time (s)
FIGURE 2.1 Signal 1: New Year’s Eve rogue wave 01/01/1995, Draupner platform, STA-
TOIL, Norway.
20
15
10
η(t) (m)
−5
−10
0 200 400 600 800 1000 1200
time (s)
FIGURE 2.2 Signal 2: Sea state recorded immediately after the New Year’s Eve rogue wave
01/01/1995, Draupner platform, STATOIL, Norway.
peak components and a relatively gentle negative energy slope adjacent to the peak
waves on the “tail” of the spectrum that approximately follows a f–5 law [23].
According to the fundamental perturbation expansion approach, which is understood
to simulate nonlinear water waves accurately, the latter part of the spectrum is usually
recognized to contain the high order nonlinear harmonics of the full wave train. This
particular spectrum of the New Year’s Eve rogue wave exhibits peak components
with frequencies between 6 × 10–2 Hz and 9 × 10–2 Hz, but highly misleading spectral
fluctuations make an overall interpretation of a biased spectrum impossible. It turned
out to be most effective to proceed with some spectral smoothing techniques.
Figure 2.4 displays the estimated power spectral density using Welch’s method.
Here, the raw input signals with N = 2,560 data points is subdivided into four, five,
and ten sections of equal length (nwin = 640, 512, and 256, respectively), to give
an idea of the most appropriate spectral smoothing method. With an overlap between
the individual windows of 32 data points (5%, 6.25%, and 12.5% of window length),
each subdivision is carried out by means of Welch’s averaged modified periodogram
method of spectral estimation. The Nyquist frequency remains constant. In respect
to the raw data series, spectral frequency resolution becomes coarser with more data
segmentation; df = 1.6667 × 10–3 Hz, 2.0833 × 10–3 Hz, and 4.1667 × 10–3 Hz for
Welch spectra with equal-length sections nwin = 640, 512, and 256, respectively.
103
102
101
PSD (m2/Hz)
100
10−1
10−2
10−3
10−2 10−1 100
Frequency (Hz)
103
pwelch(640, 32)
pwelch(512, 32)
pwelch(256, 32)
102
101
PSD (m2/Hz)
100
10−1
10−2
10−3
10−2 10−1 100
Frequency (Hz)
FIGURE 2.4 Estimated power spectral density of signal 1 made using Welch’s method.
only multidirectional waves can generate double-peaked spectra. But, in the case of
the New Year’s Eve rogue wave, a clear unidirectional wave field emerged from the
severe storm, as Trulsen [25] significantly points out, since the measured response
of the platform was found to be solely unidirectional. It is therefore concluded that
the waves were essentially long-crested. So, our result shown in Figure 2.4, a
particular double-peakedness with very closely resolved carrier wave components,
can be validated. We will show that both dominating carrier wave groups fpeak1,1st and
fpeak1,2nd are resolved as well by the EMD.
Moreover, higher order nonlinear superharmonics — either bounded or disper-
sive (free) wave components of higher frequency — are evident at fpeak1,2nd = 1.01 ×
10–1 Hz and fpeak2,2nd = 1.03 × 10–1 Hz, as shown in Figure 2.4. These specific riding
wave groups show significant smaller energy fraction than the dominant wave modes
at peak frequencies in the spectrum and are clearly evident from all three Welch
spectra. Energy portions of infra-gravity components and other higher order super-
harmonics are of minor importance.
Power spectral densities for the second data series of the sea state are shown in
Figure 2.5 and Figure 2.6. Obviously, we see no significant differences to the raw
spectrum in Figure 2.3, as high-frequency fluctuations disrupt the overall spectral
energy content and make an appropriate interpretation difficult. Smoothed estimated
103
102
101
PSD (m2/Hz)
100
10−1
10−2
−3
10
10−2 10−1 100
Frequency (Hz)
spectra obtained with Welch’s methods also identify the double-peaked dominant
components, but this time they are less meaningful. Only the first subdivided Welch
frequency representation [pwelch(640,32)] helps to reveal the preceding carrier
components, at fpeak1,1st = 5.7 × 10–2 Hz and (now slightly lower) at fpeak2,1st = 7.5 ×
10–2 Hz. Nevertheless, the rest of the spectrum from this data series is mostly identical
to the one examined before — even the riding wave components are evident at
fpeak1,2nd = 1.01 × 10–1 Hz and fpeak2,2nd = 1.03 × 10–1 Hz.
From these simple spectral estimates obtained with Welch’s method, the pre-
vailing components, which are supposed to be accountable for the occurrence of
this particular abnormal rogue wave, are evaluated, leading to a physical model of
two (or three) main carrier wave components and two riding wave components.
Their phase orientation and time-dependent dispersive propagation are supposed to
be in charge for any transient phenomena occurring at sea. At this stage, we can
conclude that it is not possible to understand the occurrence of rogue waves through
an examination of power spectral densities alone. Transient data are completely
misinterpreted, since essential time-dependent information is cancelled out through
ordinary Fourier-based techniques. The occurrence of rogue waves on the open sea
is therefore simply disregarded in Fourier-based power spectra density representa-
tions. Nonetheless, to support this mathematical approach on the whole, we now
look more closely at the corresponding phase spectra.
103
pwelch(640, 32)
pwelch(512, 32)
pwelch(256, 32)
102
101
PSD (m2/Hz)
100
10−1
10−2
10−3
10−2 10−1 10
0
Frequency (Hz)
FIGURE 2.6 Estimated power spectral density of signal 2 made using Welch’s method.
100
−100
Unwrapped phase (radians)
−200
−300
−400
−500
−600
−700
−800
10−2 10−1 100
Frequency (Hz)
phase lag between the components tends to be constant and, therefore, is mainly
independent of the window length of the Welch spectrum. Otherwise, however, this
analysis technique is categorically unqualified to examine supposed elementary
generating mechanisms, since it is mainly based on averaging the unwrapped phases
from the individual windows. Therefore, implemented variations are pronounced
excessively and yield to misleading results.
To validate this conclusion, we similarly analyzed the second signal. Calculated
results for the raw unwrapped phase spectra of the second data series, shown in
Figure 2.9, are almost identical to those of the first data series. Note that these
unfiltered phase spectra are rather unsatisfactory for physical interpretation. Again,
Welch’s estimated spectral phase representations are applied, and results are shown
in Figure 2.10. Previously proposed effects are shown to be valid for this data series
as well, since all three Welch representations produce significant dissimilar phase
spectra. This indirectly points to the fact that both underlying peaks and both higher
order superharmonic wave groups are partially out of phase. Table 2.2 summarizes
the derived results for the unwrapped phases from signal 2.
Consequently, we conclude that no anomalous effects concerning driving mech-
anisms of tentative rogue waves are apparent in the data series. Fourier-based
analysis seems to be effective as long as the mathematical model attempts to for-
mulate the wave groups as stationary and composed solely of linear components
3π
pwelch(640, 32)
pwelch(512, 32)
pwelch(256, 32)
2π
Unwrapped phase (radians)
−π
−2π
10−2 10−1 100
Frequency (Hz)
FIGURE 2.8 Estimated unwrapped phase (radians) spectra of signal 1 made using Welch’s
method.
TABLE 2.1
Unwrapped Phases of Significant
Components from Signal 1
fpeak1 ,1st fpeak2 ,1st fpeak1 ,2nd fpeak2 ,2nd
100
−100
Unwrapped phase (radians)
−200
−300
−400
−500
−600
−700
−800
10−2 10−1 100
Frequency (Hz)
3π
pwelch(640, 32)
pwelch(512, 32)
pwelch(256, 32)
2π
Unwrapped phase (radians)
−π
−2π
10−2 10−1 100
Frequency (Hz)
FIGURE 2.10 Estimated unwrapped phase (radians) spectra of signal 2 made using Welch’s
method.
TABLE 2.2
Unwrapped Phases of Significant
Components from Signal 2
fpeak1 ,1st fpeak2 ,1st fpeak1 ,2nd fpeak2 ,2nd
Consequently, these techniques are also applied to ordinary wave data and are
approved in the coastal and ocean engineering disciplines. Liu [27, 28, 29] introduces
the wavelet transform technique in its entirety and analyzes coastal wave data sets
from the Great Lakes. Chien et al. [30] analyze nearshore observed extreme wave
events off the east coast of Taiwan. Schlurmann [31, 32] previously studied time–fre-
quency representations of rogue waves in the Sea of Japan and focused on a direct
comparison between wavelet and Hilbert spectra.
However, within the present study, another time-dependent spectral analysis tool,
the empirical mode decomposition (EMD), is applied to both data series recorded
in the central North Sea. As mentioned, this technique has been developed by Huang
et al. [16, 17, 33] and recently extended by Flandrin and his group [18, 34]. The
EMD is capable of disintegrating transient, nonlinear data into its own embedded
characteristic oscillations on a non–a priori basis. This feature constitutes the essen-
tial difference from other time–frequency analysis techniques, such as wavelet trans-
formations. Huang et al. provide a general overview of EMD and its recent appli-
cations [16, 17, 33], and full details are given by Dätig and Schlurmann [36].
η(t) (m) 20
0
−20
10
c1
0
−10
10
c2
0
−10
5
c3
0
−5
2
c4
0
−2
1
c5
0
−1
0.5
c6
0
−0.5
0.5
c7
0
−0.5
0.5
c8
0
−0.5
0.1
c9
0
−0.1
0.5
rn
0
−0.5
0 200 400 600 800 1000 1200
time (s)
FIGURE 2.11 The IMF components from extrema-based sifting without intermittency testing
of signal 1.
the first IMF at around t = 280 sec, with an approximate wave height of 12 m.
Compared to this unusually large wave height in c1, the rest of this particular mode
is characterized by relatively low amplitudes. It is well accepted from preceding
investigations that it is the individual time scale of an embedded mode that makes
the EMD work successfully in disintegrating meaningful components from any given
η(t) (m) 20
0
−20
5
c1
0
−5
20
c2
0
−20
5
c3
0
−5
5
c4
0
−5
1
c5
0
−1
1
c6
0
−1
0.5
c7
0
−0.5
0.05
c8
0
−0.05
0.4
rn
0.2
0
0 200 400 600 800 1000 1200
time (s)
FIGURE 2.12 The IMF components from extrema-based sifting without intermittency testing
of signal 2.
signal, and not its time-dependent amplitude. We can conclude that the existence of
the New Year’s Eve rogue wave is a sudden and unexpected phenomenon within the
data series, being part of the high frequency component of the sea state. Spectral
analysis of sifted IMF will resolve further particulars. Regardless of this observation,
the higher order modes c2 to c4 emerge as almost of constant period, but allowing
time-dependent amplitude variations. This feature builds the main advantage of the
103
x1(t)
IMF1(t)
IMF2(t)
102
IMF3(t)
IMF4(t)
IMF5(t)
101
IMF6(t)
IMF7(t)
PSD (m2/Hz)
IMF8(t)
100
10−1
10−2
10−3
10−2 10−1 100
Frequency (Hz)
FIGURE 2.13 Estimated power spectral density made using Welch’s method [pwelch
(512,32)] from raw data (signal 1) and IMF components of signal 1.
As assumed previously, it is now proven that c1 in Figure 2.13 contains all high
frequency contents of the signal. From a cut-off frequency of about 1.1 Hz its
distribution is identical to that of the original data series. This first IMF is extremely
broad banded with no significant peak component. It ranges from 10–2 Hz up to the
Nyquist frequency fNyquist = 1.0667 Hz with only slightly varying energy distribution,
but it fails to embody any of the carrier or riding wave components mentioned earlier.
The second IMF implements a substantially smaller frequency bandwidth, clearly
constituting the higher order superharmonic components at around fpeak1,2nd = 1.01 ×
10–1 Hz and fpeak2,2nd = 1.03 × 10–1 Hz and almost the full energy content of one of
the dominant carrier waves at fpeak2,1st = 8.3 × 10–2 Hz. The subsequent mode, c3,
contains the other dominating carrier wave group at fpeak1,1st = 5.7 × 10–2 Hz. None-
theless, both components c2 and c3 show a certain degree of frequency overlapping,
but they do individually characterize a predominant wave group inside the spectrum.
At this stage, it is remarkable that the EMD automatically sifts those components
that were previously identified from ordinary Fourier-based transformations. Higher
order modes disintegrated from the EMD are situated in lower frequency regions,
but these play only marginal role in the overall energy spectra.
Although those disintegrated modes principally match observation from Fourier
analysis, suggesting that the EMD is a successful method for analyzing transient
water waves, it is cumbersome that this particular sifting procedure artificially creates
energy modes in the infra-gravity wave range of the spectra. These are not evident
in spectral representation of the original data set. Modes c1 to c8 do all show higher
energy magnitudes in the lower frequency region than the water surface elevation
from the measured sea state initially shows. Dätig and Schlurmann [36] have reported
this effect in bichromatic wave groups and concluded that it is an artifact of the
numerical sifting algorithm of the EMD. We do not clarify this mismatch within the
present paper and suggest further investigations in future research projects.
Figure 2.14 illustrates the spectral representation of characteristic oscillations
of the second data set. Again, Welch’s estimated power spectral density method is
applied, with identical specifications to the spectra shown in Figure 2.13. The bold
solid line indicates the spectral distribution of the original data series. Again, c1 is
broad banded and mainly shows wave groups with relatively high frequencies. It is
identical to the initial signal from an approximate cut-off frequency of 1.1 10–2 Hz.
In contrast to the first signal, the second IMF accords with the assumptions on carrier
and riding waves stated earlier. Note that only mode c2 implements peak carrier
wave groups as well as the assumed riding wave components. The EMD fails in
distinguishing those characteristic oscillations and finally yields just nine modes
compared to ten IMFs from the first signal. The third IMF, c3, shows a significant
overlap with c2, but it cannot exclusively occupy the carrier wave group around
frequencies of fpeak1,1st = 5.7 × 10–2 Hz. Higher order modes disintegrated from the
EMD are also located in the lower frequency regions. Their influence on the overall
appearance of the spectral decomposition is marginal and can be neglected. As
already observed, cumbersome artificially created energy modes in the infra-gravity
wave range of the spectra are also verified for the second signal and are due to
numerical artifacts of the sifting procedure within the EMD.
103
x1(t)
IMF1(t)
IMF2(t)
102
IMF3(t)
IMF4(t)
IMF5(t)
101
IMF6(t)
IMF7(t)
PSD (m2/Hz)
100
10−1
10−2
10−3
10−2 10−1 100
Frequency (Hz)
FIGURE 2.14 Estimated power spectral density made using Welch’s method [pwelch
(512,32)] from raw data (signal 2) and IMF components of signal 2.
TABLE 2.3
Data for Signal 1. (a) Number of Extrema, (b) Number of Zero-Crossings,
(c) Mean Periods of IMFs in sec, (d) Corresponding Standard Deviation (STD)
in sec
mode k 1 2 3 4 5 6 7 8 9 10
TABLE 2.4
Data for Signal 2. (a) Number of Extrema, (b) Number of Zero-Crossings,
(c) Mean Periods of IMFs in sec, (d) Corresponding Standard Deviation (STD)
in sec
mode k 1 2 3 4 5 6 7 8 9
et al. [18] make clear that the average period of these zero-crossings is a rough
indication of the mean embedded frequency of each IMF. The way this number
varies from mode to mode is further indication of the hierarchical structure of the
decomposed natural data series from the central North Sea. The standard deviation
also suggests the bandwidth of the disintegrated mode. Table 2.3 summarizes results
for the first signal, containing the abnormal wave.
The same measurements and specifications were made for the second data set;
these are listed in Table 2.4. Clearly, the natural data sets do not behave like white
noise or any other artificially created data. Flandrin et al. [18] and Wu and Huang
[19] suggest an interesting pattern in stochastic data: that the average period of any
IMF component almost exactly doubles that of the previous one. They conclude that
this behavior establishes the EMD as performing as a dyadic filter bank. This
mathematical feature is not drawn from the present data sets. However, as Table 2.2
and Table 2.4 show, the first five modes of each signal correlate quite well. At least,
the number of extrema, mean periods, and corresponding standard deviations are
strongly related. This pattern is especially interesting given that we are analyzing
two independent data series.
10
6
Log2 (period)
0
0 1 2 3 4 5 6 7 8 9 10 11
Mode k
FIGURE 2.15 Circle: Log2 mean periods versus mode number k of signal 1. Solid line: Log2
linear fit. Dashed line: confidence limits.
Tk = 2( ak + b ) (2.1)
10
6
Log2 (period)
0
0 1 2 3 4 5 6 7 8 9 10 11
Mode k
FIGURE 2.16 Circle: Log2 mean periods versus mode number k of signal 2. Solid line: Log2
linear fit. Dashed line: confidence limits.
where k is the mode number and a and b are real valued coefficients. A least squares
method determines a = 0.7877 and b = 1.7039 to fit the average period log2(Tk).
Modes c1 and the residue c10 heavily deviate from the linear fitted function. As shown
earlier, spectral distribution of the first IMF is very broad-banded and represents a
kind of half-band high pass filter. It contains more or less exclusively the high
frequency fractions of the original signal, so that the mean period of c1 is far less
than that of the linear fitted function from c2 to c9. The dotted black line, which is
constructed from simple extrapolation of the latter linear fitted function, makes this
clear.
The mathematical behavior of the other signal is shown in Figure 2.16. Here,
only modes c2 to c5 appear to obey this strict (log2) linearity of the mean periods.
Modes c1 and IMFk for 6 k 9 show strong divergence from this log2 linearity.
Coefficients for the second data series are determined to be a = 0.8709 and b =
1.4917 to fit the averaged periods log2(Tk) with 2 k 5. This behavior tends to
underestimate higher modes (6 k 9), whereas c1 is categorically overestimated, as
the dotted black line reveals — again determined through extrapolation of the
perfectly fitting function from modes c2 to c5. Flandrin et al. extend their method to
determine whether those dyadic filter modes can be classified for a possible self-sim-
ilarity behaviour of their data. This feature is (so far) unnecessary in application to
the present data, since transient water waves from the central North Sea evidently
disobey the dyadic filter bank characteristic.
Tk ⋅ Ek = p (= const ) (2.2)
or
Ek = 10 cTk + d (2.4)
where c = –2.0264 and d = 2.5001 are derived for the modes from the first data
series and c = –2.0118 and d = 2.6944 are calculated for the IMF of the second data
100
90
80
70
60
Tk* Ek
50
40
30
20
10
0
0 1 2 3 4 5 6 7 8 9 10 11
Mode k
FIGURE 2.17 Circle: product of energy density and mean period periods versus mode
number k of signal 1. Square: product of energy density and mean period versus mode number
k of signal 2.
series. Only modes c1 and the residuals c10 show significant divergence from being
fitted. Clearly, Equation 2.4 offers an excellent correlation to these fitted points.
Thus, we can prove for natural data series from the central North Sea that the energy
density of the IMF and its averaged period follows a hyperbolic function. This
outcome is in agreement with results obtained by Wu and Huang [19].
0.5
−0.5
−1
Log10 (Ek)
−1.5
−2
−2.5
−3
−3.5
−4
0 0.5 1 1.5 2 2.5 3 3.5 4
Log10 (Tk)
FIGURE 2.18 Circle: relation of energy density and mean period periods of signal 1. Square:
relation of energy density and mean period of signal 2. Solid line: Log2 linear fit, signal 1.
Dashed line: Log2 linear fit, signal 2.
wave events from real sea states. As Onorato et al. [14], Pelinovsky et al. [15], and
Trulsen and Dysthe [6] point out, the utility of those models has not yet been
determined. They propose different, more complex underlying hydrodynamic mech-
anisms for the generation of rogue waves, based on phenomena named wave break-
down and focusing and self-organization. These phenomena are described by an
envelope function on the basis of the nonlinear Schrödinger equation and higher
order corrections (Dysthe equations), yielding an equation hierarchy that attempts
to explain an abnormal wave within real sea states through a simple model.
It is found numerically that rogue wave events are mainly related to the initial
steepness and to the correlation length of the complex envelope of the initial time
series. The effects of wave breakdown and focusing and self-organization arise
through the concept of Benjamin-Feir [37] instability mechanisms, which cause local
exponential growth in the amplitude of a wave train. Still unknown is how rogue
waves can generate these instabilities since realistic oceanic conditions are not
characterized by a single monochromatic wave perturbed by two small-amplitude
parasitic waves in the adjacent side-bands, but instead by a complex spectrum whose
perturbation of the carrier wave cannot be viewed as being small. Therefore, Onorato
et al. [14] perform numerical simulations based on the envelope equations by using
an inverse scattering technique, which finally calculates the maximum wave ampli-
tude as a function of the wave steepness.
Interestingly, Pelinovski et al. [15] indicate that for shallow water, the Korteweg-de
Fries equation is an adequate tool to demonstrate wave focusing. They show that
large-amplitude abnormal short-lived impulses can be generated from weak,
so-called invisible deterministic components in the background of the random wind
wave field. An initial disturbance representing a single crest, for example, evolves
into a set of solitons (solitary waves) and an oscillating dispersive tail located in
space according to the magnitudes of the individual celerity of each embedded
component. They also make use of the nonlinear Schrödinger equation for describing
the complex amplitude of the wave envelope in a deep water environment, and they
make clear in this context that the well-known Benjamin-Feir effect of modulation
instability can explain the appearance of abnormal waves in a periodic weakly
modulated wave field with heights exceeding two to three times the unperturbed
wave height.
It is not in the scope of the present paper to deepen those nonlinear theories,
but it is of main interest to identify the embedded structure of rogue waves to discover
new physical insights about their occurrence that accord with these newly developed
theoretical explanations. Nonetheless, numerical investigations suggest that the phys-
ical model of an infinite number of free waves that coincide is not sufficient to
describe the occurrence of water waves of extreme wave height. This is in close
correspondence with the work carried out by Trulsen and Dysthe [6], who simulate
the measured time series of the New Year’s Eve rogue wave forward and backward
in space to find the time histories at neighboring locations. Based on the modified
nonlinear Schrödinger equations with exact linear dispersion, he concludes that the
abnormal wave did not appear suddenly from nowhere; instead, he suggests that a
short and tall wave group approached the platform about one minute before the
rogue wave hit the structure. Both groups appear to split up and run away from a
longer and smaller trailing group that becomes rather diffuse as it approaches the
platform. The rogue wave that hits the structure arrives in the middle of the short
leading group. After the impact, the whole wave group broadens, becomes less steep,
and slows down slightly. A large trough, a so-called hole in the sea, is observed
slightly upstream of the platform.
All numerical observations strongly correlate with a long-established model for
nonlinear wind waves created by Lake and Yuen [38]. They propose to characterize
a wind–wave system by an ensemble of single nonlinear wave trains, which are
considered to be a coherent bound-wave system. The associated energy is transported
only at the group velocity of the dominant wave and is governed by nonlinear
self-interactions of the type found in amplitude-modulated wave trains. Essentially,
all of the energy in the resulting nonlinear wind–wave system is contained in the
bound-wave components of a single dominant wave. The individual components do
not propagate as free waves and do not obey the usual dispersion relation, but they
are in spectral representations merely Fourier components needed to describe the
complicated shapes of the individual waves as well as of their modulating envelope.
Lake and Yuen propose that to a leading order the individual components still
travel at the single wave speed C0 = ω0 /k0, where ω0 is the carrier frequency and k0
is its corresponding wave number. The phase speed of the carrier wave component
and the energy (or envelope) is equal to the carrier wave group celerity Cg0 = ω0 /2k0.
Lake and Yuen further add that although the spectrum of these modulated wave
groups appears more or less complicated, covers a relatively broad frequency range,
and contains many components that are not harmonics of the carrier wave, it is in
fact questionable whether the existence of a coherent carrier wave form is proven
solely from visual inspection of the spectrum. Nevertheless, the wave system that
produced the spectrum in their analysis has a coherent carrier wave, bound-wave
components, and a single energy propagation speed, although — and this is important
— it does not propagate without change of form.
It is believed that this model is essentially valid for the present case in the central
North Sea and that the EMD is categorically capable of deriving these self-modu-
lating carrier wave components according to the Lake and Yuen model. Although
we have established that this particular sea state is tentatively composed of at least
two carrier waves and two additional riding wave components, we have not deter-
mined whether the latter ones are bounded or dispersive within the spectral repre-
sentation. To attempt to prove our assumption, we simply add both intrinsic modes
c2 and c3 and clarify how closely this reproduces the full water surface elevation.
Figure 2.19 and Figure 2.20 show this comparison for both time series. The differences
20
x1(t) (m)
10
−10
0 200 400 600 800 1000 1200
20
c2(t) + c3(t) (m)
10
−10
0 200 400 600 800 1000 1200
20
∆x(t) (m)
10
−10
0 200 400 600 800 1000 1200
Time (s)
FIGURE 2.19 Top: Signal 1. Middle: Sum of c2 and c3 of signal 1. Bottom: Difference
between signal 1 and sum of c2 and c3 of signal 1.
x2(t) (m) 20
10
−10
0 200 400 600 800 1000 1200
20
c2(t) + c3(t) (m)
10
−10
0 200 400 600 800 1000 1200
20
∆x(t) (m)
10
−10
0 200 400 600 800 1000 1200
Time (s)
FIGURE 2.20 Top: Signal 2. Middle: Sum of c2 and c3 of signal 2. Bottom: Difference
between signal 2 and sum of c2 and c3 of signal 2.
x(t) between the original signals x1(t) and x2(t), respectively, and the sum of modes
c2 and c3 are shown individually in the bottom subfigures of Figure 2.19 and Figure
2.20. Obviously, only marginal divergence is noticed for both time series, mainly
indicating that some higher frequency waves are missing. Most obvious is the omitted
peak of the New Year’s Eve rogue wave, which is hidden in the high frequency
component c1 disintegrated from the first signal.
Fourier-based transformation, discussed earlier, suggests two peak carrier wave
components located in the spectrum at fpeak1,1st = 5.7 × 10–2 Hz and fpeak2,1st = 8.3 ×
10–2 Hz. These frequencies recalculated into periods are defined as Tpeak1,1st = 17.5
sec and Tpeak1,2nd = 12.0 sec. According to Table 2.3 and Table 2.4, the EMD proposes
the mean period of components c2 and c3 to be about T2 = 9.4 sec and T3 = 16.9 sec
to 17.7 sec, respectively. Mode c2 also embeds both previously identified superhar-
monic riding waves at approximately fpeak1,2nd = 1.01 × 10–1 Hz and fpeak2,2nd = 1.03
× 10–1 Hz, so that the averaged period is significantly smaller for that mode. Nev-
ertheless, it is most interesting that both carrier waves and fractions of the riding
wave components previously detected from Fourier-based spectra are validated in
principle from the EMD. Moreover, if we now apply the Hilbert transformation to
those derived components from the EMD, we can easily confirm Lake and Yuen’s
proposed model, because time-frequency representation using the Hilbert spectral
technique allows modulating amplitudes within time and frequency domains.
0.20
c1
0.16
Instantaneous frequency (Hz)
0.12
c2
0.08
c3
0.04
c4
0
0 200 400 600 800 1000 1200
Time (s)
Figure 2.21 and Figure 2.22 present results from the Hilbert spectral analysis.
Each figure renders the time–frequency dependent representations of modes c1 to
c4. Evidently, amplitudes of the carrier wave components vary in time; a detailed
measure in the form of colormaps is omitted here. Alternating colors exclusively
indicates this tendency. Large amplitudes are designated by black colors, and smaller
time-variant wave heights are shown by gray-shaded colors. Note that each Hilbert
spectrum is created by use of an additional filter algorithm to smooth these distinct
representations to a coarser time–frequency grid (400 × 400).
Intrinsic modes c2 to c4 shown in Figure 2.21 are characterized as slowly varying
time-dependent functions. While the instantaneous frequency of c4 is nearly constant,
located at f = 0.04 Hz, both dominating carrier wave components c3 and c2 show
slightly larger time-variant spectral deviations. The latter mode alternates around an
average frequency of about f = 0.08 Hz, with the largest amplitude in the spectrum.
The third IMF characterizes the other dominant carrier wave and oscillates around
f = 0.06 Hz. As pointed out before, mode c1 contains the high frequency part of the
signal and is therefore more or less absent in the Hilbert spectrum since the upper
frequency limitation of this plot is set to f = 0.20 Hz. It seems obvious that the model
proposed by Yuen and Lake could be validated generally with Hilbert spectral
analysis. Results from the second data series are shown in Figure 2.22. Both dom-
inating modes c2 and c3 show evidence of larger frequency variations in time domain.
0.20
c1
0.16
Instantaneous frequency (Hz)
0.12
c2
0.08
c3
0.04
c4
0
0 200 400 600 800 1000 1200
Time (s)
Again, mode c4 is almost of constant frequency, located at f = 0.04 Hz, but the
second and third IMFs do exhibit significant frequency fluctuations and hardly
represent an oscillation around an average frequency. The first IMF is almost absent
in the Hilbert spectrum.
ACKNOWLEDGMENTS
The authors want to express their deep appreciation to Sverre Haver, STATOIL,
Norway, for kindly providing the whole New Year’s Eve data set from the Draupner
platform in the central North Sea. This paper was carried out within the framework
of a research project SCHL503/5-1 funded by the Deutsche Forschungsgemeinschaft
(DFG). The authors gratefully acknowledge this financial support.
REFERENCES
1. Clauss, G. 2002. Dramas of the sea: Episodic waves and their impact on offshore
structures. Applied Ocean Engineering 24, 3:147–161.
2. Dean, R. 1990. Freak waves: A possible explanation. In: Water Wave Kinematic, A.
Torum & O. T. Gudmestad, Eds. Dordrecht: Kluwer, pp. 743–756.
3. Sand, S. E., Ottesen Hansen, N. E., Klinting, P., Gudmestad, O. T., & Sterndorf, M.
J. 1989. Freak wave kinematics. In: Water Wave Kinematic, A. Torum & O. T.
Gudmestad, Eds. Dordrecht: Kluwer, pp. 535–549.
4. Rozario, J. B., Tromans, P. S., Taylor, P. H., & Efthymiou, M. 1993. Comparison of
loads predicted using “Newwave” and other wave models with measurements on the
term structure. Wave Kin. Env. Forces 29, 143–159.
5. Lavrenov, I. V. 1998. The wave energy concentration at the Agulhas current off South
Africa. Natural Hazards 17, ASCE, 117–127.
6. Trulsen, K., & Dysthe, K. 1997. Freak waves: A three-dimensional wave simulation.
Proc. 21st Symp. Naval Hydrodynamics, pp. 550–558.
7. Yasuda, T., & Mori, N. 1997. Occurrence properties of giant freak waves in sea area
around Japan. J. Water, Port, Coast. Ocean Eng. 123, ASCE, 4:209–213.
8. Mori, N., Yasuda, T., & Nakayama, S. I. Statistical properties of freak waves observed
in the Sea of Japan. 10th Int. Offshore Polar Eng. Conf. (ISOPE), Vol. 3, pp. 109–115.
9. Schlurmann, T., Lengricht, J., & Graw, K.-U. 2000. Spatial evolution of laboratory
generated freak wave in deep water depth. 10th Int. Offshore Polar Eng. Conf.
(ISOPE), Vol. 3, pp. 54–59.
10. Claus, G., & Steinhagen, U. 1999. Numerical simulation of nonlinear transient waves
and its validation by laboratory data. 9th Int. Offshore Polar Eng. Conf. (ISOPE),
Vol. 3, pp. 368–375.
11. Kway, J., Loh, Y., & Soon, C. E. 1998. Laboratory study of deep-water breaking
waves. Ocean Eng. 25, Elsevier Sc., pp. 657–676.
12. Haver, S., & Andersen, O. J. 2000. Freak waves: Rare realizations of a typical
population or typical realizations of a rare population? 10th Int. Offshore Polar Eng.
Conf. (ISOPE), Vol. 3, pp. 123–130.
13. Schlurmann, T. 1999. Nonlinear interaction processes of multichromatic wave groups
and the spatial propagation of freak waves. Ph.D. thesis, University of Wuppertal (in
German).
14. Onorato, M., Osborne, A. R., Serio, M., & Damiani, T. 2001. Occurrence of freak
waves from envelope equations in random ocean wave simulation. In: Rogue Waves
2000, Édition Ifremer, ISBN: 2-84433-063-0, pp. 181–191.
15. Pelinovsky, E., Kharif, C., Talipova, T., & Slunyaev, A. 2001. Nonlinear wave focus-
sing as a mechanism of the freak wave generation in the ocean. in: Rogue Waves
2000, Édition Ifremer, ISBN: 2-84433-063-0, pp. 193–204.
16. Huang, N. E., Shen, Z., Long, S., Wu, M. C., Shih, H. H., Zheng, Q., Yen, N.-C.,
Tung, C. C., & Liu, H. H. 1998. The empirical mode decomposition and Hilbert
spectrum for nonlinear and non-stationary time series analysis. Proc. R. Soc. London
A 454, 903–995.
17. Huang, N. E., Shen, Z., & Long, S. 1999. A new view of nonlinear water waves: The
Hilbert spectrum. Annu. Rev. Fluid Mech. 31, 417–457.
18. Flandrin, P., Rilling, G., & Gonçalves, P. 2003. Empirical mode decomposition as a
filter bank. IEEE Signal Processing Letters Vol. 11, No. 2, pp. 112–114.
19. Wu, Z., & Huang, N. E. 2002. A study of the characteristics of white noise using the
empirical mode decomposition method. Submitted to Proc. R. Soc. London A.
20. List of Sea State Parameters. 1986. International Association for Hydraulic Research
(IAHR). Suppl. to Bulletin No. 52, General Secretary of PIANC, Belgium.
21. Kjeldsen, S. P. 1997. Examples of heavy weather damages caused by giant waves.
Techno Marine 10, 24–28.
22. Pierson, W. J., & Moskowitz, L. 1964. A proposed spectral form for fully developed
wind seas based on the similarity theory of S. A. Kitaigordskii. J. Geophysical Res.
69, 5181–5190.
23. Phillips, O. M. 1977. The Dynamics of the Upper Ocean (2nd ed.). London: Cam-
bridge University Press.
24. Olagnon, M., & van Iseghem, S. 2000. Some cases of observed rogue waves and an
attempt to characterize their occurrence. In: Rogue Waves 2000, Édition Ifremer,
ISBN: 2-84433-063-0, pp. 105–116.
25. Trulsen, K. 2000. Simulating the spatial evolution of a measured time series of a
freak wave. In: Rogue Waves 2000, Édition Ifremer, ISBN: 2-84433-063-0, pp.
265–273.
26. Meyer, Y. 1993. Book review: “An introduction to wavelets” by C.K. Chui and “Ten
lectures on wavelets” by I. Daubechies. Bull. Am. Mathematical Soc. 28, 2:350–360.
27. Liu, P. L. 2000. Is the wind wave frequency spectrum outdated. Ocean Eng. 5,
577–588.
28. Liu, P. L. 2000. Wave grouping characteristics in nearshore Great Lakes. Ocean Eng.
27, 11:1221–1230.
29. Liu, P. L. 2000. Wavelet transform and new perspective on coastal and oceanic
engineering data analysis. In: Advances in Coastal and Ocean Engineering Book
Series, Vol. 6, Phillip L. F. Liu, Ed. World Scientific, pp. 57–101, Singapore.
30. Chien, H., Chuang, L., & Kao, C. C. 1999. A study on mechanisms of nearshore
rabid waves, Proc. 2nd German-Chinese Joint Sem. Recent Dev. Coastal Eng., pp.
469–483.
31. Schlurmann, T. 2001. The empirical mode decomposition and the Hilbert spectra to
analyse embedded charactersitic oscillations of extreme waves. In: Rogue Waves 2000,
Édition Ifremer, ISBN: 2-84433-063-0, pp. 157–165.
32. Schlurmann, T. 2002. Spectral analysis of nonlinear water waves based on the Hil-
bert-Huang transformation. J. Offshore Mech. Artic Eng. 124, ASME, 22–27.
33. Huang, N. E., Wu, M.-L. C., Long, S. R., Shen, S. S. P., Qu, W., Gloersen, P., & Fan,
K. L. 2003. A confidence limit for the empirical mode decomposition and Hilbert
spectral analysis. Proc. R. Soc. London A, FirstCite e-publishing, Vol. 459, pp. 2317–
2345.
34. Rilling, G., Flandrin, P., & Gonçalves, P. 2003. On empirical mode decomposition
and its algorithms. Proc. IEEE-EURASIP, Workshop on Nonlinear Signal and Image
Processing, Grado, Italy, 8–11 June 2003 (in press).
35. Veltcheva, A. D. 2002. Wave and group transformation by a Hilbert spectrum. Coastal
Eng. J. 44, 4:283–300.
36. Dätig, M., & Schlurmann, T. 2003. Performance and limitations of the Hilbert-Huang
transformation with short application to water waves. Submitted to Ocean Engineer-
ing.
37. Benjamin, T. B., & Feir, J. E. 1967. The disintegration of wave trains on deep water,
Part I: Theory. J. Fluid Mech. 27, 417–430.
38. Lake, B. M., & Yuen, H. C. 1978. A new model for nonlinear wind waves. Part I:
Physical model and experimental evidence. J. Fluid Mech. 88, 33–62.
3 Applications of
Hilbert-Huang Transform
to Ocean-Atmosphere
Remote Sensing
Research
Xiao-Hai Yan, Young-Heon Jo,
Brian Dzwonkowski, and Lide Jiang
CONTENTS
ABSTRACT
The Hilbert-Huang transform (HHT) is a newly developed method for analyzing non-
linear and nonstationary processes. Its application in oceanography and oceanatmo-
sphere remote sensing research is still in its infancy. In this chapter, we briefly introduce
the application of this method in oceanatmosphere remote sensing data analyses and
present a few examples of such applications.
59
3.1 INTRODUCTION
Spectral analysis is a very useful tool to analyze a time series signal. However, this
method does not fully describe a data set that changes with time. The spectrum gives
us the frequencies that exist over the entire duration of the data set. On the other
hand, time–frequency analysis allows us to determine the frequencies at a particular
time. Hence, the fundamental idea of time–frequency analysis is to understand and
describe phenomena where the frequency content of a signal is changing in time.
Scientists traditionally use short Fourier transform by sliding the window along
the time axis to get a time–frequency distribution. Since it relies on the traditional
Fourier spectral analysis, one has to assume the data to be piecewise stationary.
Currently, the most famous time–frequency analysis method is wavelet transform.
The most common method used is Morlet wavelet, defined as Gaussian enveloped
sine and cosine wave groups with 5.5 waves (1). The problem with Morlet wavelet
is the leakage generated by the limited length of the basic wavelet function, which
makes the quantitative definition of the energy–frequency–time distribution difficult.
Once the basic wavelet is selected, one has to apply it to analysis of all the data (2).
Recently Huang et al. (3) introduced a new and potentially more robust method
for time–frequency analysis. This method, the empirical mode decomposition–Hil-
bert-Huang transform (EMD-HHT), is applicable to both nonstationary and nonlin-
ear signals. In real ocean and ocean atmosphere coupling, most processes are non-
linear and nonstationary. One example is that at the onset of El Nino: nonlinear
Kelvin waves carry warm water from the western Pacific to the east (4). This process
is exhibited as a nonlinear pattern in altimeter data. For this reason, we use the
EMD-HHT technique in our El Nino study and in many of our other studies.
Since the EMD-HHT is relatively new to the ocean remote sensing community,
a brief summary of the technique based on Huang et al. (3) is given in this section.
Basically, the EMD-HHT method requires two steps in analyzing the data. The first
step is to decompose time series data into a number of intrinsic mode functions
(IMFs). These functions must satisfy the following two conditions: (a) within the
entire data set, the total number of extrema (as a function of time) and the total
number of zero-crossings must either be equal or differ at most by one, and (b) at
any point, the mean value of the envelope defined by the local minima (as a function
of time) and the envelope defined by the local maxima (as a function of time) must
be zero. The second step is to apply the Hilbert transform to the decomposed IMFs
and construct the energy–frequency–time distribution, designated as the Hilbert
spectrum. The presentation of the final results of the time–frequency analysis is
similar to the wavelet transform method, which is a spectrogram (time–fre-
quency–energy plot). For clarity, a spectral analysis of the corresponding signal is
shown next to the spectrogram when we present our results in the next sections.
The decomposition of the time series data (i.e. H(t)) into IMFs uses separately
defined envelopes of local maxima and minima. Once the extrema are identified, all
the local maxima are connected by a cubic spline to form the upper envelope. The
procedure is repeated for the local minima to produce the lower envelope. Their
mean is designated as m1(t), and the difference between the time series data and
m1(t) is the first component, h1(t). One can repeat this procedure k times, until hk(t)
is an IMF. Then, hk(t) = c1(t) is the first IMF component of the data. c1(t) should
contain the finest scale or the shortest period component of the signal. Then, c1(t)
is separated from the data, and the process is repeated until either the component
cn(t) or the residue rn(t) becomes so small that it is less than a predetermined value,
or when the residue rn(t) becomes a monotonic function from which no IMF can be
extracted.
After all IMFs have been determined, one can check the original data with the
sum of the IMF components
()
H t = ∑ c (t ) + r (t ) .
i =1
1 n (3.1)
Thus, a decomposition of the data into n empirical modes and a residue rn(t) is
achieved. The rn(t) can be either the mean trend or a constant.
After IMFs of the data have been generated, the next step is to apply the Hilbert
transform to each IMF time series. For an arbitrary time series X(t), one can define
its Hilbert transform, Y(t), as:
()
∞
( ) dτ .
X τ
∫
1
Y t = (3.2)
π t−τ
−∞
With this definition, X(t) and Y(t) form a complex conjugate pair, so one has an
analytic signal Z(t) as
() ()
Z t = X t + iY t , () (3.3)
() ()
a t = X 2 t + Y 2 t 2 , () (3.4)
()
Y t
θ t = arctan .
()
()
(3.5)
X t
ω=
dθ t ( ). (3.6)
dt
Once the instantaneous frequency (as function of time) of each IMF has been
generated, the final result of the time–frequency analysis is similar to the wavelet
transform method, time–frequency–energy plot, or spectrogram. The energy density
similar to Fourier transform can be generated by summing the spectrogram for the
whole time series along a constant frequency for each frequency.
From the definition of the phase function and instantaneous frequency, we can
clearly see that for a simple function such as a = sin(t), the Hilbert spectrum is
simply cos(t), and the phase function is a monotonic straight line. Hence, the
spectrogram is simply a horizontal line for a whole time along fixed frequency, and
its spectrum is simply a single peak at that particular frequency.
In the next sections, we describe a few examples of applications of the HHT in
ocean-atmosphere remote sensing data processing and research. These examples
include an analysis of TOPEX/Poseidon sea level anomaly interannual variation
using HHT and empirical orthogonal function (EOF), application of HHT to ocean
color remote sensing of the Delaware Bay, and Mediterranean outflow and Meddies
determined from satellite multisensor remote sensing.
0.3 0.1
0.25 0
0.2 −0.1
0.15 −0.2
93 94 95 96 97 98 99 00 01 02
0.1
Residual of EOF-1
0.08
0.05
0.06
0
0.04
−0.05
0.02
−0.1 0
−0.15 −0.02
93 94 95 96 97 98 99 00 01 02 93 94 95 96 97 98 99 00 01 02
FIGURE 3.1 The spatial (upper) and temporal (lower) interannual EOF-1 (left) and EMD
of the temporal mode (right) of the T/P-SLA.
Due to its important role in interannual variation, EOF1 was further analyzed
using the HHT. Empirical mode decomposition (EMD) was applied, and EOF1 was
decomposed into three IMFs plus a residue (Figure 3.1). The temporal mode of
EOF1 shows a peak in late ‘97, which correspond to the strong 1997–1998 El Niño.
To examine the effect of ENSO events on EOF1, the Southern Oscillation Index
(SOI) was also used as a reference by decomposing it using EMD. The IMF4 of the
SOI appeared to be a smoothed curve of the SOI, and Salisbury and Wimbush (6)
used this to predict future ENSO events.
Here a comparison was made between the EOF-1 and the SOI by calculating
the correlation coefficient between the IMF-3 of the EOF-1 and IMF-4 of the SOI.
The correlation coefficient was found to be as high as 0.85 (Figure 3.2). Considering
this, the EOF-1 can be considered to behave under the impact of the ENSO, and we
can further infer that the ENSO contributes to about one third of the interannual
SLA variation.
Moreover, we can reexamine the spatial mode of EOF1 (Figure 3.1) to see the
global impact of the ENSO events on human activities. Besides the coastal areas of
America, Australia, and Indonesia, which are most strongly and directly affected,
those of east Africa and Japan also are undergoing remarkable ENSO variation. The
Correlation = 0.85
0.1
0.05
−0.05
−0.1
−0.15
93 94 95 96 97 98 99 00 01 02 03
FIGURE 3.2 Correlation between negative IMF-4 of SOI (red) and IMF-3 of EOF-1 (blue)
is 0.85. It suggests that the primary interannual EOF is impacted by ENSO.
impact even spread as far as Antarctica. On the other hand, coastal areas of the
Atlantic Ocean are less affected by the ENSO events.
In addition, the HHT spectrum of the EOF1 was investigated to extract the
frequency information (Figure 3.3). The frequency distribution of the EOF1 is
between 0 and 4 cycle/yr. The 2.5 to 4 cycle/yr range corresponds to the spectrum
of the IMF1, which behaves in a relatively random pattern. The 0.5 to 2 cycle/yr
range corresponds to the spectrum of IMF2, which has more energy within the
frequency range 0.5 to 0.8 cycle/yr. The 0.2 to 0.4 cycle/yr range has much higher
energy, which is shown as a dark red curve at the bottom part of the spectrum. This
curve corresponds to the frequency feature of IMF3, i.e., a period of 2.5 to 5 yr,
which can be considered as the typical frequency of ENSO events. The lowest
frequency part of the 0 to 0.1 cycle/yr range corresponds to the frequency of the
residue, which is the longterm trend. It has a period of tens of years or longer and,
therefore, requires much longer time series to analyze.
4
Hilbert Spectrum (cycles per year)
3.5
2.5
1.5
0.5
93 94 95 96 97 98 99 00 01 02
Year
FIGURE 3.3 The HHT spectrum of EOF-1. The dark red curve indicates high energy at
frequencies between 0.2 and 0.4 cycle/yr, which corresponds to the Hilbert spectrum of IMF-3
of interannual EOF-1. It indicates that ENSO events have a typical period of 2.5 to 5 yr.
source. However, use of this data source is dependent on the premise that there are
discernable relationships between reflectance exiting the water column and the
constituents in it. A primary constituent of study has been chlorophyll-a on account
of its connection to phytoplankton and thus to primary production and biomass (7).
As coastal management strategies begin to focus on large-scale ecosystem based
programs, the relationship between chlorophyll-a and primary production and bio-
mass has the potential to provide a cost-effective alternative to traditional ship-based
or point source sampling. Monitoring and assessing the health of ecosystem-size
regions will be much more feasible with satellite-based parameters due to the
frequent repeat periods and large spatial areas covered by satellite sensors. Thus,
the quantitative assessment of water constituents from satellite-based ocean color
data holds great promise for implementing dynamic management strategies. How-
ever, to interpret ocean color data appropriately, a full understanding of the period-
icity of chlorophyll concentrations in a given area is essential. The purpose of this
study is to examine the seasonal and interannual chlorophyll-a cycles from satellite
data of the coastal region at the mouth of the Delaware Bay.
Although quantifying chlorophyll concentration from satellite ocean color data
has been successful in the open ocean, coastal areas (case II water) can still present
problems. Water constituents associated with coastal regions, such as excessive
chlorophyll-a concentrations, suspended sediments, and colored dissolved organic
25
Chlorophyll-a (mg/m3)
20
15
10
0
1998 1999 2000 2001 2002 2003
1
NAO Index
−1
−2
−3
1998 1999 2000 2001 2002 2003
Years
FIGURE 3.4 The time series of the monthly means of the chlorophyll-a concentrations (top
panel) and the North Atlantic Oscillation index (bottom panel) from September 1997 to July
2003.
0.4
0.35
0.35
0.3 0.3
Frequency (cycles/month)
0.25 0.25
0.2 0.2
0.15 0.15
0.1 0.1
0.05
0.05
0
1998 1999 2000 2001 2002 2003 0 10 20
Years Energy
FIGURE 3.5 The Hilbert spectrum (left panel) and marginal spectrum (right panel) for the
IMFs of the chlorophyll-a times series.
0.5 2
0 0
−0.5 −2
1998 1999 2000 2001 2002 2003
Years
FIGURE 3.6 Interannual variation (IMF 4) for both the NAO index (blue line) and the
chlorophyll-a (green line) time series.
the Subpolar Low. Both positive and negative phases affect basin-wide variations in
the North Atlantic jet stream and storm tracks, as well as in large-scale modulations
of the typical patterns of zonal and meridional heat and moisture transport. These
alterations will consequently affect temperature and precipitation patterns and can
extend from eastern North America to western and central Europe (17). Furthermore,
recent studies have shown links between the NAO and hydrographic properties along
the northeast Atlantic shelf slope and the Gulf of Maine, and zooplankton and
chlorophyll-a concentrations within the Gulf of Maine (18). Thus, the interannual
frequency of the NAO was examined.
Monthly mean data of the NAO index from September 1997 to July 2003 were
obtained from the NOAA Climate Prediction Center. The NAO index data are shown
in Figure 3.4. This time series was also decomposed into its component IMFs in
order to isolate the NAO interannual mode. After decomposing the NAO index time
series, the fourth IMF contained an interannual signal that exhibited features very
similar to the fourth chlorophyll-a IMF (Figure 3.6).
To quantify this relationship, cross-correlation analysis was performed on the
IMF 4 of both the chlorophyll-a concentrations and the NAO index. The results are
shown in Figure 3.7. This study focused only on positive lag time (chlorophyll-a
time series following the NAO time series), because it would not be sensible to
assume that negative lags (chlorophyll-a time series leading the NAO time series)
were meaningful. The correlation analysis reveals that the NAO and chlorophyll-a
0.6
0.4
Correlation coefficient
0.2
−0.2
−0.4
−0.6
−0.8
−1
0 10 20 30 40 50 60 70
Time lag
Chlorophyll-a follows NAO Index for lag > 0
FIGURE 3.7 Correlation coefficient (blue line) and the related 95 % confidence interval (red
dashed line). The maximum correlation coefficient with a positive lag (i.e. chlorophyll-a
interannual variation following NAO index interannual variation) is –0.67, corresponding to
a lag of 10 months.
that was significant at the 95% confidence level. This relatively high correlation
coefficient occurred at a lag of 10 months, which suggests that the possible inter-
annual link between chlorophyll-a concentrations at the mouth of the Delaware Bay
and the NAO index may result from chlorophyll-a concentrations responding to
variations in the NAO index. Finally, the observations of both the weakening of the
seasonal signal and the suggested teleconnection of the interannual variation between
chlorophyll-a concentrations and the NAO are topics worthy of future investigation
that would have been difficult to identify with traditional time series analysis tech-
niques.
η′T = α
∫ ∆Tdz.
−400
∆T is the temperature difference between two different depths (dz), and α is the
thermal expansion coefficient, which is calculated as function of salinity (S) and
water pressure (P) based on empirical measurements (24);
1 ∆ρ
α=− .
ρ0 ∆T
( s , p)
The spatial resolution of XBT data is 73° × 61° longitude/latitude, which is inter-
polated to 1° × 1° longitude/latitude data for this study. It is possible that errors
occurred due to interpolation of the time rates of change of the integrated upper
ocean heat storage anomaly (5 W/m2, on average); this is discussed by White and
Tai (25). We also checked interpolation error using 1° × 1° longitude/latitude opti-
mum interpolation sea surface temperature (OISST). We compared interpolated sea
surface temperature (SST) from XBT data and OISST by calculating correlation
coefficients. Correlation coefficients between OISST and SST from XBT were over
95% with RMS 0.2°C, except for two small areas, and correlation coefficient 80%
with RMS 1.2°C. We conclude that interpolation error of XBTs is less than 1°C at
the sea surface.
Because of scarcity of temporal and spatial salinity data, we estimated the effect
of η′S in the upper layer indirectly. Two different correlation and RMS calculations
were made: the first correlation is between η′T and OISST, and the second correlation
is between η′Total and OISST. Good temporal and spatial agreement between SST
and η′Total or η′T suggests that a robust regression between fields may have some
physical significance with respect to thermal expansion, but a low correlation with
high RMS may have some other variability in the mixed layer or below the mixed
layer due to the salinity. It appears that the difference is due to a change in ocean
salinity, which is reflected in the T/P sea level measurements but not in the η′T
measurements. The first correlation coefficients were all over 90% with smaller RMS
than the second ones, but the second ones were also over 60% to 80%, with larger
RMS than the first ones. This result is caused by a warm and salty core below the
mixed layer. Using Levitus 94 (26), the vertical structure of the salinity was examined
to see the salinity distribution. The whole upper layer in our study domain has a
uniform salinity distribution above 1000 m. We conclude that the anomaly of η′S is
spatially uniform with only small variability.
Wind effect on the η′Total signal was considered using scatterometer data from
European Remote Sensing Satellite-1/2 (ERS-1/2) from January 1993 to December
1999. First we calculated the correlations between wind stress curl and the η′Total for
the temporal variation. Negative correlation would be expected; rising sea level is
associated with negative wind stress curl. Correlation coefficients showed a relation
of about –30% in our study area, with southern areas having larger RMS than
northern areas. Second, we considered the magnitude of sea level variation due to
wind stress based on a linear barotropic vorticity equation, i.e.,
d η − g′
≈ ∇x τ,
dt f0ρg
where η is the sea level height, ρ is the water density, g′ is the reduced gravity, f0
is the Coriolis parameter, and τ is the wind stress. We also estimated ηW ′ by using
NASA’s scatterometer: NSCAT (August 1996 to June 1997) and Quikscat (June
1999 to present). The mean sea level variation due to wind stress was around 1 cm.
The result of ∆η′ = η′T P − ηUL′ is apparently some variation of the vertical water
column, which is a response according to the different incoming or outgoing water
mass under the upper layer. Sea level variation due to salinity (temperature) change
using mass conservation with 800 m thick is around 60 cm (32 cm), with 1-ppt
differences from background fluid (2°C), derived from the salt expansion coefficient
7.5 × 10–4/°C and thermal expansion coefficient 2.0 × 10–4/°C, respectively. The
bottom pressure (deep current) variation is negligible at this region (26, 27). How-
ever, since the Meddy was a weakly stratified result of extensive salt fingering (28),
there were regions of high stratification above and below the lens, where the back-
ground isopycnal surface becomes increasingly broader as it moves above the Meddy
toward the sea surface. Because of this isopycnal compensation, the O&M are not
revealed in the η′Total signal. This is why we cannot detect a Meddy with the altimeter
observation alone.
We computed the absolute differences of the sea surface height anomaly of ηUL ′
and η′Total to examine the trajectories of the O&M from January 1993 to December
1999. Figure 3.8 is an example of such trajectories from July 1998 to June 1999.
One can see a strong signature (|∆η′|) toward west and south from July to November.
Generally, southward Meddies are formed near 36°N by separation of the frictional
boundary layer at sharp corners (29) and in the Canary Basin (30, 31, 32). The
southward travelling Meddies can be explained by the strongest low frequency zonal
motions driven by baroclinic instability (33) and the influence of the neighboring
mesoscale features (cyclonic vortices or Azores Current meanders) in the regions
(34). The northward O&M over 36° to 40°N are also shown from August to October.
Furthermore, a strong signature over 40°N in February is considered to be the
influence of the returning Gulf Stream. From March to June the weak O&M signa-
tures are found because of the weak salinity deviation in 1000 m depth examined
from climatological data.
To investigate the reasons for the change in the direction of propagation of
Meddies, the stream function (ψ) was computed by using the T/P altimeter. This
representation of the stream function permits observations of the interactions
between the sea surface gradient and the Meddies’ propagation. The computation
of the sea surface height anomaly η′ in terms of the usual dynamic variables is
straightforward, if the flow is assumed to be quasigeostrophic: The stream function
(ψ) is defined as
40 40
35 35
30 30
30W 25W 20W 15W 10W 05W 30W 25W 20W 15W 10W 05W
Aug 1998 Feb 1999
45 45
40 40
35 35
30 30
30W 25W 20W 15W 10W 05W 30W 25W 20W 15W 10W 05W
Sep 1998 Mar 1999
45 45
40 40
35 35
30 30
30W 25W 20W 15W 10W 05W 30W 25W 20W 15W 10W 05W
Oct 1998 Apr 1999
45 45
40 40
35 35
30 30
30W 25W 20W 15W 10W 05W 30W 25W 20W 15W 10W 05W
Nov 1998 May 1999
45 45
40 40
35 35
30 30
30W 25W 20W 15W 10W 05W 30W 25W 20W 15W 10W 05W
Dec 1998 Jun 1999
45 45
40 40
35 35
30 30
30W 25W 20W 15W 10W 05W 30W 25W 20W 15W 10W 05W
0 1 2 3 4 5 6 7 8
(
FIGURE 3.8 Calculation of the ∆η′ ∆η′ = η′Total − ηUL
′ )
from July 1998 to June 1999.
40 40
35 35
30 30
30W 25W 20W 15W 10W 05W 0W 30W 25W 20W 15W 10W 05W 0W
40 40
35 35
30 30
30W 25W 20W 15W 10W 05W 0W 30W 25W 20W 15W 10W 05W 0W
FIGURE 3.9 Comparisons of Meddy trajectories with stream functions (Equation 3.7) for
January, April, July, and October 1994.
ψ= g η′Total , (3.7)
f 0
3rd EMD mode using HHT at A(30W, 30N) with other places
5
B(20W, 30N)
−5
93 94 95 96 97 98 99 00
5
C(10W, 30N)
−5
93 94 95 96 97 98 99 00
5
D(10W, 37N)
−5
93 94 95 96 97 98 99 00
5
E(10W, 42N)
−5
93 94 95 96 97 98 99 00
5
F(20W, 42N)
−5
93 94 95 96 97 98 99 00
5
G(30W, 42N)
−5
93 94 95 96 97 98 99 00
5
H(30W, 37N)
−5
93 94 95 96 97 98 99 00
FIGURE 3.10 A comparison of the third EMD modes using HHT are shown at given
locations. The solid curve is for location A (refer to Figure 3.11 for the location) in all panels
for comparison. The dotted curve is the individual EMD modes for those locations.
G F E 4
3.5
40
2.5
H D
35
1.5
0.5
30 A B C 0
30W 25W 20W 15W 10W 05W
FIGURE 3.11 Comparisons of annual mean |∆η′| signal from two float experiments in 1994.
The gray scale and contours show the annual mean of the |∆η′| signal estimated from our
method. All Meddies were discovered during the AMUSE experiment (stars) and SEMA-
PHORE experiment (circles) in 1994.
To find out the dominant signal of the sea surface interaction with the Meddies,
the HHT was applied to their EMD modes. We chose eight places to investigate the
dominant signal in our study area. Figure 3.10 showed that Place Group 1 (B, C,
and D, shown in Figure 3.11) and Group 2 (H and G, shown in Figure 3.11) had a
similar signal. However, Group 3 (E and F, shown in Figure 3.11) had slightly
different signals from Groups 1 and 2. Consequently, there were seasonal fluctuations
between location A and Group 3. To compute the dominant signal of the power for
locations A and E, HHT was also employed; this is shown in Figure 3.12. In both,
the dominant frequency was around f = 0.082 (1 year). However, location A had a
lower frequency, f = 0.03 (33.3 months), which seems to indicate that wind stress
with 33.3-month period produces sea surface forcing. The surface variations produce
the baroclinic instability on the Meddies, which is related to southward translation.
The southward translation of the Meddies due to the baroclinic effects were consis-
tent with that discussed by Müler and Siedler (36) and by Käse and Zenk (37). If
the current is vertically sheared in a stratified fluid, baroclinic instability can occur.
Figure 3.12 also demonstrates that the comparison between field observations from
two float experiments in 1994 and our computation was excellent.
0.4 0.4
0.35 0.35
0.3 0.3
0.25 0.25
0.2 0.2
0.15 0.15
0.1 0.1
0.05 0.05
0 0
0 10 20 30 40 50 60 70 80 0 20 40
0.4 0.4
0.35 0.35
0.3 0.3
0.25 0.25
0.2 0.2
0.15 0.15
0.1 0.1
0.05 0.05
0
0 10 20 30 40 50 60 70 80 0 20 40
3.5 CONCLUSION
Three examples of applications of the HHT method in ocean-atmosphere remote
sensing research were illustrated in this chapter. These examples show that the HHT
method is indeed a potentially very useful and powerful tool for ocean engineering
and science studies.
ACKNOWLEDGMENTS
This research was supported partially by the National Aeronautics and Space Admin-
istration (NASA) through Grant NAG5-12745 and NGT5-40024, by the Office of
Naval Research (ONR) through Grant N00014-03-1-0337, and by the National
Oceanic and Atmospheric Administration (NOAA) through Grants NA17EC2449
and NA96RG0029.
REFERENCES
1. Chan, Y. T. (1995). Wavelet Basics. Boston, MA: Academic.
2. Huang, N., Long, S., Shen, Z. (1996). The mechanism for frequency downshift in
nonlinear wave evolution. Adv. Appl. Mech. 32:59–111.
3. Huang, N., Shen, Z., Long, S., Wu, M., Shin, H., Zheng, Q., Yen, N.-C, Tung, C.,
Liu H. (1998). The empirical mode decomposition and Hilbert spectrum for nonlinear
and nonstationary time seriesanalysis. Proc. Royal Soc. Mathematical, Phys. Eng.
Science 454:903–995.
4. Yan, X.-H., Zheng, Q., Ho, C.-R., Tai, C.-K., Cheney, R. (1994). Development of the
pattern recognition and the spatial integration filtering methods for analyzing satellite
altimeter data. Remote Sens. Environ. 48:147–158.
5. Cheney, R., Miller, L., Agreen, R., Doyle, N., Lillibridge, J. (1994). TOPEX/POSEI-
DON: 2 cm solution. J. Geophys. Res. 99:24555–24563.
6. Salisbury, J. I., Wimbush, M. (2002). Using modern time series analysis techniques
to predict ENSO events from the SOI times series. Nonlinear Proc. Geophys.
9:341–345.
7. Kiddon, J. A., Paul, J. F., Buffum, H. W., Strobel, C. S., Hale, S. S., Cobb, D., Brown,
B. S. (2003). Ecological condition of U.S. mid-Atlantic estuaries. 1997–1998. Marine
Pollution Bulletin 46(10):1224–1244.
8. Keiner, L. E., Yan, X-H. (1998). A neural network model for estimating sea surface
chlorophyll and sediments from thematic mapper imagery. Remote Sens. Environ.
66(2):153–165.
9. Zhang, Y., Pulliainen, J., Koponen, S., Hallikainen, M. (2002). Application of an
empirical neural network to surface water quality estimation in the Gulf of Finland
using combined optical data and microwave data. Remote Sens. of Environ.
81:327–336.
10. Baruah, P. J., Oki., K., Nishimura, H. (2000). A neural network model for estimating
surface chlorophyll and sediment content at Lake Kasumi Gaura of Japan. In: Con-
ference Processing of GIS Development. Taipei, Taiwan, December.
11. Buckton, D., O’Mongain, E., Danaher, S. (1999). The use of neural networks for the
estimation of oceanic constituents based on the MERIS instrument. Int. J. Remote
Sensing 20(9):1841–1851.
12. Keiner, L. E., Yan, X.-H. (1997). Empirical orthogonal function analysis of sea surface
temperature patterns in Delaware Bay. IEEE Trans. Geoscience Remote Sensing
25(5):1299–1306.
13. Zhang, T., Fell, F., Liu Z.-S., Preusker, R., Fischer, J., He, M.-X. (2003). Evalating
the performance of artificial neural network techniques for pigment retrieval from
ocean color in Case I Water. J. Geophys. Res. 108(C9):3286–3298.
36. Müler, T. J., Siedler, G. (1992). Multi-year current time series in the eastern North
Atlantic Ocean, J. Mar. Res. 50:63–98.
37. Käse, R. H., Zenk W. (1996). Structure of the Mediterranean water and Meddy
characteristics in the northeastern Atlantic. In: Ocean Circulation and Climate:
Observation and Modeling the Global Ocean, W. Krauss, Ed. Berlin: Gebrüder
Bornträger, Berlin, pp. 365–395.
4 A Comparison of the
Energy Flux Computation
of Shoaling Waves
Using Hilbert and
Wavelet Spectral
Analysis Techniques
Paul A. Hwang, David W. Wang, and
James M. Kaihatu
CONTENTS
83
ABSTRACT
4.1 INTRODUCTION
Fourier transform decomposes a nonlinear waveform into a fundamental frequency
component and its harmonics. As a result, the energy of a nonlinear wave is spread
into higher frequencies. This may produce difficulties in the interpretation of wave
propagation and quantification of wave dynamic properties such as the energy flux
of the wave field. Recently, Norden Huang and his colleagues developed a new
analysis technique, the Hilbert-Huang transformation (HHT). Through analytical
examples, they demonstrated the excellent frequency and temporal resolution of
HHT for analyzing nonstationary and nonlinear signals [1, 2].
With the HHT analysis, the physical interpretation of nonlinearity is frequency
modulation, which is fundamentally different from harmonic generation. The HHT
spectrum therefore retains its energy density near the fundamental frequency of the
wave motion in comparison with the FFT or wavelet spectrum [3]. In this article,
we briefly describe the HHT technique and investigate its use in the calculation of
the energy flux of ocean waves. The resulting HHT spectrum is compared with the
counterpart obtained by the wavelet method. The wavelet technique is based on
Fourier spectral analysis but uses adjustable frequency-dependent window functions
— the mother wavelets — to provide temporal/spatial resolution for nonstationary
signals [4–7]. As expected, the Fourier-based analysis interprets wave nonlinearity
in terms of harmonic generation; thus the spectral energy leaks to higher frequency
components. The HHT interprets wave nonlinearity as frequency modulation, and
the spectral energy remains near the base frequency. The computed energy flux from
the HHT method is much higher than that from the wavelet method.
wave problem [1] and on the rapid change of the oscillation frequencies [11] have
been published and will not be repeated here.
The key ingredient in the HHT is empirical mode decomposition (EMD),
designed to reposition the riding waves at the mean water level. This is achieved
through a sifting process that repeatedly subtracts the local mean from the original
signals. The procedure decomposes the signal into many modes with different
frequency characteristics, and thus also alleviates the problem of sharp frequency
change in the original signal. Huang and his coworkers have provided extensive
discussions on the EMD [1, 2]. From experience, even for very complicated random
signals, a time series can usually be decomposed into a relatively small number of
modes, M < log2 N. Each mode is free of riding waves and is suitable for Hilbert
transformation to yield accurate local frequency of the mode. The spectrum of the
original signal can be obtained from the sum of the Hilbert spectra of all modes.
Extensive tests have been carried out, and the HHT technique proves to deliver very
high frequency and temporal/spatial resolutions in analyzing nonstationary signals.
It is also shown to be able to handle the task of analyzing nonlinear signals produced
by exact solution with modulating oscillation frequencies, whereas Fourier-based
techniques interpret the signals as superposition of harmonics [1–3].
FIGURE 4.1 (a) A 3D surface wave topography of shoaling swell in the coastal region. (b)
The HHT spectrum of the waves shown in (a). (c) The wavelet spectrum of the waves shown
in (a). (d) The water depth from bathymetry database.
()
F x =
∫ S (ω; x )C (ω; x ) dω ,
g (4.1)
3 (a)
F 2
0 200 400 600 800 1000 1200 1400 1600 1800 2000 2200
<EF(HHT)>=−6.373e−004, <EF(wavelet)>=−3.083e−004, offset=30
(b)
0.2
d(F)/dxHHT
−0.2
0 200 400 600 800 1000 1200 1400 1600 1800 2000 2200
0.02 (c)
d(F)/dxWavelet
−0.02
0 200 400 600 800 1000 1200 1400 1600 1800 2000 2200
x (m)
FIGURE 4.2 (a) Energy flux computed from the shoaling swell shown in Figure 4.1. (b) The
source function computed from the HHT spectrum. (b) The source function computed from
the wavelet spectrum.
∂F
=Q . (4.2)
∂x
Figure 4.2b and c shows the derived source function based on the Hilbert and wavelet
spectra, respectively. The magnitude of the Hilbert source term is considerably larger
than the wavelet source term (notice the difference in scale of the two ordinates).
The oscillatory nature of the source function highlights the complex nature of the
wave conditions in the field. The group structure can be introduced by many pro-
cesses, such as wave nonlinearity, interaction among different wave components,
and bathymetry-induced wave scattering. The group structure causes the source
function to fluctuate between positive and negative values, as can be visualized from
Equation 4.1 and Equation 4.2. Further discussion of the group structure will be
presented later.
In appearance, the source functions obtained by the HHT and wavelet methods,
as shown in Figure 4.2b and c, are considerably different. In reality, the apparent
difference reflects the much higher temporal/spatial resolution of the HHT method.
Figure 4.3 shows the results of the running average of HHT computation, with the
wavelet result superimposed. With an averaging bin width of 60 elements, the
running average of the HHT source function is almost identical to the wavelet source
function. The spatial average (excluding the leading and trailing 90 m of the spatial
coverage shown) of the source term is –2.49 × 10–4 for the wavelet processing, –5.64
× 10–4 for the HHT processing, and –4.78 × 10–4 for the running average of the HHT
0.04
0.02
d(F)/dx
−0.02
−0.04
−0.06
−0.08
−0.1
0 200 400 600 800 1000 1200 1400 1600 1800 2000 2200
x (m)
FIGURE 4.3 A comparison of the wavelet source function, the HHT source function, and
the ensemble average of HHT source function with a bin width of 60 spatial elements.
processing. From this limited investigation, we conclude that the energy flux com-
putation may be off by a factor of two using FFT based processing techniques.
As mentioned earlier, harmonic generation has created major difficulties for the
analysis of wave dynamics due to the dispersive nature of water waves. For example,
the phase and group velocities of each free wave component are frequency depen-
dent, yet the harmonics (of nonlinear waves) are not dispersive. Therefore, the energy
flux (the product of group velocity and spectral density) of the harmonics associated
with the nonlinear waves cannot be distinguished from the energy flux of the free
waves at the same frequency, and FFT based processing will always underestimate
the energy flux of the wave field. The underestimation will reflect on the magnitude
of parameters such as the dissipation rate or growth rate of a wave field.
(a)
(b)
FIGURE 4.4 Numerical simulations of shoaling waves. Left panels: linear simulation. Right
panels: nonlinear simulation. (a) Waveform and water depth, (b) HHT spectrum, (c) wavelet
spectrum, and (d) characteristic wave numbers kp and k2.
linear nonnolear
(a) (a)
0.4 0.4
g
E Cg
EC
0.2 0.2
HHT
wavelet
0 0
0 −3
500 1000 1500 0 −3
500 1000 1500
x 10 x 10
2 (b) 2 (b)
d(EF)/dx
d(EF)/dx
0 0
−2 −2
0 −3
500 1000 1500 0 −3
500 1000 1500
x 10 x (m) x 10 x (m)
1 (c) 1 (c)
60
<d(EF)/dx>60
<d(EF)/dx>
0 0
−1 −1
0 500 1000 1500 0 500 1000 1500
x (m) x (m)
FIGURE 4.5 Same as Figure 4.4 but for the energy flux computation. (a) Energy flux, (b)
gradient of energy flux, and (c) same as (b) but phase averaged. A: linear simulation. B:
nonlinear simulation.
scenarios by the HHT processing method is shown in Figure 4.4b. The spectral
density is concentrated in a very narrow frequency band, as expected from a mono-
chromatic wave train. The lack of frequency modulation in the absence of nonlin-
earity is clear from a comparison of Figure 4.4Ab with Figure 4.4Bb. The result
based on the wavelet method is shown in Figure 4.4c. The spectral energy is more
spread out because of the finite window size, and the effect of nonlinearity is
harmonic generation.
The characteristic wave number is typically defined as the peak wave number,
kp, or weighted wave number computed from the spectral moment, kn = ( ∫knS(k)dk/
∫S(k)dk)1/n. With the HHT analysis, kp and kn are almost identical for both linear and
nonlinear scenarios because the spectral energy is confined in a narrow frequency
range. With the wavelet analysis, kp and kn are also very similar in the linear scenario,
but they differ considerably in the nonlinear scenario as a result of harmonic gen-
eration. The results of phase-average kp and k2 for the two scenarios are shown in
Figure 4.4d. Intra-wave frequency modulation as reflected by the oscillation of the
characteristic wave number is clearly seen in the HHT analysis of the nonlinear
simulation; it is also somewhat noticeable but much weaker in the wavelet analysis.
For the linear case, the computed energy flux and the gradient of energy flux
derived from the two processing methods are very similar (Figure 4.5a,b,c). The
HHT results show small intra-wave modulation due to finite amplitude of the wave-
form. For the nonlinear case, the intra-wave modulation is greatly amplified by the
nonlinearity of the wave system. The differentiation process in calculating the gra-
dient of energy flux sometimes produces large spikes in the computation. The results
become very sensitive to small modifications in the data processing procedure. Figure
4.5 shows an example of the processed results.
4.4 DISCUSSIONS
4.4.1 RESOLUTION AND NONLINEARITY
Fourier-based spectral analysis methods have been widely used for studying random
waves. One major weakness of the Fourier-based spectral analysis methods is the
assumption of linear superposition of wave components. As a result, the energy of
a nonlinear wave is spread into many harmonics, which are phase-coupled via the
nonlinear dynamics inherent in ocean waves. In addition to the nonlinearity issue,
strictly speaking Fourier spectral analysis should be used for periodic and stationary
processes only. Wave propagation in the ocean is certainly neither stationary nor
periodic. Using the HHT analysis, the physical interpretation of nonlinearity is
frequency modulation, which is fundamentally different from the commonly
accepted concept associating nonlinearity with harmonic generation. Huang et al.
[1, 2] argued that harmonic generation results from the perturbation method used in
solving the nonlinear equation governing the physical processes, thus the harmonics
are produced by the mathematical tools used for the solution rather than being a
true physical phenomenon. Through analytical examples, they demonstrated the
excellent frequency and temporal resolutions of HHT for analyzing nonstationary
and nonlinear signals. Here we give a few computational examples to illustrate the
points discussed earlier.
Figure 4.6 presents a comparison of HHT and wavelet analysis of three different
cases [3]. Case 1 is an example of an ideal time (t) or space (x) series of sinusoidal
oscillations of constant amplitude; the frequency (f) or wave number (k) of the first
half of the signal is twice that of the second half (Figure 4.6a). The spectra computed
by the HHT and wavelet techniques are displayed in Figure 6b and c, respectively.
The HHT spectrum yields very precise frequency resolution as well as high temporal
resolution in identifying the sudden change of signal frequency at about the half
point of the time series. In comparison, the wavelet spectrum has only a mediocre
temporal resolution of the frequency change. There is also a serious leakage problem,
and the spectral energy of the simple oscillations spreads over a broad frequency
range (the contour interval is 3 dB in the spectral plots).
Case 2 is a single cycle sinusoidal oscillation occurring at the middle of the
otherwise quiescent signal stream (Figure 4.6d). The precise temporal resolution of
the HHT method is clearly demonstrated by the sharp rise and fall of the HHT
spectrum coincident with the transient signal, as illustrated in Figure 4.6e. In com-
parison, the wavelet spectrum is much more smeared, both in the frequency and
temporal resolutions (Figure 4.6f).
Case 3 is a sinusoidal function, y, with its oscillating frequencies subject to
periodic modulation (Figure 4.6g)
0 0 0
y
y
−1 −1 −1
500 1000 1500 500 1000 1500 500 550 600
3dB contours
0.3 (b) 0.2 (e) 0.3 (h)
0.2 0.2
k|f
k|f
k|f
0.1
0.1 0.1
0 0 0
500 1000 1500 500 1000 1500 500 550 600
0.3 (c) 0.2 (f) 0.3 (i)
0.2 0.2
k|f
k|f
k|f
0.1
0.1 0.1
0 0 0
500 1000 1500 500 1000 1500 500 550 600
x|t x|t x|t
FIGURE 4.6 Examples comparing HHT (middle row) and wavelet (bottom row) analysis of
nonlinear and nonstationary signals. (a,b,c) Time series of the harmonic motion with the
oscillation period doubled in the second half, and its HHT and wavelet spectra. (d,e,f) Transient
sinusoidal function of one cycle. (g,h,i) Oscillatory motion with modulated frequencies.
() (
y t = a cos ωt + ε sin ωt , ) (4.3)
d2y
( ) ( )
2 0.5
+ ω + εω cos ωt y − 1− x2 εω 2 sin ωt = 0. (4.4)
dt 2
If the perturbation method is used to solve Equation 4.2, the solution to the first
order of ε is
1
()
y1 t = cos ωt − ε sin 2ωt = cos ωt − ε 1 − cos 2ωt ( ) . (4.5)
2
The HHT spectrum (Figure 4.6h) correctly reveals the nature of oscillatory frequen-
cies of the exact solution (Equation 4.3). In contrast, the wavelet spectrum (Figure
4.6i) shows a dominant component at the base frequency and periodic oscillations
of the second harmonic component.
1 (a)
cosθ (expanded)
Hilbert transform
f(t)
0 sinθ (expanded)
−1
0 50 100 150 200 250 300
1 t (b)
f(t)
−1
0 50 100 150 200 250 300
1 t (c)
exp(−0.01t)cosθ (expanded)
Hilbert transform
f(t)
0 exp(−0.01t)sinθ (expanded)
−1
0 50 100 150 200 250 300
1 t (d)
f(t)
−1
0 50 100 150 200 250 300
t
FIGURE 4.7 An example illustrating the mirror-imaging approach to alleviate edge problems.
These three examples illustrate the excellent temporal (spatial) and frequency
(wave number) resolution of the HHT method for processing nonlinear and nonsta-
tionary signals. The result from Case 3 is especially interesting, as it illustrates the
unique ability of the HHT technique to reveal more accurately the nature of nonlinear
processes with nonstationary oscillation frequencies. Solutions obtained through the
perturbation method and analysis results obtained through Fourier-based techniques
represent those processes as superpositions of harmonic motions.
4.5 SUMMARY
Analyzing nonlinear and nonstationary signals remains a very challenging task.
Many methods developed to deal with nonstationarity are based on the concept of
Fourier decomposition; therefore, all the shortcomings associated with Fourier trans-
formation are also inherent in those methods. The recent introduction of empirical
mode decomposition [1, 2] represents a fundamentally different approach for pro-
cessing nonlinear and nonstationary signals. The associated spectral analysis (HHT)
provides excellent spatial (temporal) and wave number (frequency) resolution for
handling nonstationarity and nonlinearity [1–3]. The HHT spectrum also results in
a considerably different interpretation of nonlinearity (frequency modulation, as
compared to the traditional view of harmonic generation). Applying the technique
to the problems of ocean waves, we found that the spectral function derived from
HHT is markedly different from those obtained by the Fourier-based techniques.
The difference in the resulting spectral functions is attributed to the interpretation
of nonlinearity. The Fourier techniques decompose a nonlinear signal into sinusoidal
harmonics; therefore, some of the spectral energy at the base frequency is distributed
to the higher frequency components. The HHT interprets nonlinearity in terms of
frequency modulation, and the spectral energy remains in the neighborhood of the
base frequency. This results in a considerably higher spectral energy at lower fre-
quencies and a sharper dropoff at higher frequencies in the HHT spectrum in
comparison with the Fourier-based spectra [3]. The energy flux computed by using
the HHT spectrum is much higher than that obtained from the FFT or wavelet
methods.
ACKNOWLEDGMENTS
This work is supported by the Office of Naval Research (Naval Research Laboratory
Program Elements N61153 and N62435). This is NRL contribution BC/7330.
REFERENCES
1. Huang, N. E., Shen, Z., Long, S. R., Wu, M. C., Shih, H. H., Zheng, Q., Yuen, N.
C., Tung, C. C., and Liu, H. H. (1998). The empirical mode decomposition and the
Hilbert spectrum for nonlinear and nonstationary time series analysis. Proc. R. Soc.
Lond. 454A: 903–995.
2. Huang, N. E., Shen, Z., and Long, S. R. (1999). A new view of nonlinear water
waves: The Hilbert spectrum. Annu. Rev. Fluid Mech. 31: 417–457.
3. Hwang, P. A., Huang, N. E., and Wang, D. W. (2003). A note on analyzing nonlinear
and nonstationary ocean wave data. Appl. Ocean Res. 25: 187–193.
4. Shen, Z., and Mei, L. (1993). Equilibrium spectra of water waves forced by inter-
mittent wind turbulence. J. Phys. Oceanogr. 23(9): 2019–2026.
5. Shen, Z., Wang, W., and Mei, L. (1994). Finestructure of wind waves analyzed with
wavelet transform. J. Phys. Oceanogr. 24(5): 1085–1094.
6. Liu, P. C. (2000). Is the wind wave frequency spectrum outdated? Ocean Eng. 27:
577–588.
7. Massel, S. R. (2001). Wavelet analysis for processing of ocean surface wave records.
Ocean Eng. 28: 957–987.
8. Melville, W. K. (1983). Wave modulation and breakdown. J. Fluid Mech. 128:
489–506.
9. Bitner-Gregersen, E. M., and Gran S. (1983). Local properties of sea waves derived
from a wave record. Appl. Ocean Res. 5: 210–214.
10. Hwang, P. A., Xu, D., and Wu, J. (1989). Breaking of wind-generated waves: mea-
surements and characteristics. J. Fluid Mech. 202: 177–200.
11. Guillaume, D. W. (2002). A comparison of peak frequency-time plots produced with
Hilbert and wavelet transforms. Rev. Scientific Inst. 73(1): 98–101.
12. Hwang, P. A., Walsh, E. J., Krabill, W. B., Swift, R. N., Manizade, S. S., Scott, J. F.,
and Earle, M. D. (1998). Airborne remote sensing applications to coastal wave
research. J. Geophys. Res. 103(C9): 18791–18800.
13. Kaihatu, J. M., and Kirby, J. T. (1995). Nonlinear transformation of waves in finite
water depth. Phys. Fluids 7: 1903–1914.
5 An Application of HHT
Method to Nearshore
Sea Waves
Albena Dimitrova Veltcheva
CONTENTS
ABSTRACT
Real sea waves are nonlinear by nature, and this nonlinearity increases considerably
in the shoreward direction. The presence of wave groups in the sea wave records shows
that the wave process is also not stationary, as it is usually considered in conventional
wave data analysis. In this work, the Hilbert-Huang transform (HHT) method for
nonlinear and nonstationary time series analysis is applied to wave data from the
nearshore area. The field data were collected at Hazaki Oceanographical Research
Station (HORS), Port and Airport Research Institute, Japan, and covered different sea
states. The importance of a proper choice of an analyzing technique for the correct
understanding of the examined phenomenon is discussed. The ability of EMD, a key
part of the HHT method, to extract into IMFs the different oscillation modes embedded
in the sea surface elevation records is demonstrated. The variation of IMFs along the
beach profile is examined in an attempt to investigate cross-shore transformation of
sea waves. The dominant oscillations for each data record are determined, and their
variations during different sea stages are investigated.
97
5.1 INTRODUCTION
Real sea waves are nonlinear and nonstationary by nature, and their profile is
asymmetrical with respect to the zero level, with sharp crests and flat troughs,
especially well pronounced in shallow water. The existence of small ripples riding
longer waves, as well as an appearance of many small waves due to wave breaking,
complicates the wave profile considerably. Real sea waves also appear in groups,
which are time localized transient events. The wave data records always register a
mixture of many wave systems with different periods and energy, generated by
different sources. Through analysis of sea surface elevation record, we seek to
determine characteristic time–frequency scales of the energy. The widely used meth-
ods of wave analysis are burdened with the traditional concept of sea waves: as a
sine wave with definition domain from minus infinity to plus infinity. It is also
assumed, according to linear wave theory, that within the framework of one wave
the period is constant and the wave height is equal to twice the wave amplitude. In
addition to the positive maxima and negative minima of the wave profile, positive
minima and negative maxima are also often observed in the profile of real sea waves.
The neglect of these actual features of sea waves leads to incompleteness and
misinterpretation of the recorded sea state, and thus the analyzing method can distort
the investigated phenomenon.
The dominant approach in the existing methods of wave analysis is decompo-
sition of the time series into component basis functions. Global sinusoidal compo-
nents of fixed amplitude are the basis for data expansion in the commonly used
Fourier spectral analysis methods. The parameters of water waves, determined by
Fourier spectra, are integral characteristics of the sea state. Despite its being wide-
spread and useful, the Fourier spectral technique has an important drawback: the
Fourier spectrum contains no information about the timing of the analyzed process.
Consequently, the local time variations of determined wave characteristics are impos-
sible to track.
To understand the sea wave process correctly, we need an adequate method for
analysis. In the last decade the attention of scientists started to turn from globally
estimated wave parameters to local properties of a wave system. The attention has
been addressed to the time distribution of the wave energy, complimentary to the
frequency distribution information provided by usual Fourier spectral analysis. Liu
(1) suggested that the wave frequency spectrum approach is outdated, presenting
the advantages of the wavelet spectrum in the investigation of wave growth. The
differences between the Fourier and the wavelet spectrum are widely discussed by
Massel (2) on the basis of the application of the wavelet transform for the processing
of sea wave data. Huang et al. (3) also point out the necessity to break down the
earlier paradigm of wave analysis, emphasizing that Fourier analysis is not a good
method for studying nonlinear and nonstationary water waves.
Several Fourier-based methods for frequency–time distribution of the energy are
reviewed by Huang et al. (4): time window-width Fourier transform, evolutionary
spectrum, principal component analysis, and wavelet analysis. These methods suffer
from the demerits of Fourier spectral analysis — global definition of harmonic
components and the use of linear superposition for decomposition of the data
(assuming the stationarity and linearity of data). There is also a resolution problem
inherent in all Fourier-based analysis due to the Heisenberg Uncertainty Principle.
Wavelet analysis provides some improvement in the resolution by using adjustable
windows, but the problems of the nonadaptive nature of these windows are still not
solved completely. It has to be mentioned in addition that in these methods for
frequency–time distribution of the energy, the basis for expansion of the data must
be chosen in advance. An exception is the principal component analysis, in which
the basis is derived empirically from the data.
In this paper, we discuss the importance of the analyzing technique for the correct
understanding of the examined phenomenon and pay special attention to the non-
linear and nonstationary behavior of real sea waves. We propose a new method for
nonlinear and nonstationary time series analysis, the Hilbert-Huang transform (HHT)
method, introduced by Huang et al. (4), as an alternative one for investigation of
real sea waves. The empirical mode decomposition (EMD), a key part of the HHT
method, extracts the energy associated with various intrinsic time scales into a set
of intrinsic mode functions (IMFs). By virtue of their derivation, the IMFs have
well-behaved Hilbert transform results, and the instantaneous frequencies can be
calculated. The frequency–time distribution of energy is designated as a Hilbert
spectrum. Any event in the data series can be localized in time as well as in frequency.
We apply the HHT to the wave data from the nearshore area and demonstrate
the ability of EMD to extract into IMFs the different oscillation modes embedded
in the sea surface elevation records. The variation of IMFs along the beach profile
is examined in an attempt to investigate cross-shore transformation of sea waves.
The dominant oscillations for each data record are determined, and their variations
during different sea stages are investigated.
FIGURE 5.1 Location of wave gauges during the observations (HORS, PARI, Japan).
TABLE 5.1
Location of Wave Gauges
Number of Wave Gauge 1 2 3 4 5 6 7 8
Offshore distance (m) 3000 2000 1000 380 320 215 145 80
Water depth (m) 24.00 14.00 9.00 5.65 5.05 2.33 1.81 0.30
the wave periods became longer than 9 sec. At the end of the observation, noted
here as the post-storm stage, the wind changed its direction and the registered waves
were small in amplitude, but with long periods. The wave gauges 7, 8, and 9 were
located inside the surf zone during the calm sea stage; in other sea stages, the surf
zone became wider and also included wave gauges 4, 5, and 6.
FIGURE 5.2 Offshore characteristics of sea waves during the observation period.
2
Sea surface elevatiom (m)
-1
-2
80 90 100 110 120 130 140
Time (s)
TABLE 5.2
Discretization Criteria
Mean Wave
Criterion Definition of Individual Waves Frequency
often used for determination of individual waves from the sea surface elevation
records are briefly described in Table 5.2. The estimation of mean frequency depends
on the discretization criterion and is determined as ω ij = 2π j −i m j mi , where mk is
a kth spectral moment
∞
mk =
∫ f S( f )df ,
0
k
corresponds to the vertical displacement of sea surface elevation X(t). Here X̂ (t) is
the Hilbert transform of the sea surface elevation,
()
X t′ ( )
∫
1
X̂ t = P dt ′ , (5.2)
π t − t′
where P indicates the Cauchy principal values. The Hilbert transform is a convolution
of the sea surface elevation X(t) with 1/t and thus emphasizes the local properties of X(t).
() ()
2 2
A(t ) = X t + Xˆ t , (5.3)
Xˆ (t )
ϕ(t ) = arctg , (5.4)
X (t )
and
dϕ
ω (t ) = , (5.5)
dt
respectively. The envelope A(t), defined by Equation 5.3, provides a direct measure
of the wave amplitude and is widely used in the statistical analysis of the sea waves
(13, 14, 15, 16, 17), while the phase information is usually not considered. Being
an important physical quantity, the phase φ(t) associated with the sea wave process
can also bear fundamental information about the dynamics of sea waves. But unfor-
tunately, the phase is totally disregarded when conventional methods for wave
analysis are employed to study the wave system.
The phase φ(t) is a wrapped phase φ(t) ∈ [– π, π] or an unwrapped phase (t) ∈
[–∞, ∞], depending on its interval of definition. An example of the unwrapped and
wrapped phase of sea surface elevation X(t) is shown in Figure 5.4. The unwrapped
phase in Figure 5.4a increases smoothly with time, in contrast to the wrapped phase,
which varies widely in the range [– π, π].
The frequency ω(t), defined by Equation 5.5 as the rate of phase change, is
called the instantaneous frequency, and its value changes with time. This definition
of frequency was adopted before by Huang et al. (18) and used by Cherneva and
Veltcheva (19) for investigation of the local properties of sea waves and their group
structure. Later, Huang et al. (4) emphasized the considerable controversy in this
definition of instantaneous frequency, even with utilization of the Hilbert transform.
The instantaneous frequency is, as mentioned by Cohen (20), “one of the most
intuitive concepts, since we are surrounded by light of changing color, by sounds
of varying pitch, and by many other phenomena whose periodicity changes. The
exact mathematical description and understanding of the concept of changing fre-
quency is far from obvious.…” Blindly applying the definition Equation 5.5 of
instantaneous frequency to any analytic function may lead to one of the five para-
doxes listed by Cohen (20). One of these paradoxes concerns the appearance of
negative frequency: “Although the spectrum of the analytic signal is zero for negative
frequencies, the instantaneous frequency may be negative.” The enlarged portion of
the unwrapped phase shown in the upper left-hand corner of Figure 5.4a reveals the
jumps up and down in the time variation of the unwrapped phase. The unwrapped
phase function ϕ(t) is not always increasing function of time, and consequently the
instantaneous frequency ω(t) will have negative values, which have no physical
meaning.
1000
100
70
60
600
Unwrapped phase [rad]
50
80 100 120 140
Time [s]
400
200
0
0 200 400 600 800 1000 1200
Time [s]
4
Wrapped phase [rad]
-2
-4
0 200 400 600 800 1000 1200
Time [s]
FIGURE 5.4 The phase φ(t) of the analytical process. (a) Unwrapped phase; (b) wrapped
phase.
The negative advances of the phase ϕ(t) produce loops or phase reversals in a
polar diagram of the wave process (Figure 5.5), where the analytical function ξ(t)
is presented as a radius-vector rotating in the complex plane. The magnitude of the
vector is equal to the wave envelope A(t), estimated by Huang et al. (3), and the
angle of rotation is equal to the unwrapped phase function ϕ(t). The arrows show
the direction of rotation of the radius-vector. The rotation of the radius-vector with
positive instantaneous frequency is called a proper rotation by Yalcinkaya and Lai
(21) in their investigation of the phase dynamics of continuous chaotic flow.
1.5
0
-1.5 0 1.5
-1 .5
For a clearer illustration, Figure 5.5 shows the polar diagram of the part of the
wave record (thick, solid line) of Figure 5.3 between 85 sec and 112 sec. The
trajectory of ξ(t) in the complex plane generally may have multiple centers of
rotation. The majority of the individual zero-crossing waves in Figure 5.3 correspond
to a completely closed trajectory of rotation around the origin of the coordinate
system in Figure 5.5. The small ZC waves, around 90 sec and between 105 sec and
110 sec in Figure 5.3, are represented by loops in Figure 5.5. The centers of rotation
of these loops differ from the origin of the coordinate system. The small riding
waves in the wave profile also produce loops in the polar diagram, like the riding
wave between 91 sec and 95 sec in Figure 5.3. These phase reversals, as seen in
Equation 5.5, produce negative frequencies, which have no physical meaning.
According Gimenez et al.’s (12) orbital criterion, any discrete waves that do not
correspond to a 2π advance of the phase in the complex plane are considered as
false waves and are simply removed. In this way, the wave process is again assumed
to be a linear one, since only the completely closed orbit of water particles with 2π
advance of phase angle is accepted as an individual wave.
The small zero-crossing waves and ripples are indeed distinctive features of real
sea waves from the surf zone, and thus it is incorrect to eliminate them by the orbital
criterion for the purpose of achieving the positive frequency. Band-pass filtering,
proposed by Melville (22), also does not resolve the problem with negative frequen-
cies. Huang et al. (4) especially stressed the fact that the filtering operation itself
contaminated the data with spurious harmonics. These unsuccessful attempts to
resolve the problem of negative frequencies are probably attributable to an inappro-
priately chosen tool for investigation of nonlinear and nonstationary sea waves.
()
X t = ∑ C (t ) + R (t ) .
j =1
j n (5.6)
By definition, the IMF has to satisfy two conditions. First, the number of extrema
must be equal or differ at most by one from the number of zero-crossings; this is
similar to the traditional narrow band requirements for the stationary Gaussian
process. Practically, this condition corresponds to elimination of riding waves or
small waves in the time series with multiple centers of rotation. The derived IMF
corresponds to a proper rotation in the complex plane, as the center of rotation is
the origin of the coordinate system. Second, the IMF has to have symmetric envelopes,
defined by the local maxima and minima respectively. These upper and lower envelopes
are determined by using cubic splines, as suggested by Huang et al. (4). This second
condition ensures that the instantaneous frequency will not have fluctuations arising
from an asymmetric wave profile. The important condition for the IMF is that only
one maximum or minimum exists between successive zeros. Each IMF is determined
by a sifting procedure, which is repeated several times in order to ensure that all
the requirements of IMFs are satisfied.
The set of IMFs obtained in this way is unique and specific to the particular
time series, since it is based on and derived from the local characteristics of these
data. IMFs could be considered as a more general case of the simple harmonic
functions, but IMFs are claimed to have a physical meaning in additional to a
mathematical one due to their specific derivation.
In the second step, the Hilbert transform is applied to these IMFs:
()
∞
C j t′ ( )dt′
∫
1
Ĉ j t = P (5.7)
π t − t′
−∞
where P indicates the Cauchy principal value. The amplitude aj, the phase ϕj, and
the instantaneous frequency ωj are calculated by
()
a j (t ) = C 2j t + Cˆ 2j t , () (5.8)
Cˆ (t )
ϕ j (t ) = arctg , (5.9)
C (t )
and
ω j (t ) =
dϕ j t ( ). (5.10)
dt
By using Equation 5.8, Equation 5.9, and Equation 5.10, we can express the original
data X(t) as a real part (Re) of the complex expansion
()
X t = Re ∑ a (t ) e ∫
j =1
j
()
i ω j t dt
, (5.11)
which is considered a generalized form of the Fourier expansion. Here, both ampli-
tude aj(t) and instantaneous frequency ωj(t) are functions of time t. Whereas the
Fourier expansion is made in a global sense, the expansion by HHT is made in a
local sense.
It must be mentioned that, by definition, IMFs always have positive frequencies,
because the oscillations in IMFs are symmetric with respect to the local mean. Thus,
the HHT method solves the problem of negative frequencies, and there is no need
for any additional elimination of small waves, which are indeed observed peculiar-
ities of real sea waves. All the distinguishing features of sea surface elevation are
taken into account, and the subjectivity in estimation of wave characteristics probably
will be reduced. This is a direct result of the use of a more appropriate tool for
examining nonlinear and nonstationary real sea waves.
The frequency–time distribution of the amplitude or squared amplitude is des-
ignated as a Hilbert amplitude spectrum or Hilbert energy spectrum. In the present
work, we use the Hilbert energy spectrum, determined as the frequency–time dis-
tribution of the squared amplitude. For simplicity, the Hilbert energy spectrum is
denoted as the Hilbert spectrum H(ω,t). The frequency–time distribution of the
energy allows us to determine which frequency exists at a particular time, whereas
the Fourier frequency spectrum provides information as to which frequency exists
generally in given data series.
The time resolution of the Hilbert spectrum can be as precise as the sampling
data rate. Since the Hilbert spectrum gives the best fit of a local sine or cosine
function to the data, the frequency is uniformly defined by a local derivative of the
phase function. The lowest extractable frequency is 1/T, where T is the duration of
the record, and the highest frequency is 1/(l∆t), where l = 5 is the minimum number
of points necessary to define frequency accurately, and ∆t is the sampling rate.
Based on the Hilbert spectrum, the marginal Hilbert spectrum h(ω) can be
defined as
h(ω ) =
∫ H (ω, t)dt .
0
(5.12)
The marginal Hilbert spectrum represents the cumulated squared amplitude over
the entire data span and offers a measure of total energy contribution from each
frequency.
3 0.2
X(t) (a) (f )
C5
0 0
−3 −0.2
0 1000 2000 3000 4000 5000 6000 7000 0 1000 2000 3000 4000 5000 6000 7000
3 0.1
(b) (g)
C6
C1
−3 −0.1
0 1000 2000 3000 4000 5000 6000 7000 0 1000 2000 3000 4000 5000 6000 7000
3 0.1
(c) (h)
C2
0
C7
−3 −0.1
0 1000 2000 3000 4000 5000 6000 7000 0 1000 2000 3000 4000 5000 6000 7000
1.5 0.1
(d) (i)
C3
C8
0 0
−1.5 −0.1
0 1000 2000 3000 4000 5000 6000 7000 0 1000 2000 3000 4000 5000 6000 7000
0.5 0.2
(e) (j)
C4
0 0
R
−0.5 −0.2
0 1000 2000 3000 4000 5000 6000 7000 0 1000 2000 3000 4000 5000 6000 7000
Time (s) Time (s)
FIGURE 5.6 EMD of wave record. (a) Data of sea surface elevation; (b) through (i) eight
IMFs (C1 through C8); (j) residue R.
rotation in the complex plane of its own analytic function ψj(t) = aj(t) exp [iφj(t)]. The
average rotation frequencies ω j (t ) = d φ j (t ) dt obey the order ω1 ≥ ω 2 … ≥ ω n because
the sifting procedure picks the components with the fastest variations embedded in the
original data first and those with the slowest variations last. The mean periods of the
IMFs, shown in Figure 5.6, are correspondingly 3.4 sec, 8.4 sec, 16.6 sec, 35.3 sec,
89.2 sec, 235 sec, 572 sec, and 1540 sec.
These oscillations not only have different time scales, but they also have a
different range of energy. The limits of the vertical axes in Figure 5.6b through 5.6i
are different for different IMFs. The most energetic in the decomposition is the
second IMF, C2. The shortest oscillations in decomposition, presented by the first
IMF, C1, have smaller amplitudes than those of C2. Physically, these high frequency
oscillations, extracted in C1, may represent short period waves, such as wind waves
in the growing stage, as well as the small ripples that ride longer waves. The third
IMF, C3, takes second place in the energy hierarchy. The first three constituents in
the decomposition, C1, C2, and C3, probably represent the contribution of the wind
waves oscillations, while the lower frequency oscillations are characterized by the
IMFs with higher index: C4, C5, and so forth. The complete presentation of the sea
wave state recorded in this data can be considered as a composition of dominant
wave oscillations, extracted in the second IMF and several (but finite number)
components with smaller amplitudes.
The contribution of different Intrinsic Mode Functions to the energy and fre-
quency contents of wave data is investigated. The spectrum of wave record is
compared with spectra of its Intrinsic Mode Functions in Figure 5.7 as estimated
Fourier spectrum is shown in Figure 5.7a, while a marginal Hilbert spectrum h(ω)
is presented in Figure 5.7b. Under a sampling rate of these data, the highest frequency
that can be extracted by Empirical Mode Decomposition is 0.4 Hz. The highest
frequency of Fourier spectrum is 1 Hz, but for simplicity and easy comparison with
marginal spectrum, only until 0.35 Hz is shown in Figure 5.7a. The Fourier spectrum
is estimated as an average of the raw spectra, calculated in the overlapped segments,
by which the wave record is divided. The spectral estimations are also smoothed
with a Hamming window.
The Intrinsic Mode Functions represent different as period and energy oscilla-
tions, extracted by EMD from the wave data. The spectrum of the first IMF C1 (stars
line) covers very well the tail of wave spectrum (thick solid line) in the both ñ
Fourier and marginal spectrum. The peak of second IMF C2 (open circle line)
coincides with the major peak of the spectrum of sea surface elevation, both Fourier
and marginal spectrum.
The marginal Hilbert spectra are projections of real frequency-time distribution
of the energy, provided by Hilbert spectrum. There is basic difference in the meaning
of frequency in Fourier and Hilbert spectrum. The presence of energy at a given
frequency in Fourier spectrum means that a trigonometric component with this
frequency and amplitude exists through the whole time span of the data. In the
marginal Hilbert spectrum the existence of energy at a given frequency means that
in the whole time span of the data, there is a higher probability for local appearance
of such a wave.
C8
10–2
10–3
10–4
C7
C8
10–2
10–3
10–4
FIGURE 5.7 Spectra of wave data and its IMFs. (a) Fourier spectra; (b) marginal Hilbert
spectra.
∫ h (ω)dω .
C
m0 j = j (5.13)
TABLE 5.3
Characteristics of Offshore Sea Waves and Their IMFs
during Calm, Wave Growing, Decay, and Post-Storm Conditions
113
DK342X_book.fm Page 114 Sunday, May 1, 2005 12:34 PM
1.4
1.2
m0 m(m s )
1
C 2 −1
0.8
↑
m0(m s )
0.6
2 −1
Cm
0.4
0.2
0
0 0.2 0.4 0.6 0.8 1 1.2 1.4
2 -1
m0 (m s )
FIGURE 5.8 Comparison of energy of the dominant IMF (m0Cm) and wave data (m0).
0.8
24.0m (a)
0.7 14.0m
9.00m
0.6 5.65m
5.05m
2.33m
Energy ( m 2s-1)
0.5
1.81m
0.80m
0.4
0.3
0.2
0.1
0
0 1 2 3 4 5 6 7 8 9
Ci, i =1,8
0.3
24.0m (b)
14.0m
9.00m
5.65m
0.2 5.05m
Peak frequency(Hz)
2.33m
1.81m
0.80m
0.1
0
0 1 2 3 4 5 6 7 8 9
Ci, i =1,8
FIGURE 5.9 Decomposition of wave data measured in the cross-shore. (a) Energy of IMFs;
(b) frequency of IMFs.
considerably. At the same time, the oscillations extracted in C2 become shorter. The
energy and frequency of the lower frequency IMFs, C4, C5, C6, and C7, remain
approximately constant in the process of cross-shore transformation.
These peculiarities in the cross-shore variations of IMFs are reflected in the
frequency–time distribution of the energy. A section of the Hilbert spectrum of sea
waves in three different zones — offshore with water depth 24 m, in the seaward
surf zone at 5.05 m depth, and in the shoreward surf zone at 1.81 m — is shown in
Frequency (Hz)
0.25
H (m2)
10
10
5
0.2
0 8
0.4 0.15
0.35 1000 6
0.3 0.1
800
Fr 0.25 4
eq 0.2 600 0.05
ue
nc 0.15 400 s)
y( 0.1 e( 2
H m 0
z) 0.05 200 Ti 0 200 400 600 800 1000
0 0 0
Time (s)
0.4
(b) 5.05 m water depth
3 0.35
0.3
2.5
Frequency (Hz)
4 0.25
H (m2)
3 2
0.2
1
0 1.5 0.15
0.4 0.1
1000 1
0.3
Fr 800 0.05
eq 0.5
ue 0.2 600
nc 0
y( 400 ( s) 0 200 400 600 800 1000
H 0.1 me
z) 200 Ti 0 Time (s)
0 0
0.4
(c) 1.81 m water depth 1.5
0.35
0.3
1.5
Frequency (Hz)
1 0.25
H (m2)
1
0.5 0.2
0
0.15
0.4
1000 0.5
0.3 0.1
800
Fr
eq 600 0.05
ue 0.2
nc 400 s)
y(
Hz 0.1 e( 0
)
200 Tim 0 0 200 400 600 800 1000
0 0
Time (s)
FIGURE 5.10 A section of the Hilbert spectrum of sea waves, as a 3D plot in the left panels
and as a 9 × 9 Gaussian filtered color coded map in the right panels. (a) Offshore zone; (b)
seaward surf zone; (c) shoreward surf zone.
5.10a shows an energy concentration around 0.092 Hz, corresponding with the
second component, C2, which has the highest energy in the decomposition in Figure
5.6. The magnitude of energy decreases, as revealed by changes in the energy scale
of the color bars. A wide variation in the local frequency in the surf zone is evident
in Figure 10b and c, in contrast with the offshore region. This broadening of the
wave frequency range is partially due to wave breaking. The strong nonlinearity of
the sea waves in shallow water, shown by the intra-wave frequency modulation or
the variation of the local frequency within one wave, also contributes to a large
variation in the local frequency in the Hilbert spectrum.
On the basis of the results of EMD, real sea waves could be considered as a
system of oscillations with different periods and amplitudes. It has to be stressed
that the amplitudes and periods of these oscillations are not constant but instead are
functions of time according to Equation 5.11. The wave system, as a whole, changes
its content in cross-shore transformation. The process of wave breaking mainly
affects the IMFs with the highest energy in the decomposition of offshore waves.
Inside the surf zone, the frequency of each IMF tends to increase. These peculiarities
in the spatial distribution of different IMFs agree with general principles of sea wave
transformation in the cross-shore direction. The analysis of IMFs and the Hilbert
spectrum provides information on the cross-shore transformation of sea waves.
The EMD method, with its great ability to extract specific oscillation data
embedded in the data, can facilitate the investigation of sea waves. Attention can be
concentrated on a finite number of IMFs, each characterizing a dynamical system
that is much simpler than the original one. The application of the HHT method to
nearshore waves is a fertile area of research, as future studies will shed more light
on the complicated dynamics of the coastal zone.
5.6 CONCLUSIONS
We have discussed the importance of the analyzing technique for the correct under-
standing of the examined phenomenon and paid special attention to the nonlinear
and nonstationary behavior of real sea waves. A review of widely used conventional
methods for wave analysis shows that all of them assume a priori linearity and
stationarity of sea waves; these assumptions consequently affect the estimation of
wave characteristics.
The Hilbert-Huang transform method for the analysis of nonlinear and nonsta-
tionary time series is proposed as an alternative for the investigation of the nonlinear
and nonstationary nature of real sea waves. In this method, the oscillations embedded
in the data are extracted by EMD into a set of IMFs without placing any subjective
preliminary limitations on the nature of the investigated phenomenon. Easy separa-
tion of the different, as time and as energy associated with the oscillation in the data
is well achieved by EMD.
The dominant oscillations for each data record were determined and their vari-
ations during different sea stages were investigated. The energy hierarchies of the
IMFs in decomposition of wave records observed during different sea stages reflect
the specific peculiarities of the particular wave data. The process of wave breaking
mainly affects the IMFs with the highest energy in the decomposition of offshore
waves. Inside the surf zone, the frequency of each IMF tends to increase.
The EMD method, with its great ability to extract oscillation modes embedded
in the data, can facilitate the investigation of sea waves. Attention can be concentrated
on a finite number of IMFs, each characterizing a dynamical system that is much
simpler than the original one. The application of the HHT method to nearshore
waves is a fertile area of research, as future studies will shed more light on the
complicated dynamics of the coastal zone.
REFERENCES
1. Liu, P. C. (2000). Is the wind wave frequency spectrum outdated? Ocean Eng. 27:
577–588.
2. Massel, S. R. (2001). Wavelet analysis for processing of ocean surface wave records.
Ocean Eng. 28: 957–987.
3. Huang, N. E., Shen, Z., and Long, S. R. (1999). A new view of nonlinear water
waves: The Hilbert spectrum. Annu. Rev. Fluid Mech. 31: 417–457.
4. Huang, N. E., Shen, Z., Long, S..R., Wu, M. C., Shin, H. S., Zheng, Q., Yuen, Y.,
Tung, C. C., and Liu, H. H. (1998). The EMD and Hilbert spectrum for nonlinear
and non-stationary time series analysis. Proc. R. Soc. London. Vol. 454: 903–995.
5. Katoh, K., Nakamura, S., and Ikeda, N. (1991). Estimation of infragravity waves in
consideration of wave groups: An examination on basis of field observation at HORF.
Rep. Port Harbour Res. Inst. Vol. 30, 1: 137–163.
6. Ochi, M. K. (1998). Ocean Waves: The Stochastic Approach. Cambridge Ocean
Technology Series 6, 319 p. New York: Cambridge University Press.
7. Gimenez, M. H., Sanchez-Carratala, C. R., and Medina, J. R. (1998). Influence of
false waves on wave records statistics. Proc. ICCE 1998, pp. 920–933.
8. Kitano, T., Mase, H., and Nakano, S. (1998): Statistical properties of random wave
periods in shallow water: Analysis by utilizing false waves. Proc. Coast. Eng, JSCE,
Vol. 45, 221–225.
9. Veltcheva, A., and Nakamura, S. (2000). Wave groups and low frequency waves in
the coastal zone. Rep. Port Harbour Res. Inst. Vol. 39, 4: 75–94.
10. Mizuguchi, M. (1982). Individual wave analysis of irregular wave deformation in the
nearshore zone. Proc. ICCE 82, pp. 485–504.
11. Hamm, L., and Peronnard, C. (1997). Wave parameters in the nearshore: A clarifica-
tion. Coastal Eng. 32: 119–135.
12. Gimenez, M. H., Sanchez-Carratala, C. R., and Medina, J. R. (1994). Analysis of
false waves in numerical sea simulations. Ocean Eng. Vol. 21, 8: 751–764.
13. Longuet-Higgins, M. S. (1958). On the intervals between successive zeros of a
random function. Proc. R. Soc. Ser. A, 246: 99–118.
14. Tayfun, M. A. (1983). Frequency analysis of wave heights based on wave envelope.
J. Geophys. Res. Vol. 88, C12: 7573–7587.
15. Bitner-Gregersen, E. M., and Gran, S. (1983). Local properties of sea waves derived
from a wave record. Appl. Ocean Res. Vol. 5, 4: 210–214.
16. Hudspeth, R. T., and Medina, J. R. (1988). Wave group analysis by Hilbert transform.
Coastal Eng. Vol. 1: 885–898.
17. Tayfun, M. A., and Lo, J. M. (1989). Envelope, phase, and narrow-band models of
sea waves. J. Waterway Port Coastal Ocean Eng. ASCE, Vol. 115, 5: 594–613.
18. Huang, N. E., Long, S. R., Tung, C. C., Donelan, M. A., Yuan, Y., and Lai, R. J.
(1992). The local properties of ocean surface waves by the phase-time method.
Geophysical Res. Lett. Vol. 19, 7: 685–688.
19. Cherneva, Z., and Veltcheva, A. (1993). Wave group analysis based on phase prop-
erties. Proc. 1st Int. Conf. Mediterranean Coastal Environ. (MEDCOAST ’93), Tur-
key, pp. 1213–1220.
20. Cohen, L. (1995). Time-Frequency Analysis. Prentice Hall Signal Processing Series,
Alan V. Oppenheim, series editor, 299 p.
21. Yalcinkaya, T., and Lai, Ying-Cheng (1997). Phase characterization of chaos. Phys.
Review Lett. Vol. 79, 20: 3885–3888.
22. Melville, K. (1983). Wave modulation and breakdown. J. Fluid Mech. 128: 489–506.
23. Farge, M. (1992). Wavelet transforms and their applications to turbulence. Annu. Rev.
Fluid Mech. 24: 395–457.
24. Veltcheva, A. D. (2001). Wave groupiness in the nearshore by Hilbert spectrum. Proc.
4th Int. Symp. Ocean Wave Meas. Anal. (WAVES 2001), San Francisco, pp. 367–376.
25. Veltcheva, A. D. (2002). Wave and group transformation by a Hilbert spectrum.
Coastal Eng. J. Vol. 44, 4: 283–300.
26. Salisbury, J. I., and Wimbush, M. (2002). Using modern time series analysis tech-
niques to predict ENSO events from the SOI time series. Nonlinear Processes Geo-
physics, Vol. 9: 341–345.
6 Transient Signal
Detection Using the
Empirical Mode
Decomposition
Michael L. Larsen, Jeffrey Ridgway,
Cye H. Waldman, Michael Gabbay,
Rodney R. Buntzen, and Brad Battista
CONTENTS
ABSTRACT
121
6.1 INTRODUCTION
In this paper, we discuss application of nonlinear signal processing techniques to
underwater electromagnetic data. Electromagnetic measurements provide a poten-
tially useful complement to acoustic detection methods in littoral zones where high
noise levels and reverberation in the littoral limit acoustic detection ranges. In
shallow waters, for example, the ambient noise levels from wind-generated breaking
waves are typically 10 to 100 times greater than those in deep water. Electromagnetic
measurements are affected by noise as well. One source of electromagnetic fields
in seawater is electrical currents induced by hydrodynamic flow in the Earth’s
magnetic field. The ocean is a constantly moving conducting fluid [26], and ocean
flows caused by gravity waves (wind waves and far-field swell), internal waves,
turbulence, tides, and currents all result in electromagnetic field fluctuations.
While the standard way of viewing the complexity of wave phenomena is to
consider the ocean to be a random, Gaussian process, the importance of nonlinear
interactions in water wave dynamics has been conclusively demonstrated [28]. Fur-
thermore, turbulence is also often present at the bottom boundary layer due to
nonlinear interactions between flows arising from tidal and wind-driven currents,
wind waves, swell, and internal waves [4]. In the random sea assumption (used
primarily in the deep ocean), the evolution of the sea surface is obtained as a
superposition of waves of different frequencies and directions with random phases.
The wave amplitudes are determined according to a measured power spectrum, the
sea surface is modeled as a linear system, and energy transfer between wave modes
is not accounted for. In littoral regimes, however, nonlinear wave interactions are
significant and cannot be neglected. Higher-order statistics show that wave behavior
in shallow water zones displays a distinctly non-Gaussian regime, where wave–wave
interaction plays an important role and energy transfer between wave components
occurs [18, 12].
A typical power spectrum of naturally occurring ocean-bottom magnetic and
electric field data is shown in Figure 6.1. The most obvious peaks in both electric
and magnetic spectra in Figure 6.1 are visible at 62 MHz, the dominant oceanic
swell frequency. Swells have characteristic frequencies in the 50 to 100 MHz range,
and surface waves driven by local winds are found at frequencies above the swell
frequency, up to about 0.5 Hz. The effect of ocean swell on the magnetic field is
significantly more pronounced than its effect on the electric field. An additional peak
is also seen in the spectrum of the magnetic field data at twice the swell frequency
(125 MHz). Frequency doubling effects due to the nonlinear interference of the swell
with wave fields reflecting off local land masses have been previously observed [27,
24]. Spectral peaks above these frequencies are likely due to local turbulence, with
the exception of the Schumann resonance discussed below.
Another major component of underwater electromagnetic noise is actually due
to sources of non-oceanic origin — geomagnetic, atmospheric, and ionospheric
electromagnetic radiation. The Schumann resonance at 8 Hz is caused by worldwide
lightning, which resonates in the earth-ionosphere cavity. Geomagnetic noise at
lower frequencies originates in the ionosphere, is nonstationary, and increases in
Electric field spectrum, with and Magnetic field spectrum, with and
without ionospheric mitigation without ionospheric mitigation
3 3
10 10
ionosphere unmitigated ionosphere unmitigated
102 ocean swell 102 ocean swell
101 101
Power (uV/m2/Hz)
0
100
Power (nT2/Hz)
10
−1
instr. artifact double freq.
10 10−1
10−2 ionosphere mitigated 10−2 ionosphere instrument artifacts.
mitigated
10−3 10−3
−4
Schmn res. −4
10 10
−5
10 (a) E-field 10−5 (b) B-field
10−6 10−6
10−3 10−2 10−1 100 101 10−3 10−2 10−1 100 101
Frequency (Hz) Frequency (Hz)
FIGURE 6.1 Power spectra for ocean-bottom electromagnetic fields before and after noise
mitigation.
power at lower frequencies, imparting the strong slope to the spectrum seen in Figure
6.1. These effects were observed in our experiments and in nearshore tests done by
other researchers [16]. Geomagnetic noise is typically coherent over large distances
and can be mitigated by using a remote reference station [17]. A transfer function
can be calculated to account for the difference in the conductivity of the underlying
geology and used to cancel out the effect of magnetospheric sources observed in
the local sensor data. Figure 6.1 shows the spectrum before and after such noise
reduction was performed.
Characterizing the nonstationary and nonlinear character of ocean-bottom elec-
tromagnetic noise is important for detection of manmade electromagnetic (EM)
anomalies. Ships and submarines generate electric fields in the surrounding water
due to corrosion currents caused by the dissimilar metals used in their construction
and also by onboard cathodic protection systems designed to mitigate such currents.
The bronze propeller acts as one node of the vessel’s galvanic cell, while the
sacrificial zincs on the hull act as nodes of opposite polarity. A current flows from
the propeller through the drive shaft, bearings, and hull into the water, returning to
the propeller. This current flow in the surrounding water can be well modeled as an
electric dipole. The horizontal motion of the ship or submarine imparts characteristic
ultra-low frequency signals to the field components, which change with the speed
and range of the submarine. This signature is described as the “quasi-static” horizontal
electric dipole (HED) signature. Impressed on the quasi-static dipole field is an AC
modulation, which is caused by variations in the electrical resistance in the shaft as it
rotates. A vessel passing a stationary electric field sensor generates a transient signal
in the nonstationary background ambient electric field. Figure 6.2 shows an example
of simulated horizontal electric field (E-field) components (Ex and Ey).
For a stationary sensor, the measured dipole signal is a function of the vessel
speed and its lateral distance from the sensor at the point of closest approach. The
spectra of moving dipole signatures are generally confined to ultra-low frequencies
(0.5 to 5 MHz) and thus fall into the underwater electromagnetic frequency regime
1
Electric Field, E (µV/m)
ELFE Detail
0.5
−0.5
−1
0 5 10 15 20 25 30 35 40 45
Time (mins)
FIGURE 6.2 Simulated time series of the two horizontal components of the electric field
from a moving vessel. The dipole appears as a slow transient. The inset is an enlarged view
of the cyclic modulation of the dipole.
∞
x (τ)
∫
1
y(t ) = PV d τ, (6.1)
π −∞ t−τ
where PV denotes the Cauchy principal value. Hence, the Hilbert transform is the
convolution of x(t) with 1/t, which consequently stresses the local nature of the
signal. From the complex analytic signal formed via z(t) = x(t) + iy(t), an instanta-
neous amplitude and phase can be calculated. The instantaneous frequency corre-
sponds to the time derivative of the phase of z(t).
Direct application of the Hilbert transform to data may yield negative frequen-
cies. Only positive instantaneous frequencies, however, correspond to physically
meaningful oscillatory behavior. In order to avoid nonphysical instantaneous fre-
quencies, the empirical mode decomposition is used to separate the data into
well-behaved intrinsic mode functions (IMFs) from which the Hilbert transform will
yield positive instantaneous frequencies. An IMF is a function satisfying two con-
ditions: i) the number of extrema and the number of zero-crossings of the function
must either be equal or differ at most by one, and ii) at any instant in time, the mean
value of the envelopes defined by the function’s local maxima and minima must be
zero. Each IMF has a characteristic frequency, although IMFs can overlap in fre-
quency content. IMFs are found by using a recursive sifting procedure that generates
the highest-frequency IMF first. This IMF is subtracted from the time series, and
the process is iteratively applied to the difference until only a non-oscillatory resid-
ual, r, which represents the trend in the data, remains. Each IMF xn(t) has a variable
instantaneous amplitude, an(t), and frequency, ωn(t), calculated via the Hilbert trans-
form. The time–frequency distribution H(ωn,t) of the amplitude is known as the Hilbert
spectrum. The original time series can be written as a sum of a finite number of IMFs:
N
x (t ) = Re ∑
n=1
an (t )eiω n (t ) dt + r .
(6.2)
∫
T
HMS (ω ) = H (ω, t )dt,
0
which represents the accumulated energy at each frequency over the entire data span
and is related to the fraction of time that a given frequency can be observed in the
system. The marginal spectrum is analogous to the Fourier power spectral density.
The EMD leads to compact representation of nonlinear or nonstationary signals
without higher-order harmonics or a proliferation of global modes needed to account
for rapid transients. It is often observed that individual IMFs often correspond to
identifiable physical processes in the data [13,14].
Recent work is yielding greater understanding of the behavior of the EMD and
its potential for signal processing applications [8,7]. It appears to operate on frac-
tional Gaussian noise as an almost dyadic filterbank, resembling a wavelet decom-
position [10,9]. Furthermore, numerical experiments indicate that its impulse
response is similar to a cubic spline wavelet [25].
−0.5
0.02
c1
0
−0.02
0.02
c2
0
−0.02
0.5
c3
0
−0.5
0.2
c4
0
−0.2
0.2
0
c5
−0.2
0 100 200 300 400 500
Time
FIGURE 6.3 Simulated, combined dipole and cyclic signals at different ranges and their
derived IMF components. The lower five graphs display the individual IMFs, showing the
ability of the EMD procedure to separate signals with different characteristics.
Similar results are obtained for noisy signals. A typical track time series from
the experiment, containing both the HED and cyclic signal generated by a towed
source, is shown in the top graph of Figure 6.4. The cyclic signal (in the second)
and the HED (in the seventh panel), highlighted in red, are easily identified in this
close-pass track, which still contains noise.
We investigated the performance of the matched filter for detecting a moving
HED in the ocean using synthetic marine signals, which simulate a single dipole
moving in a straight path past a single receiver. The expected signal is a function
of the dipole amplitude and target range, bearing, and speed. In this study we elected
to simplify the situation by assuming a priori knowledge of speed and range, in
order to focus on the comparison between the matched filter and the empirical mode
(EMD)-assisted matched filter (EMMF) to be described later. In practice, one would
sweep through the set of realizations of speed, range, and HED dipole strength, and
determine the parameters that maximize the matched filter output. The two compo-
nents of the electric field were processed together in quadrature as a complex signal,
with Ex as the real component and Ey as the imaginary component. A correlation
filter was used with the magnitude of the output being the detection criterion,
normalized to a maximum unity output. Through simulation it was determined that
x (t)
0
−2
0.2
x1–2
0
−0.2
0.1
0
x3
−0.1
0.2
x4
0
−0.2
0.5
x5
0
−0.5
µ
0.5
x6
0
−0.5
2
x7–10
0
−2
0.5
x11
0
−0.5
0.1
x11
0
−0.1
0.1
0
r
−0.1
0 500 1000 1500 2000 2500 3000
Time (sec.)
Known target
s(t)
signature
FIGURE 6.5 Block diagram of EMD-assisted matched filter, where the matched filter process
is preceded by a least-squares fit of the IMFs (generated from the time series data).
the expected signal magnitude depends on the CPA range and target speed but not
the target bearing. All bearings will have the same magnitude response, with differing
phase responses. Thus the matched filter output depends only upon using the correct
range and speed of the vessel.
Building on the matched filter, an improved algorithm, the empirical mode
matched filter, was developed to exploit the EMD’s ability to capture the signal in
just a few of the IMFs. This method applies the matched filter to a signal generated
using the IMFs as basis functions, rather than to the target signal x(t) itself. This
improved detector performance. A block diagram of the processing steps of the
algorithm is shown in Figure 6.5. Empirical mode decomposition was first applied
to the time series to generate the IMFs αi(t), i = 1, …, N, where N is the number
of IMFs. In the next step, the linear combination of the IMFs containing the target
signal s(t) which best approximates the known target signature in the least-square
sense was determined:
min || s(t ) −
ci ≥ 0 ∑ c α || .
i =1
i i
The least-squares optimization was constrained so that the weights ci are positive.
A matched filter based on the target signature s(t) was then applied to the signal
x̂ (t), which is composed of the weighted linear combination of the IMFs,
xˆ (t ) = ∑ cα ,
i =1
i i
1.5
raw signal
LSQ wt signal
model
0.5
0
µV/m
0.5
1.5
2.5
0 100 200 300 400 500 600 700
time (sec)
3000 overlapping samples in the simulation. The window size was set to τ = 2.7
(τ = V/RCPA), which from experience yields an effective template for the matched
filter. The desired Ex and Ey signals are shown in Figure 6.7.
A τ parameter of 2.7 yields signals that are 9 min long for this geometry. Longer
τ values include the tails of the signals, which contain little information, and shorter
τ values may cut out valuable information about a signal’s shape that makes it
different from the background. For this τ and constant source speed V, the signals
lengthen in time proportional to their CPA offset RCPA.
Receiver operating curves (ROC) curves were generated for the white Gaussian
noise and signal by calculating distributions of the matched filter output (normalized
to 0–1) for inputs of noise only and the noise plus signal. Once the distributions
were created, ROC curves were extracted, utilizing a sliding threshold, to determine
the probability of detection (PD) at different probabilities of false alarm (PFAs). The
result, shown in Figure 6.8, confirms the expected theoretical behavior: that the
standard LMF technique yields the best detection statistics for this type of noise.
The LMF procedure outperforms the EMMF (also labeled as the “LSQ” in the
figures that follow because the method involves a least-squares fit of the IMFs to
the desired matched filter). As the source strength and signal-to-noise ratio decrease,
both the LMF and EMMF ROC performances deteriorate, but the LMF maintains
a slight advantage, as predicted by theory. Signal-to-noise ratio (SNR) is defined
0.5
E-field (uV/m)
−0.5
−1
−1.5
−2
0 100 200 300 400 500 600
Time (seconds)
FIGURE 6.7 Ex (solid) and Ey (dotted) signals for a north–south track, where τ = 2.7, which
yields a 9-min signal length at this CPA and speed.
1 1
0.9 0.9
0.8 0.8
0.7 0.7
0.6 0.6
PD
PD
0.5 0.5
0.4 0.4
0.3 0.3
0.2 0.2
0.1 LSQ 0.1 LSQ
LMF LMF
0 0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
PFA PFA
FIGURE 6.8 Comparative ROC curves for HED moving dipole detection in Gaussian white
noise, for LMF method (lower curve) and EMMF/LSQ method (upper curve). The original
source strength has been reduced to correspond to SNR values of –14 and –16.4. The LMF
method outperforms the EMMF in Gaussian white noise, in accordance with signal-processing
theory.
here as the peak excursion of the target signal versus the standard deviation of the
background noise [11], normalized to a logarithmic dB number:
SNR = 20 log10 ( P / σ ) ,
where P is the peak excursion of the target signal (of a specified component), and
σ is the standard deviation of the background noise. This broadband definition is
SNR 10
0.9
0.8 SNR 7
0.7
SNR 5
0.6
PD (01)
0.5
0.4
0.3
0.2
0.1 EMMF/LSQ
LMF
0
0 0.2 0.4 0.6 0.8 1
PFA (01)
FIGURE 6.9 ROC curves for LMF process (light) and EMMF/LSQ process (dark), for
several different SNRs. When the SNR is strong, the two processes achieve nearly identical
detection results, but when the SNR is weaker, the EMMF process has a definite advantage.
calculated independently of frequency. The SNR for the ROC curves in Figure 6.8
is negative (printed on curves) and means that the signal amplitude is lower than
the standard deviation of the noise. A SNR of –14, for example, corresponds to the
peak signal value being one-fifth the noise standard deviation.
The advantage of the LMF over the EMMF method is lost when an actual noise
background is used. The background used in the following analysis consists of 52
h of experimentally measured underwater electromagnetic data, which is non-Gaus-
sian with a colored spectrum and a nonstationary variance. In actual oceanic and
geomagnetic noise, the EMMF method is superior. This is demonstrated in Figure
6.9, where we have displayed three representative curves together on one graph,
utilizing SNRs between –5 and +14, much higher than the SNRs used in the white
Gaussian noise process. A major characteristic of the real-world noise is its nonsta-
tionarity and high power at low frequencies, which can lead to anomalies that mimic
the transient HED.
The ROC performance curves in Figure 6.9 have been enhanced by two sig-
nal-conditioning procedures: high-pass filtering of the data to produce first-order
stationarity (see above), and removal of a linear trend within each window being
examined. The window length is determined by the τ value being sought. (For the
0.6
PD (01)
0.5
0.4
0.3
0.2
0.1
0
0 10 20 30 40 50
Source Strength (dB Am)
FIGURE 6.10 Relative gain for HED detection using EMMF at set PFA (= 0.1) for two
different ranges. PD is plotted as a function of source strength in dB. The EMMF method is
in blue. For a given PD, the necessary source strength is lowered by 2 to 3 dB, for moderate PD.
EMMF
1 LMF
0.8
PD (01)
0.6
0.4
0.2
Range
FIGURE 6.11 Probability of detection versus CPA, for LMF (light) and EMMF (dark)
methods. As the range is further increased and SNR decreases, the EMMF method excels,
extending the detection range by 10 to 15% for a set PD, or increasing the PD by 10 to 25%
for a given range.
We repeated this procedure for a greater range for unreduced data, which yields
a nearly identical curve, but shifted to the right about 15 dB. Thus the EMMF method
maintains its increase in effectiveness across range. The detection versus range
behavior is summarized in Figure 6.11, where we plot PD versus CPA range for a
set PFA = 0.1. Up to a certain range, the two methods again yield an identical high
probability of detection, but as the range is further increased (and SNR decreases),
the EMMF method (dark curve) is more effective, extending the detection range by
10 to 15% for a set PD, or increasing the PD by 0.1 to 0.2 for a given range. We
see that the EMMF method flattens the falloff curve for low SNR regimes, thereby
extending detection ranges. The inclusion of the EMD into the matched filter process
adds detection gain in high-amplitude, colored, and nonstationary noise such as
observed in the underwater electromagnetic environment. This effectiveness
enhancement is seen generally in low to medium SNRs, whether caused by weaker
source strength or a farther CPA.
We also investigated the effectiveness of the EMMF method in a noise regime
where the common-mode geomagnetic noise has been partially mitigated by using
remote-reference techniques. This yields improved noise reduction so that an algorithm
2 0.8
1.5 0.7
0.6
PD (01)
1
0.5
0.5
0.4
0
0.3
−0.5 0.2
−1 sigma = 0.090 0.1 EMMF/LSQ
LMF
−1.5 0
0 5 10 15 20 25 30 35 40 0 0.2 0.4 0.6 0.8 1
Time (hours) PFA (01)
FIGURE 6.12 Effect of remote-reference noise reduction on matched filter methods. (A)
Unreduced (top) and reduced (bottom) time series for background, X component, over 40 h.
The geomagnetic noise removal not only lowers the variance, but also makes the resulting
series more stationary. (B) ROC curves generated from reduced data, for EMMF (dark) and
LMF (light) methods. The LMF method is more effective at low ranges/high SNR, whereas
the EMMF method increasingly dominates at greater ranges.
comparison can be made in a background where oceanic and other processes dom-
inate, rather than ionospheric sources. Figure 6.12A shows a plot of the unreduced
and reduced time series. Visually, the reduced time series is more uniform across
its entire length and has both lower variance and fewer outliers than the original
time series, which is simply high-pass filtered. Histogram analysis at various posi-
tions in the time series also indicates that it has more a stationary variance, and
spectral analysis shows it to have a flatter power spectrum at low frequencies and
thus to be closer to white Gaussian noise.
We repeated the Monte Carlo procedure for generating ROC curves as a function
of range, with the results shown in Figure 6.12B. Detection ranges are noticeably
extended. Also, the LMF method is actually slightly superior to the EMMF, at ranges
where the SNR is high. However, as the range increases and the SNR decreases, the
EMMF method becomes more effective than the linear matched filter, providing a
flatter rolloff of PD versus range.
Figure 6.13 shows PD performance versus range for the EMMF and LMF
methods for reduced data. Here we have set the PFA to a constant (= 0.1), and
calculated the PD as a function of range for both methods. The linear method has
a slight advantage at lower ranges and higher SNR, whereas the HHT-enhanced
method is more effective at long ranges and a small SNR. Overall, the EMMF
method provides superior performance with flatter rolloff versus range.
0.9
0.8
0.7
0.6
PD (01)
0.5
0.4
0.3
0.2
0.1
Range
FIGURE 6.13 Summary graph of behavior of PD versus range for EMMF (dark) and LMF
(light), after coherent geomagnetic noise removal. The LMF method is slightly improved at
lower ranges, but overall the EMMF method provides a superior curve with flatter rolloff
versus range.
versus the LMF method of detection for the moving HED signal on unreduced data.
A total of 36 track/sensor combinations existed in this CPA range. For each individual
track, assuming knowledge of the track bearing and CPA, we performed the algo-
rithms discussed above and measured the correlation signal. Such signals generally
produced a peak very near the CPA. To determine whether a detection occurred, we
applied the matched filter to data measured at a remote sensor which contains only
background noise, and calculated a threshold for a given false alarm level (a PFA
level of 5% was used).
The results are displayed in Figure 6.14A for the LMF method and Figure 6.14B
for the EMMF method. As expected from the ROC analysis, the EMMF method is
superior. EMMF correlations are uniformly higher with a much flatter falloff versus
CPA than for the LMF method. Additionally, the threshold (actually a mean value
of the thresholds over all 36 cases) for the EMMF method raises only slightly, which
results in an overall increase in detectability using the EMMF. As displayed in Figure
6.15, in the 36 cases tested, the EMD-based method results in 24 detections, versus
only 14 for the linear matched filter (points very near to the detection threshold were
counted as detections), using a PFA of 5%.
Correlation (0–1)
0.7 threshold for PFA = 0.05 0.7 threshold for PFA = 0.05
0.6 0.6
0.5 0.5
0.4 0.4
0.3 A): 14/36 detections 0.3 A): 24/36 detections
0.2 0.2
0.1 0.1
FIGURE 6.14 Comparison of (A) the LMF method with (B) the EMMF method, utilizing
signals created by the towed electric source. The EMMF has a flatter falloff with range and
more detections (24/36 versus only 14 for the LMF).
Detections vs. range: LMF vs EMMF for data with coherent noise reduction
Possible chances
Number of detections, chances
Detections EMD
Detections LMF
Range
FIGURE 6.15 Comparison of number of detections of the LMF method (light) with the
EMMF method (medium), utilizing signals created by the towed electric source.
6.5 CONCLUSION
In summary, we have exploited the empirical mode decomposition to isolate physical
signals into a discrete number of intrinsic mode functions in order to enhance the
detection of transient HED signals in a background of geomagnetic and hydrody-
namic noise. We have numerically compared this method to the standard matched
filter by deriving ROC curves via a Monte Carlo simulation for experimental data.
In white Gaussian noise, the LMF method yields superior ROC curves, in accordance
ACKNOWLEDGMENT
This work was funded by the Office of Naval Research and others. The authors are
grateful for the experimental assistance of the Swedish Defense Ministry and the
FOI, and for the assistance of Dave Rees of SPAWARSYSCEN San Diego.
REFERENCES
1. J. Berthier, F. Robach, B. Flament, and R. Blanpain. Geomagnetic noise reduction
from sea surface, high sensitivity magnetic signals using horizontal and vertical
gradients. In Proc. 1998 Oceans Conf., July 2001.
2. R. Blampain. Traitement en temps réel du signal issu d'une sonde magnetometrique.
Doctoral thesis of engineering, INPG, Grenoble, 1979.
3. C. Chichereau, J.L. Lacoume, R. Blanpain, and G. Dassot. Short delay detection of
a transient in additive Gaussian noise via higher order statistic test. In IEEE Digital
Signal Proc. Workshop, pp. 299–302, 1996.
4. C.S. Cox, X. Zhang, S.C. Webb, and D.C. Jacobs. Electro-magnetic fluctuations
induced by geomagnetic induction in turbulent flow of seawater in a tidal channel.
In MARELEC, 3rd Int. Conf. Marine Electromagnetics, July 2001.
5. B. Deniel. Undersea magnetic noise reduction. In Proc. 1997 MTS/IEEE Conf., vol.
2, pp. 784–788, 1997.
6. R. Donati. Detection of the electric field due to corrosion phenomena. In Proc. 2nd
Int. Conf. Marine Electromagnetics (MARELEC), Vol. 2, pp. 322–325, 1999.
7. P. Flandrin, P. Gonclaves, and G. Rilling. Detrending and denoising with empirical
mode decompositions. In Proc. IEEE-EURASIP Workshop Nonlinear Signal Image
Process. NSIP-04, to appear in 2004, available at http://www.inrialpes.fr/is2/peo-
ple/pgoncalv/pub/emd-eusipco04.pdf.
8. P. Flandrin and P. Gonclaves. Sur la decomposition modale empirique. In Proc.
GRETSI-03, 2003.
9. P. Flandrin, G. Rilling, and P. Gonclaves. Empirical mode decomposition as a filter
bank. IEEE Sig. Proc. Lett., 11(2):112–114, 2004.
10. P. Flandrin. Empirical mode decompositions as data-driven wavelet-like expansions
for stochastic processes. In Proc. Conf. Wavelets Statistics, 2003.
11. B.C. Fowler, H.W. Smith, and F.W. Bostick Jr. Magnetic anomaly detection utilizing
component differencing techniques. Technical Report 154, Electrical Geophysics
Research Lab., Univerisity of Austin Texas, 1973.
12. K. Hasselmann. On the nonlinear energy transfer in a gravity wave spectrum. Jour.
Fluid Mech., 12:481–500, 1962.
13. N.E. Huang, Z. Shen, S.R. Long, M.L. Wu, H.H. Shih, Q. Zheng, N. Yen, C.C. Tung,
and H.H. Liu. The empirical mode decomposition and Hilbert spectrum for nonlinear
and non-stationary time series analysis. Proc. R. Soc. London A, 454: 903–995, 1998.
14. N.E. Huang, Z. Shen, and S.R. Long. A new view of nonlinear water waves: the
Hilbert spectrum. Annu. Rev. Fluid Mech., 31:417–57, 1999.
15. S.M. Kay. Fundamentals of Statistical Signal Processing, Volume 2: Detection Theory.
Prentice Hall, Englewood Cliffs, N.J., 1998.
16. M.J. Morey, T. Bentson, and C. Osborn. TET E-field data analysis, summer 1991
experiment. Tech. Document 2482, NCCOSC RDT & E Divsion, 1993.
17. E.A. Nichols, J. Clarke, and H.F. Morrison. Signals and noise in measurements of
low-frequency geomagnetic fields. J. Geophys. Res., 93:13743–13754, 1988.
18. M.K. Ochi. Ocean Waves: The Stochastic Approach. Cambridge Ocean Technology,
Series 6. University Press, Cambridge, 1998.
19. A.M. Orr. Computational techniques for evaluating extremely low frequency electro-
magnetic fields produced by a horizontal electrical dipole in seawater. Dept. of
Physics, King's College London, 2001.
20. A. Quinquis and D. Declerq. Magnetic noise subtraction with wavelet packets. In
Proc. OCEANS '95. MTS/IEEE. Challenges of Our Changing Global Environment,
Vol. 1, pp. 227–232, 1995.
21. A. Quinquis and C. Dugal. Using the wavelet transform for the detection of magnetic
underwater transient signals. In Proc. OCEANS '94: Oceans Engineering for Today's
Technology and Tomorrow's Preservation, Vol. 2, pp. 548–543, 1992.
22. A. Quinquis and S. Rossignol. Detection of underwater magnetic transient signals
by higher order analysis. In Proc. Int. Symp. Signal Process. Appl., pp. 741–749, 1996.
23. J. Rantakokko. Localization of static electric and magnetic underwater sources. Tech-
nical Report FOA report No. FOA-R–99-01265-409–SE, Swedish Defense Research
Establishment (F.O.A.), Stockholm, Sweden, 1999.
24. J. Ridgway, M.L. Larsen, C.H. Waldman, M. Gabbay, R.R. Buntzen, and C.D. Rees.
Analysis of ocean electromagnetic data using a Hilbert spectrum approach. In Proc.
7th Exper. Chaos Conf., pp. 333–338, August, 2003.
25. G. Rilling, P. Flandrin, and P. Gonclaves. On empirical mode decomposition and its
algorithms. In Proc. IEEE-EURASIP Workshop Nonlinear Signal Image Process.
NSIP-03, 2003.
26. T.B. Sanford. Motionally induced electric and magnetic fields in the sea. J. Geophys.
Res., 76:3476–3492, 1971.
27. S. Webb and C.S. Cox. Pressure and electric fluctuations on the deep seafloor:
background noise for seismic detection. Geophys. Res. Lett., 11:967–970, 1984.
28. H.C. Yuen and B.M. Lake. Nonlinear dynamics of deep-water gravity waves. Adv.
Appl. Mech., 22:67–229, 1982.
7 Coherent Structures
Analysis in Turbulent
Open Channel Flow
Using Hilbert-Huang and
Wavelets Transforms
Athanasios Zeris and Panayotis Prinos
CONTENTS
ABSTRACT
141
7.1 INTRODUCTION
Details for the instrumentation and experimental results from measurements of the
application of suction from the bed in open channel have been presented in previous
articles [3,4,5]. We mention for completeness of the present study that measurements
were conducted with LDA (tracker) and hot film methods, the suction rate is based
on discharge ratio (Qs/Qtot, Qs = suction discharge, Qtot = total discharge), the length
of suction region L = 0.035 m, and the Re number based on upstream velocity Um
and flow depth was Re = Umh/ = 59,000.
The empirical mode decomposition (EMD) method proposed by Huang [1] and
the Hilbert transform of the decomposed signal are summarized as follows:
All the extrema of the signal x(t) are identified, and an interpolation for the
maxima and the minima is attempted with cubic splines curves, following recom-
mendations and proposals from previous work [1,2]. The local mean m(t) is com-
puted, and the difference between the local mean m(t) and the initial signal is
extracted. Until a stopping criterion [2] is achieved, an iteration process takes place.
In this study the criterion that we adopt is that the number of extrema equals (or
differs by one from) the number of zero-crossings. When this criterion has been
achieved, the remainder is considered as an intrinsic mode function (IMF) and
denoted c1; this IMF contains the shortest periods of the signal. The residual
() ()
x t − c1 = r t ,
is considered as a new signal and subjected to the iterative process repeatedly, until the
last residue is smaller than a predetermined small value or is a monotonic function.
Preliminarily, the problem of extrema interpolation is checked from the view-
point of (a) it is time consuming especially when the signal belongs to the multi-
frequency ones as turbulent signals that needs a large number of iterations for a
correct decomposition and (b) because of end effects propagation of the splines into
the interior of the signal. We used different approaches as linear, linear with empirical
modifications, and iterative process with preselected iteration numbers. With the
criterion of the most perfect composition of the original signal we chose, for the
present article, to extract IMFs with the originally proposed method [2] cubic spline
with the addition of sine waveforms at the beginning and at the end of the analyzed
data.
With the aid of the Hilbert transform, Huang et al. [1,2] proposed the construction
of the analytic signal for every IMF component. From them we get the time distri-
butions of amplitude and phase
() () ()
2
h t = c t + H c t
()
φ t = tan −1
( ( )) ,
H c t
()
c t
where h(t) and φ(t) are the instantaneous amplitude and phase respectively. From
this quantity, values for the instantaneous frequency are obtained as
()
f t =
()
dφ τ
dt
With the extraction of IMFs as a time series, the tools for the HHT ampli-
tude–frequency–time representation of a signal — the instantaneous frequency and
the amplitude — are available, and, depending on the signal’s nature, their quality
is better than or equal to the results of other well documented methods, such as
short FFT, proper orthogonal decomposition, Wigner-Ville, and the wavelet decom-
position method.
7.2 RESULTS
7.2.1 ACADEMIC SIGNALS
Some academic signals are used for the estimation of the quality of decomposition
applied in the present study. The traditional frequency shift sine signal (Figure 7.1a)
is examined through frequency time variation of the first IMF with the point of
frequency shift well localized. Figure 7.1c also gives the Hilbert spectrum and
Figure 7.1d presents the FFD distribution.
In Figure 7.2a the sawtooth signal is presented along with the distribution of
the instantaneous frequency (Figure 7.1c) because this kind of variation is a candidate
(and as the edge detection problem) as part of an average representative pattern
indicating the presence of organised structures.
(a)
1
0.8
0.6
0.4
0.2
0
−0.2
−0.4
−0.6
−0.8
−1
0 1 2 3 4 5
t (sec)
(b)
40
35
30
25
f (Hz)
20
15
10
0
0 1 2 3 4 5
t (sec)
(c) (d)
2
0.7
0.6
1.5
0.5
EH (f )
0.4
E (t)
1
0.3
0.5 0.2
0.1
0 0
1 10 100 0 5 10 15 20
f (Hz) f (Hz)
FIGURE 7.1 (a) Frequency shift signal; (b) instantaneous frequency variation; (c) Hilbert
spectrum; (d) FFT distribution.
FIGURE 7.2 Sawtooth signal with (a) instantaneous frequency (b) variation with time.
2
0
18
16
14
12
10
8
6
4
2
0
(b)
f (t)
6
4
2
-0.5 0
-1
1.5
0.5
-1.5
(a)
© 2005 by Taylor & Francis Group, LLC
DK342X_book.fm Page 146 Sunday, May 1, 2005 12:34 PM
u(m/s) signal
0.4
0.2
0
0 5 10 15 20 25 30 35 t(sec) 40
interval time average (VITA) technique [6,7], quadrant splitting [8] of Re stresses,
the temporal pattern average (TPAV) [10], linear stochastic estimation [9], and the
proper orthogonal decomposition (POD) and wavelet decomposition–based tech-
niques [11].
Traditional techniques, such as VITA (focusing on local variance that exceeds
part of the whole signal variance) and quadrant splitting for velocity and Re stress,
with empirical threshold criteria k = 1.0, predetermined time scale T* = 10, and
H = 1 for quadrant splitting, respectively, have been applied in previous studies [3,4]
to extract intense events in the flow field inside the porous suction region (x/L =
0.5) and beyond the suction region (x/L = 1.2) from hot film and LDA results. Figure
7.8 gives an example of results from the VITA technique. Here the effect of suction
is evident at the exit of the suction region from the frequency appearance of the
intense events.
With the tool of signal decomposition, it is possible to overcome the shortcom-
ings resulting from the single scale limitations of the VITA method. The VITA
method does not permit the identification of events separated by a time smaller than
that imposed by the VITA treatment. Also, a main disadvantage for VITA is the fact
that it requires the selection of two rather subjective criteria (time scale and thresh-
old).
Wavelets decomposition allows intense events to be identified based on the value
of the wavelet coefficients in the representation of the translation — scale plane
using a nonsubjective threshold criterion.
Analysis in the scale time plane of turbulent statistical measures gives scale
dependent behavior for quantities such skewness and flatness factors. Strong
non-Gaussian behavior of the smaller scales is an indication of the intermittency.
For intermittency estimation at each scale, Farge [11] proposed the local intermit-
tency index Im,n , the ratio of local energy to the mean energy at the respective scale.
As this index takes an extreme value at a time instant, a strong percentage of energy
and intermittency is found in the corresponding time and scale.
With the above in mind, it is possible to use EMD to determine the ratio of local
energy (amplitude squared) to mean energy for every IMF in the energy–fre-
quency–time distribution. Figure 7.9 shows the above ratios with the 4th order Daub4
wavelet and for the HHT method.
The next step in the identification of coherent events is to apply the appropriate
thresholds. Universal proposed relations from wavelets such as Donoho and
Johnstone [12] is considered that cannot contribute to these turbulent data because
of the Gaussian consideration of the incoherent part.
In this study, we applied different values for thresholding the ratio of instanta-
neous energy for every IMF mode. We propose the selection of a threshold value in
the region where the number of the detected events are not independent from the
0.1 IMF2
0.05
0
–0.05
–0.1
IMF3
0.1
0.05
0
–0.05
–0.1
0.1 IMF4
0.05
0
–0.05
–0.1
0.1 IMF5
0.05
0
–0.05
–0.1
0.1 IMF6
0.05
0
–0.05
–0.1
0.1 IMF7
0.05
0
–0.05
–0.1
0.1 IMF8
0.05
0
–0.05
–0.1
0.1 IMF9
0.05
0
–0.05
–0.1
0.1 IMF10
0.05
0
–0.05
–0.1
a)
1.2E-00
1.3E-00
1.4E-00
1.5E-00
1 10 f(H) 100
b)
1.E-01
1.E-02
1.E-03
1.E-04
1.E-05
1 10 f(H) 100
60
50
40
f(Hz)
30
20
10
200 400 600 800 1000 1200 1400 1600 1800 2000
N
select the corresponding points at every IMF that allow the coherent signal to be
distinguished from the noncoherent signal in the reconstruction formula. As an
example, Figure 7.11 presents the reconstruction of the organized and disorganized
part of the signal applying a hard thresholding method for the Hilbert-Huang trans-
form.
With well-known strategies [14,15], it is possible to average the coherent pattern
[6,7,13,14] and correct the phase jitter (through an iterative process with cross-cor-
relation and a time shift between the ensemble average pattern and each individual
pattern), in order to extract the averaged patterns for every turbulent quantity as
streamwise, vertical velocities, and Re stresses.
Figure 7.12 illustrates the shortcomings of the corresponding wavelets recon-
struction in the case of the Daub4 wavelet, where the signature of the wavelet itself
is observed in the reconstruction formula (N = 1024, m = 2, scale reconstruction).
Figure 7.13 presents an IMF analysis (sum from coarser to finer scales Ri) of the
intense events for two regions of the flow field where suction is applied in open
channel flow. A stronger effect of suction is observed near the wall and at the exit
of the suction region.
Because of the indication of organized activity of the non-Gaussian behaviour of
the turbulent data, we propose the analysis of the skewness and kirtosis coefficients,
(a)
1
0.8
0.6
u (t)
0.4
0.2
0
0 2 4 6 8 10
(b)
0.5
0.15
IE (t)
0.1
0.05
0
0 2 4 6 8 10
from the reconstruction formula (starting the composition of the initial signal from
residual and the larger scales). Figure 7.14 a presents the skewness coefficient Sk
for a short signal of N = 8192. The value of the whole signal is –0.23. Figure 7.14b
shows a scale analysis of the same signal with Daub4 wavelet.
Figure 7.15 shows, analysis with distributions for every IMF, of the turbulent
quantities with the same length of signal for comparative reasons. Figure 7.15a
presents the energy content of streamwise component Eiu and the vertical component
Eiv, as percentage of the whole energy resulting from hot film measurements that
lead to conclusions for the principal time scales. Figure 7.15b shows the ratio Eiu/Eiv.
Figure 7.15c gives the percentage of energy for every IMF, from LDA measurements
with and without suction application near the wall at the exit of the suction region.
7.3 CONCLUSIONS
Having established an approach using empirical mode decomposition to study open
channel flow, we must mention the need for a systematic application in different
flow situations. Nonstationary flows, vortex shedding over obstacle, and wake flows
may comprise an ideal domain for application of this new technique. Further doc-
umentation of the sifting process is necessary, including the stopping criteria and
extrema interpolation, especially for difficult environments such as turbulent flow
where the appearance of all the frequency bands makes the overall process time
consuming. Applications that demand real-time processing, such as active control,
must be investigated under this new method with concepts such as the modulated
1 Cq = 0
Cq = 0.056
0.9
0.8
0.7
0.6
0.5
f
0.4
0.3
0.2
0.1
0
0.001 0.01 0.1 1
y/h
FIGURE 7.8 Frequency of events with VITA analysis (y+ = 8, x/L = 1.2).
intra-wave frequency. Nonstationary flows with short time variations would be the
field for comparison of HHT with other older techniques, such as those of entropy
(MEM)–based methods. However, the final general impression remains the ability
of this novel technique to exploit the Hilbert transform and the analytic signal for
multifrequency signals to achieve better localization in the frequency time domain
and avoid the shortcomings of wavelets coefficients, such as the strong smear effect.
(a)
700
600
500
400
HHTLIM
300
200
100
(b)
100
90
80
70
60
LIM
50
40
30
20
10
(a)
1
0.8
0.6
0.4
0.2
0
0 2 4 6 8 10 12 14 16
t (sec)
(b)
35 IMF2
30 IMF3
25 IMF4
f (Hz)
20 IMF5
15
10
5
0
0 2 4 6 8 10 12 14 16
t (sec)
(c)
10
HHTLIM IMF2
8
6
4
2
0
0 2 4 6 8 10 12 14 16
t (sec)
10
HHTLIM IMF3
8
6
4
2
0
0 2 4 6 8 10 12 14 16
t (sec)
10
HHTLIM IMF4
8
6
4
2
0
0 2 4 6 8 10 12 14 16
t (sec)
10
HHTLIM IMF5
8
6
4
2
0
0 2 4 6 8 10 12 14 16
t (sec)
FIGURE 7.10 (a) Velocity signal; (b) instantaneous frequency variations; (c) HHTLIM with
thresholds.
(a)
0.2
0.1
−0.1
−0.2
(b)
0.2
0.15
0.1
0.05
0
−0.05
−0.1
−0.15
−0.2
0.4
0.35
u (m/s)
0.3
0.25
0.2
0 0.5 1 1.5 2 2.5
t (sec)
1.6
Y+ = 8, x/L = 1.2
1.4
Ns/N0 events
1.2 Y+ = 100, x/L = 0.5
0.8
0.6
0.4
0.2
0
1 2 3 4 5
IMF
FIGURE 7.13 Ratio of intense events for suction and no-suction cases.
1 Sk a) 1 Sk b)
0.5 0.5
Ri m
0 0
1 2 3 4 5 6 7 8 9 10 11 1 2 3 4 5 6 7 8 9 10 11 12
-0.5 -0.5
-1 -1
FIGURE 7.14 (a) Sk distributions with residuals; (b) Sk distribution with scales Daub4.
30 Eiu/Eiv
0.2 Ei/Eiall u 25
0.16 v 20
0.12 15
0.08 10
0.04 5
IMF IMF
0 0
1 2 3 4 5 6 7 8 9 10 11 12 13 14
1 2 3 4 5 6 7 8 9 10 11 12 13
FIGURE 7.15 (a) Percentage contribution of Eiu and Eiv; (b) ratio Eiu/Eiv; (c) effect of
suction near the wall (y+ = 10).
REFERENCES
1. Huang N., Shen Z., Long S. (1999). A new view of nonlinear water waves: The
Hilbert Spectrum. Annu. Rev. Fluid Mech. 31: 417–457.
2. Huang N. E., Shen Z., Long S., Wu M., Shih H., Zheng Q. et al. (1998). The Empirical
Mode Decomposition and the Hilbert Spectrum for non linear and non stationary
time series analysis. Proc. R. Soc. London A 454, 903–995.
3. Zeris A. (2001). Turbulent flow in open channel with suction from the bed. Ph.D.
thesis, Aristotle University, Thessaloniki, Greece, 2001.
4. Zeris A., Prinos P. (2001). Measurements of bed suction effects on an open channel
flow with Laser–Doppler/Hot-Film Anemometry. EFHT Congr., pp. 1099–1103.
5. Zeris A., Prinos P. Open channel flow with suction from the bed. J. Hydraulic Eng.,
submitted 2004.
6. Blackwelder R., Haritonidis J. Scaling of the bursting frequency in turbulent boundary
layers. J. Fluid Mech. Vol. 132, 87–103, 1983.
7. Bogard, D. G., Tiederman W. G. (1986). Burst detection with single point velocimetry
measurements. J. Fluid Mech. Vol. 162, 389–413.
8. Lu, S. S., Wilmarth, W. W. (1973). Measurements of structure of Reynolds stress in
a turbulent boundary layer. J. Fluid Mech. Vol. 60, 481–511.
9. Adrian R. J. On the role of conditional averages in turbulence theory. In Turbulence
in Liquids, Zakin and Patterson, Eds., Princeton Science Press, 1977.
10. Wallace, J. M., Eckelmann, H., Brodkey, R. S. (1972). The wall region in turbulent
shear flow. J. Fluid Mech. Vol. 54, 39–48.
11. Farge M. (1992). Wavelet transforms and their applications to turbulence. Annu. Rev.
Fluid Mech. Vol. 24, 395–497.
12. Donoho D., Johnstone I. M. (1994). Ideal spatial adaptation by wavelet shrinkage.
Biometrika, Vol. 81, 425–455.
13. Camussi R., Gui G. (1997). Orthonormal wavelet decomposition of turbulent flows:
intermittency and coherent structure. J. Fluid Mech. Vol. 348, 177–199.
14. Raupach, M. (1981). Conditional statistics of Reynolds stress in rough wall and
smooth wall turbulent flow. J. Fluid Mech. Vol. 108, 363–382.
15. Subramanian C. S., Rajagopalan S., Antonia R., Chambers A. (1982). Comparison
of conditional sampling and averaging techniques in a turbulent boundary layer. J.
Fluid Mech. Vol. 23, 335–362.
8 An HHT-Based Approach
to Quantify Nonlinear
Soil Amplification and
Damping
Ray Ruichong Zhang
CONTENTS
ABSTRACT
This study proposes to use a method of nonlinear, nonstationary data processing and
analysis, i.e., the Hilbert-Huang transform (HHT), to quantify influences of soil non-
linearity in earthquake recordings. The paper first summarizes symptoms of soil non-
linearity shown in earthquake ground motion recordings. It also reviews the Fou-
rier-based approach to characterizing the nonlinearity in the recordings and
demonstrates the deficiencies. It then offers the justifications of the HHT in addressing
the nonlinearity issues. With the use of the 2001 Nisqually earthquake recordings and
results of the Fourier-based approach as a reference, this study shows that the
159
The first HHT-based component and the Hilbert amplitude spectra can identify
abnormal high-frequency spikes in the recording at sites where strong soil
nonlinearity occurs; this can help to detect the nonlinear sites at a glance.
The HHT-based factor for site amplification is defined as the ratio of marginal
Hilbert amplitude spectra, similar to the Fourier-based one that is the ratio of
Fourier amplitude spectra. The HHT-based factor is effective in quantifying soil
nonlinearity in terms of frequency downshift in the low-frequency range and
amplitude downshift in the intermediate-frequency range.
Hilbert and marginal damping spectra are identified in ways similar to Hilbert
and marginal amplitude spectra. Consequently, the HHT-based factor for site
damping is found as the difference of marginal Hilbert damping spectra, which
can be extracted from the HHT-based factor for site amplification and used as
an alternative index to measure the influences of soil nonlinearity in seismic
ground responses.
8.1 INTRODUCTION
Site amplification is the phenomenon in which the amplitude of seismic waves
increases significantly when they pass through soil layers near the earth’s surface.
It can be illustrated by considering the seismic energy flux along a tube of seismic
rays, which is proportional to the impedance (density × wave speed) and squared
shaking velocity. Since the energy should be constant in the absence of damping,
any reduction in the impedance is compensated by an increase in the shaking velocity,
thus yielding site amplification for seismic waves in soil layers. Site amplification
is a key factor in mapping seismic hazard in urban areas (e.g., [1]) and designing
geotechnical and structural engineering systems on soils (e.g., [2]).
In general, site amplification is not linearly proportional to the intensity of input
seismic motion at bedrock because of soil nonlinearity under large-amplitude earth-
quakes. The extent of soil nonlinearity can be characterized by the change of two
dynamic features of soil layer, i.e., soil resonant frequency and damping, in the
frequency-dependent site amplification. Consensus has been building that the
site-amplification factors in the current codes overemphasize the extent of soil
nonlinearity and thus potentially underestimate the level of site amplification. It has
been demonstrated [3] that the recording-based amplification factors are larger than
those in codes for a certain range of base acceleration intensity. In addition, some
features of site-amplification factors used in codes and guidance for structural design
contradict recent findings from the 1994 Northridge ground motion data set [4].
The aforementioned problem might exist partly because seismologists and engi-
neers lack sufficient understanding of the underlying causes in nonlinear soil. For
example, the influence of soil heterogeneity does not scale linearly, even when the
soil is perfectly linear [5]. In other words, a linear elastic medium with random
heterogeneity can change ground motion in a way similar to that caused by medium
N N
X (t ) = ℜ ∑j =1
Aje
iΩ j t
=ℜ ∑[ A sin(Ω t) + iA cos(Ω t)] ,
j =1
j j j j (8.1)
where ℜ denotes the real part of the value to be calculated, i = (–1)∫ is an imaginary
unit, amplitudes Aj are a function of time-independent frequency Ωj that is defined
over the window in which the data is analyzed, and the Fourier amplitude spectrum
is defined as
F (Ω) = ∑A .
j =1
j (8.2)
To apply this Fourier spectral analysis to estimate the influences of soil nonlinearity
in the seismic wave responses at soil site or simply site amplification, two sets of
recordings are typically needed [18], one at a soil site and the other at a referenced
site such as bedrock or outcrop. For a frequency, the Fourier-based factor of site
amplification (FF) for an earthquake event (either mainshock or aftershock) can then
be found by
Fs2,h1 + Fs2,h 2
FFs Ω = ( ) , (8.3)
Fr2,h1 + Fr2,h 2
where subscripts s and r denote respectively the soil and referenced sites, and
subscripts h1 and h2 denote the two horizontal components. Note that Equation 8.3
is one of many representatives for site-amplification factor that can be the ratio of
characteristics of seismic waves or spectral responses at a site versus referenced site.
Since the wave paths and earth structures except the soil layer are almost the
same for the soil and referenced sites, the factor for site amplification in Equation
8.3 eliminates approximately the influences of source from the earthquake event and
thus provides essentially the dynamic characteristics of the soil. In addition, the
recordings at the referenced site are generally believed to be the results of linear
wave responses and the recordings at the soil site subject to the large-amplitude
mainshock to be the results of nonlinear wave responses. Accordingly, comparing
the factors from the mainshock and the aftershock could help us explore and quantify
the influences of soil nonlinearity in site amplification.
While the Fourier-based approach given here and similar methods are widely
used, they have the following deficiencies in characterizing the nonstationarity of
the earthquake motion that is caused by source, different types of propagating waves,
and soil nonlinearity if the earthquake magnitude is large enough.
A Fourier-based approach defines harmonic components globally and thus yields
average characteristics over the entire duration of the data. However, some charac-
teristics of data, such as the downshift of soil resonant frequency at a nonlinear site,
may occur only over a short portion of a record. This is particularly true when the
intensity of the seismic input to a soil layer is not strong, such that the soil becomes
nonlinear over only a portion of the entire duration of motion and in only a certain
frequency band. As a result, the averaging characteristic in Fourier spectral analysis
makes it insensitive for identifying time-dependent frequency content. While a
windowed (or short-time) Fourier-based approach can be used to improve the above
analysis to a certain extent, it also reduces frequency resolution as the length of the
window shortens. Thus, one is faced with a tradeoff. The shorter the window, the
better the temporal localization of the Fourier amplitude spectrum, but the poorer
the frequency resolution, which directly influences the measurement of downshift
of soil resonance that typically arises in a low to intermediate frequency band.
More important, a Fourier-based approach explains data in terms of a linear
superposition of harmonic functions. Therefore, it is an appropriate, effective method
for characterizing linear phenomena such as waves with time-independent frequency,
rather than nonlinear phenomena with time-dependent frequency. An example of
time-independent and time-dependent frequency waves is a hypothetical wave record
y(t) = y1(t) + y2(t), where decaying waves y1(t) = cos[2πt + εsin(2πt)]e–0.2t have
time-dependent frequency of 1 + εcos(2πt) Hz, with ε denoting a constant factor,
and noise y2(t) = 0.05sin(30t) has time-independent frequency of 15 Hz. Note that
the waves shown in Figure 8.1 with ε = 0.5 are physically related to one type of
1.5
0.5
Amplitude
−0.5
−1
−1.5
0 1 2 3 4 5 6 7 8 9 10
Time (sec)
FIGURE 8.1 A hypothetical wave recording, consisting of nonlinear waves and noise with
frequencies 1 + 0.5cos(2πt) Hz and 15 Hz, respectively.
FIGURE 8.2 Fourier and marginal Hilbert amplitude spectra of the recording in Figure 8.1.
water waves that result from a nonlinear dynamic process and are also representative
of seismic responses at a nonlinear soil site (to be elaborated).
The time-dependent frequency waves can be expanded into and thus interpreted
by a series of time-independent frequency waves, as done by the Fourier spectral
analysis in which y(t), or y1(t) in particular, can be interpreted as to contain Fourier
components at all frequencies (see Equation 8.1, Figure 8.2, and Figure 8.3).
Alternatively, the expansion of y1(t), i.e.,
suggests that the Fourier transform of y1(t) consists primarily of two harmonic
functions centered respectively at 1 Hz and 2 Hz for ε << 1, and the widths of these
harmonic functions are proportional to the exponential parameter 0.2, which is
related to the damping factor. Note that Figure 8.1 and Figure 8.2 use ε = 0.5, which
is not a small number in comparison to unity, and thus they have the third observable
harmonic function at 3 Hz in Figure 8.2. Therefore, one can equally well describe
y1(t) by saying that it consists of just two frequency components, each component
having a time-varying amplitude that is proportional to e–0.2t. Indeed, if one were to
examine the local behavior of y1(t) in the neighborhood of a given time instant, say
FIGURE 8.3 Fourier components (fj, j = 1, 2, 3, 4, 5) of the recording in Figure 8.1 at selected
frequencies (i.e., 10 Hz, 5 Hz, 2 Hz, 1 Hz, and 0.5 Hz).
t0, this is precisely what one would observe. The Fourier-based analysis or interpre-
tation given here can also be seen in Priestley [19] and Zhang et al. [20], among
others.
Because the true frequency content of the waves y1(t) is bounded between 1 –
ε and 1 + ε, much less than 2 Hz, analysis of the above example suggests that Fourier
spectral analysis typically needs higher-frequency harmonics (at least 2 Hz for the
example) to simulate the nonlinear waveform of the data. Stated differently, Fourier
spectral analysis distorts the nonlinear data. Consequently, the Fourier-based
approach in Equation 8.3 twists the influences of soil nonlinearity in site amplifica-
tion. The above assertions are confirmed in Huang et al. [21] and Worden and
Tomlinson [22], among others, with the aid of solutions to classic nonlinear systems
such as the Duffing equation in general, and in Zhang et al. [23] with the nonlinear
site amplification in particular.
In theory, Fourier spectral analysis in general and Fourier-based approaches for
site amplification in particular can be further used for evaluating damping factor.
For example, the resonant amplification method or half-power method uses the
amplitude change or width of the peaks at a certain frequency in the Fourier ampli-
tude spectrum to find the damping factor of dynamic systems such as a soil layer
(e.g., [24]). However, the distorted Fourier amplitude spectrum for nonlinear data
will mislead the subsequent use for damping evaluation with nonlinear soil. For
example, the damping factor evaluated at the first and second peaks in the Fourier
amplitude spectrum in Figure 8.2 suggests that the damping is associated with
frequency at 1 Hz and 2 Hz. In fact, the damping of the hypothetical record is
dependent only on the true frequency content of the waves y1(t) bounded between
1 – ε and 1 + ε, or 0.5 Hz and 1.5 Hz with ε = 0.5 in Figure 8.1. Accordingly, a
Fourier-based approach would misrepresent the influences of damping factor as it
relates to soil nonlinearity.
n n
X (t ) = ℜ ∑ a (t)e
j =1
j
iθ j ( t )
=ℜ ∑[C (t) + iY (t)] ,
j =1
j j (8.5)
where Cj(t) and Yj(t) are respectively the jth IMF component of X(t) and its Hilbert
transform,
C j (t ′)
∫
1
Yj (t ) = P dt ′ ,
π t − t′
where P denotes the Cauchy principal value, and the time-dependent amplitudes
aj(t) and phases θj(t) are the polar-coordinate expression of Cartesian-coordinate
expression of Cj(t) and Yj(t), from which the instantaneous frequency is defined as
d θ j (t )
ω j (t ) = . (8.6)
dt
H (ω , t ) = ∑ a (t) ,
j =1
j (8.7)
and its square gives the temporal evolution of the energy distribution. The marginal
Hilbert amplitude spectrum, h(ω), defined as
h(ω ) =
∫ H (ω, t)dt ,
0
(8.8)
provides a measure of the total amplitude or energy contribution from each frequency
value, in which T denotes the time duration of the data.
In comparison with Equation 8.2, the Hilbert amplitude spectrum H(ω,t) provides
an extra dimension by including time t in motion frequency and is thus more general
than the Fourier amplitude spectrum F(Ω). While the marginal amplitude spectrum
h(ω) provides information similar to the Fourier amplitude spectrum, its frequency
term is different. The Fourier-based frequency (Ω) is constant over the sinusoidal
harmonics persisting through the data window, as seen in Equation 8.1, while the
HHT-based frequency ω varies with time based on Equation 8.6. As the Fourier
transformation window length reduces to zero, the Fourier-based frequency (Ω)
approaches the HHT-based frequency (ω). The Fourier-based frequency is locally
averaged and not truly instantaneous, for it depends on window length, which is
controlled by the uncertainty principle and the sampling rate of data.
Recordings that are stationary and linear can typically be decomposed or rep-
resented by a series of time-independent frequency waves through the Fourier-based
approach in Equation 8.1. If the jth IMF component, i.e., Cj(t) in Equation 8.5,
corresponds to a Fourier component with a sine function at a time-independent
frequency, the Hilbert transform of the sine function, i.e., Yj(t) in Equation 8.5, can
be found to equal the cosine function at the same frequency in opposite sign. Because
the sign can be changed with adding a constant phase, the above analysis essentially
leads to the consistence between Fourier- and HHT-based approaches in general,
hs2,h1 + hs2,h 2
FH s ω = ( ) . (8.9)
hr2,h1 + hr2,h 2
n n
X (t ) = ℜ ∑
j =1
a j (t )e
iθ j ( t )
=ℜ ∑ Λ (t)e
j =1
j
− ϕ j ( t )+ iθ j ( t )
(8.10)
− ϕ j (t )
a j (t ) = Λ j (t )e . (8.11)
d ϕ j (t )
η j (t ) = . (8.12)
dt
With the aid of Equation 8.11, the Hilbert damping spectrum can be found as
n n
a j (t ) (t )
Λ
D(ω , t ) = ∑j =1
η j (t ) = ∑ − a (t) + Λ (t) .
j =1 j
j
j
(8.13)
T n T
(t )
a j (t ) Λ
d (ω ) =
∫ D(ω , t )dt = ∑∫
j =1 0
− + j dt = d a (ω ) + d Λ (ω ) .
a j (t ) Λ j (t )
(8.14)
0
Equation 8.14 indicates that the marginal Hilbert damping spectrum consists of
two terms: one is from the time-dependent amplitudes aj(t) that are related to
marginal and Hilbert amplitude spectra, and the other is from source-related intensity,
i.e., time-dependent amplitudes Λ j(t).
It is of interest to note that the definition of instantaneous damping factor in
Equation 8.12 and subsequent spectra in Equation 8.13 and Equation 8.14 are
different from those in Salvino [26] and Loh et al. [27]. For recordings of
impulse-induced or ambient linear vibration responses, some IMF components can
be extracted from the data that are related to certain vibration modes [28–30].
Consequently, Λj(t) are constant and ηj(t) are proportional to the damping ratio and
damped frequency. The modal damping ratio can then be found. This is essentially
the same as those in Salvino [26] and Loh et al. [27], if the latter can judicially
relate the IMF components to the vibration/wave modes.
For recordings to an earthquake, Λj are functions of time and unknown, which
are dependent upon the seismic source. Their influences in the site amplification,
however, can be removed if two recordings at soil and referenced sites are used.
Similar to the HHT-based factor of site amplification, the difference of marginal
Hilbert damping spectra at soil and referenced sites, or HHT-based factor of site
damping, could approximately eliminate the influences of the source that is associated
0.5
0
c1
−0.5
0.5
c2
0
−0.5
0.02
c3
0
−0.02
0.02
c4
0
−0.02
×10−3
10
5
c5
0
0 1 2 3 4 5 6 7 8 9 10
Time (sec)
FIGURE 8.4 The five IMF components of the recording in Figure 8.1.
with Λj and thus provide essentially the characterization of the damping in the soil
site. The HHT-based factor of site damping can be found as
where use has been made in the last approximation of the fact that the source-related
damping terms at the soil and referenced sites are approximately equal, i.e., dsΛ (ω)
drΛ (ω).
Finally, comparing the HHT-based factors of site damping from the mainshock
and the aftershock could help quantify the influences of nonlinear soil damping in
site responses.
To illustrate the HHT-based characterization of nonlinearity, the hypothetical
record in Figure 8.1 is analyzed again. Figure 8.4 shows the five IMF components
decomposed from the data by EMD. The first and second components (c1 and c2)
capture the noise and primary waveform, while the other three (c3 to c5), with
negligible amplitudes, represent the numerical error in the EMD process. Comparing
Figure 8.3 and Figure 8.4 suggests that some IMF components can be not only more
physically meaningful than the Fourier components, they can also be in principle
used to explore the damping factor with the use of, e.g., the consecutive peak values
18 0.9
16 0.8
14 0.7
12 0.6
Frequency (Hz)
10 0.5
8 0.4
6 0.3
4 0.2
2 0.1
0
0 1 2 3 4 5 6 7 8 9 10
Time (sec)
and corresponding time elapse in the second IMF component, if that component is
related to a free-vibration response or forced response at a certain mode.
The Hilbert amplitude spectrum in Figure 8.5 shows a clear picture of tempo-
ral–frequency energy distribution of the data, i.e., primary waves with frequency
dependence modulated around 1 Hz and bounded by 0.5 Hz and 1.5 Hz, noise at
15 Hz, and the decaying energy of the primary waves with the color changing from
the red/yellow at the beginning to the dark blue at the end of the record. In contrast,
the Fourier amplitude spectrum in Figure 8.2 not only loses the information pertain-
ing to temporal characteristics of the motion, but, more important, it also distorts
the information of the record by introducing higher-order harmonics, notably at 2
Hz and 3 Hz. For comparison, the marginal amplitude spectrum of the recording is
also plotted in Figure 8.2, showing truthfully the energy distribution of the motion
in frequency.
Figure 8.6 shows the marginal Hilbert damping spectrum, calculated without
using any smooth function, which was not so done in the calculation of the above
marginal Hilbert amplitude spectrum with the use of Hilbert-Huang Transformation
Toolbox [31]. While oscillation of the curve in Figure 8.6 is originally due to the
numerical calculation of a· j(t) and a· j(t)/aj(t) that subsequently influences the compu-
tation of the spectra in Equation 8.13 and Equation 8.14, the estimated mean damping
factor at frequency 0.5 to 1 Hz is around the true value of 0.2. Compared with no
100
10−1
10−2
100 101
Frequency (Hz)
FIGURE 8.6 Marginal Hilbert damping spectrum of the recording in Figure 8.1.
damping in the record at frequency larger than 1.5 Hz, the very small mean damping
factor in Figure 8.6 (about 0.04, lower at a factor of five than the damping factor of
waves) due to the aforementioned unavoidable, cumulated numerical error is still
acceptable. It is believed that the oscillated curve in Figure 8.6 can be improved by
replacing it with the instantaneous mean curve. The mean curve can be found by
many methods, one of which is the use of the summation of all the IMF components,
excluding the first IMF component, that are extracted from the data of the oscillated
curve with one sifting process. The large damping factors at around 1.5 Hz are due
to the numerical error caused by the transition of two damping factors from 0.2 to
0.04, although such abrupt damping change is unlikely in practice. The large damp-
ing factor below 0.5 Hz is caused by the high-order, low-frequency IMF components
(primarily from the third to fifth IMFs). The error in overestimating damping factor
at very low frequency can be theoretically minimized if high-order, low-frequency
IMF components with very small amplitudes are judicially not used in the damping
calculation.
0.2
Acceleration (g)
0.1
−0.1
−0.2
0 5 10 15 20 25 30 35 40
0.05
Acceleration (g)
−0.05
−0.1
0 5 10 15 20 25 30 35 40
Time (sec)
FIGURE 8.7 (a, top) NS-acceleration recording and (b, bottom) its first IMF of the Nisqually
mainshock at SDS (soft soil).
148 m/s. The LAP is located over a stiff soil with Vs30 = 367 m/s. To examine the
soil nonlinearity from recordings at the two stations, recordings at SEW are used
as referenced ones, because Vs30 = 433 m/s at SEW is within the range of Vs30
values for typical rock sites in the western United States. Previous Fourier-based
studies (e.g., [17]) suggest that SDS experienced strong soil nonlinearity during the
mainshock while the LAP did not, with recordings at SEW as a reference.
0.2
Acceleration (g)
−0.2
−0.4
0 5 10 15 20 25 30 35 40
0.04
Acceleration (g)
0.02
−0.02
−0.04
−0.06
0 5 10 15 20 25 30 35 40
Time (sec)
FIGURE 8.8 (a, top) EW-acceleration recording and (b, bottom) its first IMF of the Nisqually
mainshock at SDS (soft soil).
spikes. It should be noted that there exist other tools to identify the spikes. For
example, Hou et al. [32] used a wavelet-based approach to characterize the spikes
from nonlinear vibration recordings in the vicinity of a damaged-structure location
subject to a severe earthquake. The disadvantages of these approaches, however,
reside with the subjective selection of frequency band in a Fourier-based approach
and subjective selection of another wavelet in a wavelet-based approach, among
others.
This study presents the effectiveness of the HHT-based approach in identifying
the spikes. Figure 8.7b depicts the first IMF component of the NS-component of
motion, clearly showing the two largest spikes between 20 sec and 25 sec. While
the spikes in the recording of Figure 8.7a can be visualized without using any tools,
the first IMF component in Figure 8.7b is shown simply for validation of the
HHT-based approach in identifying the spikes. Indeed, the EW-component of the
same recording in Figure 8.8a does not clearly show the spikes between 20 sec and
25 sec. The corresponding first IMF component in Figure 8.8b is, however, able to
reveal them explicitly, suggesting that the first IMF component is effective at detect-
ing the high-frequency spikes. This study also analyzed the horizontal recordings
of the ML3.4 aftershock at the same location in Figure 8.9 and Figure 8.10, in which
one does not observe the spikes in the S-coda waves in the corresponding first IMF
0.5
Acceleration (g)
−0.5
−1
0 5 10 15 20 25 30 35 40
1
Acceleration (g)
−1
−2
0 5 10 15 20 25 30 35 40
Time (sec)
FIGURE 8.9 (a, top) NS-acceleration recording and (b, bottom) its first IMF of the Nisqually
aftershock at SDS (soft soil).
0.5
Acceleration (g)
−0.5
−1
0 5 10 15 20 25 30 35 40
−4
×10 1st Component Sds2
8
6
Acceleration (g)
−2
−4
0 5 10 15 20 25 30 35 40
Time (sec)
FIGURE 8.10 (a, top) EW-acceleration recording and (b, bottom) its first IMF of the
Nisqually aftershock at SDS (soft soil).
0.18
101 0.16
0.14
0.12
Frequency (Hz)
0.1
100 0.08
0.06
0.04
0.02
10−1
0 5 10 15 20 25 30 35 40
Time (s)
FIGURE 8.11 Hilbert amplitude spectra of NS-acceleration recording of the 2001 Nisqually
earthquake mainshock at SDS (soft soil).
• In the low-frequency range (below 2.5 Hz), Figure 8.16a shows a down-
shift profile in both frequency and amplitude from the aftershock to
mainshock that is similar to Figure 8.15a in the low to intermediate
frequency range, but the former shows a smaller shift (about 0.1 Hz in
0.1
101
0.09
0.08
0.07
Frequency (Hz)
0.06
100 0.05
0.04
0.03
0.02
0.01
−1
10
0 5 10 15 20 25 30 35 40
Time (s)
FIGURE 8.12 Hilbert amplitude spectra of EW-acceleration recording of the 2001 Nisqually
earthquake mainshock at SDS (soft soil).
frequency and a factor of 0.2) than the latter (about 0.7 Hz and a factor
of 0.43).
• In the intermediate to high frequency range, there is almost no difference
in the two factors between the mainshock and the aftershock.
Comparison of the HHT-based factors at SDS and LAP suggests that SDS had
severe soil nonlinearity during the mainshock and LAP had slight soil nonlinearity.
Site LAP can also be regarded as having no soil nonlinearity under the mainshock
if the aforementioned small downshift in both frequency and amplitude in the
low-frequency range is the result of variation of data collection and sampling.
−4
Nisqually Aftershock – 2-D Hilbert Spectrum - SDSns ×10
6
101
5
4
Frequency (Hz)
3
100
10−1
0 5 10 15 20 25 30 35 40
Time (s)
FIGURE 8.13 Hilbert amplitude spectra of NS-acceleration recording of the 2001 Nisqually
earthquake aftershock at SDS (soft soil).
−4
Nisqually Aftershock – 2-D Hilbert Spectrum - SDSew ×10
5.5
101 5
4.5
3.5
Frequency (Hz)
3
0
10 2.5
1.5
0.5
10−1
0 5 10 15 20 25 30 35 40
Time (s)
FIGURE 8.14 Hilbert amplitude spectra of EW-acceleration recording of the 2001 Nisqually
earthquake aftershock at SDS (soft soil).
the mainshock. The downshift in both amplitude and frequency from the HHT-based
factors between mainshock and aftershock is clearly seen in Figure 8.17b.
To further illustrate the characteristics of the HHT approach, this study compares
the HHT-based factors of site amplification at SDS shown in Figure 8.15a with the
Fourier-based ones shown in Figure 8.15b (i.e., Figure 8.7 in Frankel et al. [17]).
In the low-frequency range, Figure 8.15b shows a frequency-downshift profile
from the aftershock to mainshock that is similar to Figure 8.15a, but the former
shows a smaller shift (about 0.2 Hz) than the latter (about 0.7 Hz). Because of the
averaging characteristic in Fourier spectral analysis, as indicated in Section 8.3, the
frequency downshift measured from the HHT-based factors in Figure 8.15a may
give a more truthful indication of the soil nonlinearity than that measured from
Fourier-based factors in Figure 8.15b. In addition, the factor in the low-frequency
range in Figure 8.15a is generally somewhat larger than the factors in Figure 8.15b.
In the intermediate-frequency range, Figure 8.15b shows an amplitude-reduction
profile from the aftershock to mainshock that is similar to Figure 8.15a, but the
1
10
Spectral Ratio
0
10
Aftershock
Mainshock
−1
10
10−1 100 1
10
Frequency (Hz)
1
10
Spectral Ratio
100
Aftershock
Mainshock
10−1
10−1 100 1
10
Frequency (Hz)
FIGURE 8.15 (A) HHT-based factor for site amplification at SDS (soft soil) for mainshock
and aftershock of the 2001 Nisqually earthquake. (B) Fourier-based factor for site amplifica-
tion at SDS (soft soil) for mainshock and aftershock of the 2001 Nisqually earthquake.
1
10
Spectral Ratio
0
10
Aftershock
Mainshock
10−1
10−1 100 101
Frequency (Hz)
101
Spectral Ratio
100
Aftershock
Mainshock
10−1
10−1 100 101
Frequency (Hz)
FIGURE 8.16 (A) HHT-based factor for site amplification at LAP (stiff soil) for mainshock
and aftershock of the 2001 Nisqually earthquake. (B) Fourier-ased factor for site amplification
at LAP (stiff soil) for mainshock and aftershock of the 2001 Nisqually earthquake.
2
10
Spectral Ratio
1
10
0
10
Aftershock
Mainshock
−1 0
10 10 101
Frequency (Hz)
2
10
Spectral Ratio
1
10
0
10
Aftershock
Mainshock
−1 0
10 10 101
Frequency (Hz)
FIGURE 8.17 (A) HHT-based factor for site amplification with the use of recordings in
window 0 to 10 sec at SDS (soft soil) for mainshock and aftershock of the 2001 Nisqually
earthquake. (B) HHT-based factor for site amplification with the use of recordings in window
10 to 20 sec at SDS (soft soil) for mainshock and aftershock of the 2001 Nisqually earthquake.
Aftershock
Mainshock
10−1
Difference of Damping Factor
10−2
0 2 4 6 8 10 12 14 16 18 20
Frequency (Hz)
FIGURE 8.18 Difference of marginal Hilbert damping spectra at SDS and SEW for main-
shock and aftershock of the 2001 Nisqually earthquake.
former shows a smaller reduction (about a factor of 0.2) than the latter (about a
factor of 0.43).
In the high-frequency range, Figure 8.15b shows a significant increase in the
Fourier-based factor from the aftershock to mainshock, while Figure 8.15a does not.
Following the discussion in Section 8.3 and numerical illustration in Figure 8.1 and
Figure 8.2, the high-frequency content in the Fourier amplitude spectra may be
influenced by higher-order harmonics used to represent a nonlinear waveform, which
consequently increases the Fourier-based factor of the mainshock in the high-fre-
quency range.
This study also compares HHT- and Fourier-based factors at LAP in Figure 8.16a
and b. Almost no fundamental difference in the factors is observed in terms of overall
profile, amplitude value, frequency downshift, and amplitude reduction between the
mainshock and aftershock, indicating that the two approaches are essentially consistent
with each other in estimating linear or approximately linear site amplification.
reveals that the site damping during the mainshock is much larger than that in the
aftershock at frequency 0.5 to 5 Hz, suggesting that strong soil nonlinearity occurred
during the mainshock in this frequency band. The increased damping will decrease
the amplified magnitude of seismic wave responses through the nonlinear soil and
thus reduce the site amplification factor. This can be confirmed from Figure 8.15a,
which shows that the HHT-based factor for site amplification is observably reduced
for the mainshock from the aftershock in the similar frequency band of 0.5 to 7 Hz.
Figure 8.18 also shows that the second largest increased site damping for the
mainshock is in the frequency band 8 to 16 Hz. However, the HHT-based factors
for site amplification in Figure 8.15a do not appear to change between the mainshock
and aftershock. This can be explained as follows: On the one hand, the increased
site damping for the mainshock over the frequencies 8 to 16 Hz will reduce the
seismic wave responses in the same frequency band. Note that the soil nonlinearity
typically occurs under large-amplitude seismic waves incident to the soil layer, and
that the amplitude of the motion changes with time. This suggests that the soil
nonlinearity is likely most prevalent in the strong S-wave motion or following surface
or S-coda waves, but not in the P-wave motion during the mainshock. Figure 8.7a
and Figure 8.8a reveal that except for the abnormal high-frequency spikes between
20 sec and 25 sec, the mainshock accelerations after the S-wave arrival at about 10
sec contain less high-frequency motion than does the aftershock in general, and in
the high-frequency band of about 8 to 16 Hz in particular. This fact is consistent
with the increased damping of the nonlinear soil during the mainshock in the
high-frequency band of 8 to 16 Hz, which damped out the high-frequency wave
responses in the recordings.
On the other hand, the soil nonlinearity indeed introduces large-amplitude
high-frequency spikes from 20 to 25 sec in the mainshock recordings in Figure 8.7a
and Figure 8.8a.
Together with the fact that the HHT-based factors in Figure 8.15a show the
averaged characteristics of soil linearity and nonlinearity, these observations suggest
that the overall high-frequency (around 8 to 16 Hz) content of the mainshock may
be equivalent to that of the aftershock. This yields the same factors in Figure 8.15a
in the frequency range 8 to 16 Hz.
For further clarification, the HHT-based factor for site damping at LAP is
examined. Figure 8.19 shows that the profiles of soil damping are essentially no
different between the mainshock and aftershock, suggesting that site LAP is linear
during the mainshock. This can be further verified from the HHT-based site-ampli-
fication factors shown in Figure 8.16a.
Aftershock
Mainshock
10−1
Difference of Damping Factor
10−2
0 2 4 6 8 10 12 14 16 18 20
Frequency (Hz)
FIGURE 8.19 Difference of marginal Hilbert damping spectra at LAP and SEW for main-
shock and aftershock of the 2001 Nisqually earthquake.
The HHT-based factor for site damping can be extracted from the HHT-based
factor for site amplification and used as an alternative index to measure the influences
of soil nonlinearity in seismic wave responses at sites.
It should be pointed out that the results from this study rely mainly on the use
of advanced signal processing techniques to explore and then quantify the signature
of soil nonlinearity from recordings. They must, therefore, be validated by
model-based simulation. Recently, significant advances have been made in borehole
data collection and simulation techniques. These include data from 17 borehole
arrays in Southern California [33] and from the Port Island vertical array for the
1995 Hyogoken Nanbu earthquake, geophysical data including the S-wave velocity
profile in the top layer(s) at key strong motion station sites (Rosrine 2002 at
http://geoinfo.usc.edu/rosrine, USGS open file reports), and simulated broadband
ground motion for scenario earthquakes including nonlinear soil effects [34]. The
above information should allow the further validation of the observations and results
from this study.
ACKNOWLEDGMENTS
The author would like to express sincere gratitude to Norden E. Huang at NASA;
Stephen Hartzell, Authur Frankel, and Erdal Safak at USGS; Lance VanDemark
from Colorado School of Mines; and Yuxian Hu from China Seismological Bureau
and Jianwen Liang from Tianjin University of China for providing data, Fourier
analysis and calculations, and more important, constructive suggestions. This work
was supported by the National Science Foundation with Grant Nos. 0085272 and
0414363, and by the US-PRC Researcher Exchange Program administered by Mul-
tidisciplinary Center for Earthquake Engineering Research. The opinions, findings,
and conclusions expressed herein are those of the author and do not necessarily
reflect the views of the sponsors.
REFERENCES
1. Frankel, A., C. Mueller, T. Barnhard, D. Perkins, E. Leyendecker, N. Dickman, S.
Hanson, and M. Hopper (2000), “USGS national seismic hazard maps,” Earthquake
Spectra, v. 16, pp. 1–19.
2. Kramer, S.L. (1996) Geotechnical Earthquake Engineering, Prentice-Hall, Inc. Upper
Saddle River, NJ.
3. Borcherdt, R.D. (2002) “Empirical evidence for site coefficients in building code
provisions,” Earthquake Spectra, Vol. 18, pp. 189–217.
4. Hartzell, S. (1998) “Variability in nonlinear sediment response during the 1994
Northridge, California, earthquake,” Bull. Seism. Soc. Am. Vol. 88, 1426–1437.
5. O’Connell, D.R.H. (1999) “Influence of random-correlated crustal velocity fluctua-
tions on the scaling and dispersion of near-source peak ground motions, Science, Vol.
283, pp. 2045–2050, March 26.
6. Yoshida, N. and S. Iai (1998) “Nonlinear site response and its evaluation and predic-
tion,” The Effects of Surface Geology on Seismic Motion, (Irikura, Kudo, Okada and
Sasatani, eds.), Balkema, Rotterdam, pp. 71–90.
7. Huang, N.E., S. Zheng, S.R. Long, M.C. Wu, H,H. Shih, Q. Zheng, N-C. Yen, C.C.
Tung, and M.H. Liu (1998). “The empirical mode decomposition and Hilbert spec-
trum for nonlinear and nonstationary time series analysis." Proc. R. Soc Lond., A 454
903–995.
8. Hardin, B.O. and V.P. Drnevich (1972a) “Shear modulus and damping in soils:
measurement and parameter effects,” J. Soil. Mech. Foundations Div. ASCE, 98,
603–624.
9. Hardin, B.O. and V.P. Drnevich (1972b) “Shear modulus and damping in soils: design
equation and curves measurement and parameter effects,” J. Soil. Mech. Foundations
Div. ASCE, 98, 667–692.
10. Erdik, M. (1987) Site response analysis. In Strong Ground Motion Seismology (Erdik,
M. and Toksoz, M., eds.), D. Reidel Publishing Company, Dordrecht, 479–543.
11. Vucetic, M. and R. Dobry (1991) “Effect of soil plasticity on cyclic response, J.
Geotech. Eng., 117, 89–107.
12. Silva, W., S. Li, B. Darragh, and N. Gregor (1999). “Surface geology based motion
amplification factors for the San Francisco Bay and Los Angeles areas,” A Pearl
Report.
13. Field, E.H., P.A. Johnson, I.A. Beresnev, and Y. Zeng (1997) “Nonlinear
ground-motion amplification by sediments during the 1994 Northridge earthquake,
Nature, 390, 599–602.
14. Beresnev, I.A., E.H. Field, P.A. Johnson, and K.E.A. Van Den Abeele (1998) “Mag-
nitude of nonlinear sediment response in Los Angeles basin during the 1994
Northridge, California, earthquake, Bull. Seis. Soc. Am., Vol. 88, pp. 1097–1084.
15. Joyner, W.B. (1999). “Equivalent-linear ground-response calculations with fre-
quency-dependent damping,” Proceedings of the 31st Joint Meeting of the US-Japan
Panel on Wind and Seismic Effects (UJNR), Tsukuba, Japan, May 11–14, 1999, pp.
258–264.
16. Bonilla, L.F., D. Lavallee, and R.J. Archuleta (1998). “Nonlinear site response:
laboratory modeling as a constraint for modeling accelerograms,” The Effects of
Surface Geology on Seismic Motion, (Irikura, Kudo, Okada and Sasatani, eds.),
Balkema, Rotterdam.
17. Frankel, A.D., D.L. Carver, and R.A., Williams (2002) “Nonlinear and linear site
response and basin effects in Seattle for the M6.8 Nisqually, Washington, Earthquake,”
Bull. Seism. Soc. Am., v. 92, pp. 2090–2109.
18. Safak, E. (1997). “Models and methods to characterize site amplification from a pair
of records,” Earthquake Spectra, Vol. 13, pp. 97–129.
19. Priestley, M.B. (1981). Spectral Analysis and Time Series, Vol. 2, Multivariate Series,
Prediction and Control, Academic Press, London, 1981.
20. Zhang, R., S. Ma, E. Safak, and S. Hartzell (2003b) “Hilbert-Huang transform
analysis of dynamic and earthquake motion recordings,” ASCE Journal of Engineer-
ing Mechanics, Vol. 129, No. 8, pp 861–875.
21. Huang, N.E., Z. Shen, and R.S. Long, (1999). “A new view of nonlinear water waves:
Hilbert Spectrum.” Annu. Rev. Fluid Mech. 31,417–457.
22. Worden, K. and Tomlinson, G.R. (2001) Nonlinearity in Structural Dynamics: Detec-
tion, Identification and Modeling, Institute of Physics Publishing, Bristol and Phila-
delphia.
23. Zhang, R., L. VanDemark, J. Liang, and Y.X. Hu (2003) “Detecting and quantifying
nonlinear soil amplification from earthquake recordings,” Advances in Stochastic
Structural Dynamics (Zhu, Cai and Zhang, eds.), pp. 543–550.
24. Clough, R.W. and J. Penzien (1993). Dynamics of Structures, Second Edition,
McGraw-Hill, Inc., New York.
25. Zhang, R., S. Ma, and S. Hartzell (2003) “Signatures of the seismic source in
EMD-based characterization of the 1994 Northridge, California, earthquake record-
ings,” Bulletin of the Seismological Society of America, Vol. 93, No. 1, pp 501–518.
26. Salvino, L.W. (2001) “Evaluation of structural response and damping using empirical
mode analysis and HHT,” in CD ROM, 5th World Multiconference on Systemics,
Cybernetics and Informatics, July 22–25, 2001, Orlando, Florida.
27. Loh, C.H., T.C. Wu and N.E. Huang (2001). Application of the empirical mode
decomposition-Hilbert spectrum method to identify near-fault ground-motion char-
acteristics and structural responses, Bull. Seism. Soc. Am. 91, 1339–1357.
28. Yang, J.N., Y. Lei, S. Pan and N. Huang (2002a). “System identification of linear
structures based on Hilbert-Huang spectral analysis. Part I: Normal modes,” Earth-
quake Eng. Structural Dynamics, Vol. 32, pp. 1443–1467.
29. Yang, J.N., Y. Lei, S. Pan and N. Huang (2002b). “System identification of linear
structures based on Hilbert-Huang spectral analysis. Part II: Complex modes,” Earth-
quake Eng. Structural Dynamics, Vol. 32, pp. 1533–1554.
30. Xu, Y.L., S.W. Chen, and R. Zhang (2003) “Modal identification of Di Wang building
under typhoon York using HHT method,” The Structural Design of Tall and Special
Buildings, Vol. 12, No. 1, pp. 21–47.
31. Hilbert-Huang Transformation Toolbox (2000). Professional Edition V1.0, Princeton
Satellite Systems, Inc., Princeton, New Jersey.
32. Hou, Z., M. Noori and R. St. Amand (2000) “Wavelet-based approach for structural
damage detection,” ASCE Journal of Engineering Mechanics, 126(7), pp. 677–683.
33. Archuleta, R.J. and J.H. Steidl (1998). “ESG studies in the United States: results
from borehole arrays,” The Effects of Surface Geology on Seismic Motion, (Irikura,
Kudo, Okada and Sasatani, eds.)., Balkema, Rotterdam, pp. 3–14.
34. Hartzell, S., A. Leeds, A. Frankel, R.A. Williams, J. Odum, W. Stephenson and W.
Silva (2002). “Simulation of broadband ground motion including nonlinear soil
effects for a magnitude 6.5 earthquake on the Seattle fault, Seattle, Washington,” Bull.
Seism. Soc. Am. Vol. 92, No. 2, pp. 831–853.
9 Simulation of
Nonstationary Random
Processes Using
Instantaneous Frequency
and Amplitude from
Hilbert-Huang Transform
Ping Gu and Y. Kwei Wen
CONTENTS
9.1 INTRODUCTION
With the ever-increasing power of the computer and of the capacity of computational
mechanics, simulation has become an indispensable tool in engineering. It is more
so in earthquake engineering since earthquake ground motions are highly random and
structural behaviors are nonlinear and inelastic, making random vibration solutions
191
difficult to obtain. Generally speaking, structural responses are functions of the entire
ground excitation time history. Current codes use a scalar intensity measure such as
spectral acceleration or displacement for the earthquake demand on structures. These
measures do not reflect satisfactorily the effects of near source or effects due to
higher modes and other important structural response behaviors caused by ground
motions. For performance evaluation of complex structural systems, simulation of
earthquake ground motions and time history response analysis have played a more
and more important role.
Many models for simulation based on earthquake records have been proposed.
For example, a stationary white noise model was first proposed [1], followed by
filtered white noise models to reflect the site characteristics [2, 3, 4]. A high-pass
filter was later added to remove the undesirable zero frequency content of the power
spectral density function [5]. To account for the intensity variation with time, a
uniformly (amplitude) modulated random process was first used [6, 7]. Nonstationary
models with frequency content variation with time — in particular, the long-duration
acceleration pulses observed in many near-fault earthquake records — have also
been developed.
Saragoni and Hart [8] proposed a modulated filtered Gaussian white noise model
with different intensity and frequency content in three consecutive time intervals.
The arbitrary selection of time intervals and the resultant abrupt change of frequency
content in this model, however, are difficult to justify physically. The oscillatory
process described by an evolutionary power spectral density suggested by Priestley
[9], allowing both amplitude and frequency content variation with time, has been
widely used in earthquake ground motion simulation. For example, Lin and Yong
[10] formulated an evolutionary Kanai-Tajimi [2, 3] model as a filtered pulse train,
where the filter represents the ground medium and the pulse train represents inter-
mittent ruptures at the earthquake source.
Deodatis and Shinozuka [11] simulated the earthquake ground motions using
an autoregressive (AR) model based on selected parametric forms of the evolutionary
power spectral density for given earthquake records. Li and Kareem [12, 13] further
extended the method using the fast Fourier transform (FFT) and digital filtering
technique to make it more computationally efficient. Der Kiureghian and Crempien
[14] proposed an evolutionary earthquake model composed of individually modu-
lated band-limited white noise processes and estimated the parameters of evolution-
ary power spectral density by matching the moments of each component process.
Papadimitriou [15] produced a parsimonious nonstationary model by applying a
second-order filter with slowly varying parameters to time-modulated white noise.
Conte and Peng [16] proposed to use Thomson’s spectrum to estimate the evolu-
tionary power spectra and proposed a model for simulation based on this. By
extending the energy principle of wavelet transform developed by Iyama and Kuwa-
mura [17], Spanos and Failla [18] and Spanos et al. [19] proposed methods using
the wavelet spectrum to estimate the evolutionary power spectral density.
The main challenge in evolutionary process-based methods lies in parameter
estimation. In addition to estimating the power spectral density function, one also
needs to know the modulation functions of frequency and time, which are not known
a priori and are difficult to estimate. Motivated by an amplitude- and frequency-mod-
ulated process proposed by Grigoriu et al. [20], Yeh and Wen [21] proposed to model
ground motion as I(t)ς(φ(t)), where I(t) is a deterministic intensity envelope that
controls the amplitude of the ground motion, ς(φ(t)) is a stationary filtered white
noise in time scale φ and depicts the spectral density form of the process, and φ(t)
is a smooth, strictly increasing function that depicts the frequency modulation. The
stationary white noise was filtered to have a Clough-Penzien spectrum [5] with zero
mean and unit variance. By changing the time scale of the stationary filtered white
noise, the spectral density of the stationary process becomes time varying:
1 ω
Sςς (t , ω ) = SCP . (9.1)
φ′(t ) φ′(t )
This method does not require the “slowly time-varying” assumption implied in
the evolutionary spectrum method, and it gives a time-dependent power spectral
density as a simple modification of the spectral density of the stationary process.
However, the frequency modulation function is estimated by a zero-crossing rate
procedure and is independent of the amplitude variation. It is therefore somewhat
restrictive in nature and only suited for a certain special class of random processes.
It has been applied to earthquake ground motion simulation with some success but
lacks general theoretical basis for wider applications.
Recently, Huang et al. [22] introduced a new method of spectral analysis of
nonstationary time series based on empirical modal decomposition (EMD), Hilbert
transform, and the concept of instantaneous frequency, commonly referred to as the
method of Hilbert-Huang transform (HHT). Unlike the Fourier transform, the Hilbert
transform emphasizes local behavior and does not decompose the signal into sine
and cosine waves of fixed frequencies; thus it is suitable for nonstationary processes.
HHT decomposes the signal into orthogonal functions called intrinsic mode func-
tions (IMFs) directly based on data. The Hilbert transform of IMFs yields the
instantaneous amplitude and instantaneous frequency. HHT can be used to overcome
the difficulties of the Wen-Yeh simulation method since EMD provides a reasonable
and physically meaningful decomposition of the nonstationary signal. The Hilbert
transform yields instantaneous amplitude and frequency of each IMF, which can be
used directly in the Wen-Yeh method. It bypasses the difficult tasks of estimating
and constructing analytical functions for frequency and amplitude modulation. Also,
since these two modulation functions are obtained from Hilbert transform of IMFs,
the dependence between the two functions is preserved.
In this study, an IMF recombination method for simulation of nonstationary
processes is first proposed, followed by an improved Wen-Yeh method. After a brief
introduction of HHT, we describe the IMF recombination method, followed by the
improved Wen-Yeh Method. We then compare the two. These methods are demon-
strated by numerical examples of simulation of earthquake ground motions.
The extraction of IMFs out of a time series requires a repeated “sifting” proce-
dure called empirical mode decomposition. EMD is described briefly as follows:
Connect all the local maxima by a cubic spline as the upper envelope and connect
all the local minima by a cubic spline as the lower envelope. Then calculate the
mean of the envelopes and subtract the mean from the time series. The result is the
first component. Ideally it should have satisfied the requirements for an IMF, but in
practice a gentle hump on a slope can be amplified to become a local extreme. So
the procedure is repeated until an IMF is obtained. This is the first IMF. Subtract
the first IMF from the original time series, then continue the procedure on the residue
to obtain the second IMF. Repeat the process to obtain the third, fourth, fifth, and
subsequent IMFs, until the residue is too small to be of any consequence or becomes
a monotonic function from which no more IMF can be extracted.
After the EMD, the time series X(t) can be expressed in terms of IMFs as follows:
X (t ) = ∑ C (t) + r (t)
j =1
j n (9.2)
in which Cj(t) is the jth IMF, and rn(t) is the residue that can be the mean trend or
a constant. Taking Hilbert transform of Cj(t),
C j (τ)
∫
1
D j (t ) = P dτ (9.3)
π t−τ
in which P indicates Cauchy principal value. One can then form an analytical
function
iθ j ( t )
Zj(t) = Cj(t) + iDj(t) = aj(t) e . (9.4)
d θ j (t )
ω j (t ) = . (9.5)
dt
∑ a (t)e
X (t ) = Re{
j =1
j
iθ j ( t )
} + rn (t ) (9.6)
in which Re denotes the real part. Therefore, aj(t), j = 1, …, n, therefore control the
amplitude variation; each has an instantaneous frequency at a given time. Collec-
tively, the functions aj(t) and j(t) define the Hilbert amplitude spectrum. Similarly,
aj2(t) and j(t) define the energy distribution with respect to time and frequency, called
the Hilbert energy spectrum or simply Hilbert spectrum, H(,t). The marginal Hilbert
energy spectrum is obtained by integration over the duration:
h(ω ) =
∫ H (ω, t)dt .
0
(9.7)
Because of the local nature of the Hilbert transform and the concept of instan-
taneous frequency, the Hilbert marginal spectrum describes an energy contribution
corresponding to a given frequency at a certain time but not the entire duration of
the process, as would be the case for a Fourier spectrum. This is a fundamental
difference between these two approaches.
As an example, the method is applied to the well-known N-S component of the
El Centro earthquake record of 1940 and the Newhall ground motion of the 1994
Northridge earthquake in SAC-2 ground motion records [23], representing far-source
and near-source records respectively. The time histories of the ground accelerations
(in cm/sec2) are shown at the top of Figure 9.1 (At the bottom of the figure, four
simulated samples of the ground accelerations are shown based on the IMF recom-
bination method described later in this paper.) The IMFs of the Newhall record are
shown in Figure 9.2. The Hilbert spectra are calculated and shown in Figure 9.3.
As can be seen, one distinct advantage of the Hilbert spectrum is that the energy
distribution over time and frequency is always sharp. For example, a darker shade
of color in the plot indicates higher energy, which can be clearly seen in the Newhall
record from 4 to 10 seconds and below 20 rad/sec (3 cps), a frequently observed
feature of near-source records; it can also be seen over a longer time period in the
El Centro record. Figure 9.4 compares the marginal Hilbert energy spectrum with
the Fourier energy spectrum obtained from the same record. There are some obvious
differences due to the two different methods of spectral description. The scale of
the Hilbert marginal spectra has been adjusted by a factor of π for better comparison.
Details of comparison of these two spectra and the conversion factor are shown in
the Appendix. Note that the Hilbert transform can be carried out by first performing
the Fourier transform on the original signal, changing the phase of each positive
frequency component by /2 and the phase of each negative frequency component
by –/2, and finally performing an inverse Fourier transform to obtain the Hilbert
transform of the sample [24]. One can use FFT to make the procedure more efficient.
FIGURE 9.1 Recorded (top) and simulated El Centro and Newhall ground motions using
IMF recombination method.
FIGURE 9.2 IMFs of the 1994 Northridge Newhall record (SAC LA14).
n
x (t ) = Re
∑ a (t)e
j =1
j
iθ j ( t )
+ rn (t )
n
X (t ) = Re ∑ a j (t )e
i [ θ j ( t )+ φ j ]
+ rn (t ) , (9.8)
j =1
140
120
Frequency (rad/s)
100
80
60
40
20
0 5 10 15 20 25
Time (s)
140
120
Frequency (rad/s)
100
80
60
40
20
0 5 10 15 20 25 30
Time (s)
FIGURE 9.3 Hilbert spectra of 1940 El Centro and 1994 Northridge Newhall ground motion.
n
µ X (t ) Re
∑
j =1
a j (t )e
iθ j ( t ) iφ
E (e j ) + rn (t ) = rn (t )
(9.9)
Hilbert Marginal Spectrum and Fourier Energy Spectrum for Newhall Record
106
105
104
103
2
10
101
100
10−1
10−2
Fourier Energy Spectrum
−3
Hilbert Marginal Spectrum (adjusted)
10
10−2 10−1 100 101 102 103
rad/s
Hilbert Marginal Spectrum and Fourier Energy Spectrum for EI Centro Record
10−1
10−2
10−3
10−4
10−5
10−6
FIGURE 9.4 Comparison of Hilbert marginal spectrum (solid line) with Fourier energy
spectrum (dashed line).
K XX (t1 , t2 ) =
1
2 ∑ a (t )a (t ) cos[θ (t ) − θ (t )]
j =1
j 1 j 2 j 1 j 2 (9.10)
{ }
n
() () 1
∑ a (t )
2
σ 2x (t) = E X t − uX t = 2
j (9.11)
2 j =1
X(t), as given in Equation 9.8, has a Hilbert energy spectrum characterized by aj2(t)
with instantaneous frequency d(j(t))/dt, for j = 1, …, n. Due to the Central Limit
Theorem, X(t) approaches a Gaussian process for large n. This Hilbert spectral
representation model suggests a simple method for simulation of a nonstationary
process as shown in Equation 9.8. One can generate the random phase angles on
computer and use Equation 9.8 to simulate the process. The name reflects the fact
that it is essentially a method of recombination of the IMFs. The Hilbert spectrum
of each simulated sample can be shown to be the same as that of the recorded ground
motion.
The physical basis for this method is that the frequency and amplitude of each
IMF represent those of a sub-wave caused by a given physical mechanism of the
rupture surface of the earthquake, as shown by Zhang [25]. For comparison with
the traditional Fourier-based source inversion solution obtained by Hartzell et al.
[34] for the Northridge earthquake based on 70 records, Zhang studied the same 70
earthquake records using HHT. The spatial distributions of the slip amplitude over
the finite fault, each of which corresponds respectively to the second to fifth IMFs
of those records, were plotted. It was found that the large slip amplitude regions are
located at almost the same regions as those in Hartzell’s study, indicating that wave
motions in each IMF are generated by a fraction of the source. In addition, the
rupture with large-slip region scatters out to the northwest from the hypocenter, with
the IMF changing from the second to the fifth, indicating that the seismic waves are
generated sequentially from the dominant short-period signals to main long-period
signals as the rupture propagates. Therefore, the IMFs are closely related to the
source mechanism in general and the source heterogeneity and rupture process in
particular. For future earthquakes of similar characteristics, the rupture surface spatial
features are random and unpredictable. The IMF recombination method represents
an efficient way to capture these variabilities.
Simulations of the N-S component of the El Centro and the Northridge Newhall
records (SAC LA14) were carried out, and sample time histories are shown in Figure
9.1. The simulated samples represent a reasonably statistical image of the intensity
and frequency variation with time of the records. The method has also been extended
to vector processes.
are reflected well in the simulated samples. The response spectra generally fluctuate
around that of the record. Unlike samples generated by methods based on analytical
models of the spectra, the response spectra of the simulated samples retain the jagged
look of those of the records. The artificial smoothness of the response spectra of the
simulated ground motions obtained by existing power spectrum-based methods has
been criticized by seismologists and practicing engineers.
To further illustrate the procedure, the steps leading to the first sample in Figure
9.5, shown below by figures, are as follows:
1.5
0.5
0
0 1 2 3 4 5 6 7 8 9 10
Period (sec)
FIGURE 9.6 Response spectra of Newhall LA14 simulations by improved Wen-Yeh method.
5. Restore the time scale and the amplitude modulation to obtain the simu-
lated jth IMF Sj(t). The restored sample in Step 4 of the second IMF is
shown in Figure 9.12 and compared with the original second IMF.
6. Add all Sj(t), j =1 to n. The result is Sample 1 in Figure 9.5.
9.5 CONCLUSIONS
Two new methods of simulation of nonstationary random processes are presented
based on the Hilbert-Huang transform of sample observations, namely the IMF recom-
bination method and the improved Wen-Yeh method. Both take advantage of EMD and
the instantaneous frequency and amplitude of the Hilbert transform, thus overcoming
difficulties in the estimation of the frequency modulation and interdependence of
400
350
300
250
200
150
100
50
0
0 5 10 15 20 25 30
FIGURE 9.7 Original (dashed line) and smoothed (solid line) instantaneous amplitude func-
tion of the second IMF.
100
80
60
40
20
0
0 5 10 15 20 25 30
FIGURE 9.8 Original (dashed line) and smoothed (solid line) instantaneous frequency func-
tion of the second IMF.
0.5
−0.5
−1
−1.5
−2
0 50 100 150 200 250 300 350
1.5
0.5
−0.5
−1
−1.5
3 4 5 6 7 8 9 10
1.5
0.5
−0.5
−1
−1.5
40 50 60 70 80 90 100 110 120 130
FIGURE 9.10 Effect of change in time scale to remove frequency modulation: original (top);
frequency modulation removed (bottom).
1.5
0.5
−0.5
−1
−1.5
−2
0 50 100 150 200 250 300 350
400
300
200
100
−100
−200
−300
−400
−500
0 5 10 15 20 25 30
FIGURE 9.12 Comparison of original (dotted line) and simulated (solid line) second IMF.
ACKNOWLEDGMENTS
This research is supported by the University of Illinois Research Board and the NSF
Earthquake Engineering Research Center Program under Award Number
EEC-9701785. Information exchange with C. H. Loh of National Taiwan University
is greatly appreciated.
REFERENCES
1. Bycroft, G. N. (1960). White noise representation of earthquakes. J. Eng. Mech. Div.,
ASCE, Vol. 86, EM2: 1–16.
2. Kanai, K. (1957). Semi-empirical formula for the seismic characteristics of the
ground. Bull. Earthquake Res. Inst., Univ. Tokyo, Vol. 35, 309–325.
3. Tajimi, H. (1960). A statistical method of determining the maximum responses of a
building structure during an earthquake. Proc. Second World Conf. Earthquake Eng.,
Tokyo and Kyoto, Japan, Vol. II.
4. Housner, G. W. and Jenning, P. C. (1964). Generation of Artificial Earthquakes, J.
Eng. Mech. Div., ASCE, Vol. 90, EM1: 113–150.
5. Clough, R. W. and Penzien, J. (1975). Dynamics of Structures. McGraw-Hill Book
Co., New York.
6. Shinozuka, M. and Sato, Y. (1967). Simulation of nonstationary random processes.
J. Eng. Mech. Div., ASCE, Vol. 93, EM1: 11–40.
7. Amin, M. and Ang, A. H.-S. (1968). Nonstationary stochastic model for earthquake
motions. J. Eng. Mech. Div., ASCE, Vol. 94, EM2: 559–583.
8. Saragoni, G. R. and Hart, G. C. (1974). Simulation of artificial earthquakes. Earth-
quake Eng. Structural Dynamics, Vol. 2, 249–267.
9. Priestley, M. B. (1967). Power spectral analysis of non-stationary random processes.
J. Sound Vibration, Vol. 6, 1: 86–97.
10. Lin Y. K. and Yong, Y. (1987). Evolutionary Kanai-Tajimi earthquake models. J. Eng.
Mech. Div., ASCE, Vol. 113, 8: 1119–1137.
11. Deodatis, G. and Shinozuka, M. (1988). Auto-regressive model for nonstationary
stochastic processes. J. Eng. Mech., ASCE, Vol. 114, 4: 1995–2012.
31. Zhang R. R. (2001). The signature of structural damage: An HHT view. In: Structural
Safety and Reliability, Corotis et al. (eds.), Swets & Zeitlinger, ISBN 90 5809 197 X.
32. Salvino, L. W. (2000). Empirical mode analysis of structural response and damping.
Proc. SPIE Int. Soc. Optical Eng., Vol. 4062, 503–509.
33. Gravier, B. M. et al. (2001). An assessment of the application of the Hilbert spectrum
to the fatigue analysis of marine risers. Proc. Int. Offshore Polar Eng. Conf., Vol. 2,
268–275.
34. Hartzell, S., Liu, P., and Mendoza, C. (1996). The 1994 Northridge, California
earthquake: Investigation of rupture velocity, risetime, and high-frequency radiation.
J. Geophys. Res., Vol. 101, 20091–20108.
35. Shinozuka, M. and Deodatis, G. (1991). Simulation of stochastic processes by spectral
representation. App. Mech. Reviews, Vol. 44, 4: 191–203
∞
y(ω ) =
∫−∞
y(t )e − iωt dt (9A.1)
∞
y∗ (ω ) =
∫ −∞
y(t )eiωt dt (9A.2)
2
E f (ω ) = y(ω ) (9A.3)
then the area under the Fourier energy spectrum for all frequency is given by
∞ ∞ ∞ ∞
E=
∫ y(ω) y (ω)dω = ∫ ∫ ∫ y(t ) y(t )e
−∞
∗
−∞ −∞ −∞
1 2
− iωt1 + iωt2
d ωdt1dt2
∞ ∞
=
∫ ∫ y(t ) y(t )2πδ(t − t )dt dt
−∞ −∞
1 2 1 2 1 2
∞ T
= 2π
∫ y (t)dt = 2π ∫ y (t)dt
−∞
2
0
2
(9A.4)
if y(t) is defined for 0 < t < T, and zero otherwise. Therefore, the area under the
Fourier energy spectrum defined by Equation 9A.3 is equal to 2π times the total
energy of the time series, defined by the integral of the square of the time series.
The area under the Hilbert energy spectrum can be illustrated by a simple case
of a time series that is a summation of n sine functions defined for a given time
period 0 to T,
∑ A T.
j =1
2
j
∫ ∑ Aj
T
1
y (t )dt = T
2 2
(9A.6)
0 2 j =1
∑ ∫
T
A2j T = 2 y 2 (t )dt (9A.7)
0
j =1
if we neglect the residual term from Equation 9.5. Its total area under the Hilbert
marginal spectrum is
n T
∑ ∫ a (t)dt.
j =1 0
2
j
n n
∫ ∑∑ ∫
T T
y 2 (t )dt = ai (t )a j (t ) cos[θi (t )]cos[θ j (t )]dt (9A.9)
0 0
i =1 j =1
Due to the orthogonality of IMFs (see Huang et al. [22]), the cross-terms (i ≠ j) in
Equation 9A.9 can be neglected. Equation 9A.9 reduces to
∫ ∑∫
T T
y 2 (t )dt = a 2j (t ) cos2 [θ j (t )]dt (9A.10)
0 0
j =1
When the IMFs have relatively smooth and sinusoidal shapes, for integer number
of quarter-waves, we have
∫ ∫
T T
1
a 2j (t ) cos2 [θ j (t )]dt ≈ a 2j (t )dt (9A.11)
0 2 0
∫ ∑∫
T T
1
y 2 (t )dt ≈ a 2j (t )dt (9A.12)
0 2 j =1
0
∑∫ ∫
T T
a 2j (t )dt ≈ 2 y 2 (t )dt (9A.13)
0 0
j =1
Comparing Equation 9A.13 with Equation 9A.4, we conclude that the total area
under the Hilbert marginal spectrum needs to be multiplied by a factor, to be
comparable with the area under the Fourier energy spectrum defined by Equation
9A.3.
10 Comparison of
Hilbert-Huang, Wavelet,
and Fourier Transforms
for Selected Applications
Ser-Tong Quek, Puat-Siong Tua, and Quan Wang
CONTENTS
ABSTRACT
213
WT give similar results in terms of precision and accuracy. Last, a modified Fourier
transform (FT) is investigated based on the concept of HHT and applied to the signals
collected from the free vibration of an aluminum beam under three different support
conditions. The proposed method is able to reveal the presence of higher vibration
modes than can be revealed by conventional FT of the original signal. The results
produced are close to that of the more common WT. It is believed that supporters of
FT will be more receptive to this modified HHT method.
10.1 INTRODUCTION
Signal processing has gradually become an indispensable tool in the field of engi-
neering for extraction of important information from raw signals. In civil engineer-
ing, many practical and robust nondestructive evaluation (NDE) techniques devel-
oped for assessment of structural performance involve two key components, namely:
(1) data acquisition and (2) signal processing and interpretation. Although the use
of vibration measurements is a simpler and less costly method with respect to
instrumentation system in comparison with infrared thermography, ground-penetrat-
ing radar, acoustic emission monitoring, and eddy current detection, a key factor for
identifying damage precisely lies in having an appropriate data analysis method.
The most well known conventional signal processing technique is the Fourier
transform (FT). Fourier spectral analysis provides a general method for examining
the global energy–frequency distributions of a given signal. FT has high resolution
in the frequency domain but loses the resolution in the time domain. Despite this,
FT has been employed in health monitoring. For example, Crema et al. [1] used a
fast Fourier transform (FFT) analyzer to perform modal analysis to detect damage
in composite material structures. A broadband impulse input was applied, and the
damage that was induced by loads well above the working load caused a small
decrease in the eigen-frequencies for several modes. However, it can be problematic
to locate damage by using FT-based modal analysis techniques.
A few methods are available for processing signals of nonlinear and nonstation-
ary systems, of which the short-time Fourier transform (STFT) method and wavelet
transform (WT) method are widely adopted. Cook and Berthelot [2] used the
time-dependent energy spectrum from STFT to distinguish backscattered signals
from a single crack against those from a series of closely spaced cracks on steel
surface. Hurlebaus et al. [3] used the time–frequency spectrum obtained by perform-
ing STFT on the Lamb waves generated by laser to obtain the group velocity–fre-
quency domain. The notches are located from the autocorrelation plot of a series of
group velocity spectra at different assumed propagation distances.
Wavelet analysis is another adjustable window signal-processing technique to
handle nonstationary signals. Although WT appears similar to STFT, the basic wave
components are not limited to sinusoidal functions. Rather, classes of functions have
been proposed, from the simplest Haar wavelet to the more complicated higher order
Debauchies wavelet functions, to address various problems of time–frequency
resolution. WT has been widely used in the field of damage detection, such as in
modal-based detection (Staszewski [4]). An early work using WT to obtain a dis-
persive relation of group velocity of flexural waves in beam was presented by
Kishimoto et al. [5]. Lee and Liew [6] studied the sensitivity of WT in space domain
for identifying the damage location of a beam for small damage not easily detected
via conventional eigenvalue analysis. Quek et al. [7], on the other hand, used wavelet
coefficients over the time domain to locate cracks in beams via the time-of-flight
analysis of the wave propagation along the beam. However, WT has its shortcomings.
Experience shows that it can produce many spurious harmonics under different scales
that makes the analysis difficult or sometimes meaningless.
The most recently introduced signal-processing technique for analyzing nonlin-
ear and nonstationary time series data is the Hilbert-Huang transform (HHT) [8].
Empirical mode decomposition (EMD) of the data is first performed to segregate
the signals into narrow band components, and the Hilbert transform (HT) is then
applied on each mode. Its ability to analyze nonlinear and nonstationary time series
data [8–10] has attracted many applications. For example, Zhang et al. [11] adopted
HHT to assess structural damage from vibration recordings. Shim et al. [12]
exploited HHT’s ability to exhibit nonlinear and nonstationary behavior to locate
damages on bridges by using frequency sweep and controlled excitations. Quek et
al. [13] also employed HHT to analyze structural dynamic responses to detect and
locate anomalies in beams and plates. In addition, HHT has been used to analyze
experimental data collected by oscilloscope to investigate the dominance of the
lowest antisymmetric (A0) and symmetric (S0) modes of Lamb wave vary across the
frequency range adopted for excitation and good agreement with, which agrees well
with theoretical response prediction in terms of their relative dominance [14].
Individual signal-processing techniques have their own strength and weaknesses.
Suitability of a particular technique may, therefore, be problem dependent, especially
with respect to structural health monitoring problems. This paper compares the three
signal-processing techniques, namely, HHT, WT, and FFT, for signals obtained
experimentally. The first example concerns the time-of-flight analysis of flexural
wave propagation in an aluminum beam, where WT and HHT are compared. Next,
we compare the analysis from the two techniques on a two-dimensional problem
involving the propagation of acoustic Lamb waves in an aluminum plate. Last, we
apply a modified FT based on the EMD concept of HHT to the free vibration signals
of an aluminum beam under three different support conditions, we compare these
results with results from using conventional direct FFT, HHT, and WT to obtain the
modal frequencies.
The distinct features of a narrow-band signal, and hence an IMF, are that (a)
over its entire length, the number of extrema and the number of zero-crossings either
are equal or differ at most by one; and (b) at any point, the mean value of the
envelope of the signal defined by the local maxima and the envelope defined by the
local minima is zero. This forms the basis on which the EMD technique was
developed. For the EMD to be applicable, it is assumed that (a) the signal has at
least one maximum and one minimum; (b) the characteristic time scale is defined
by the time lapse between the extrema; and (c) if the data are totally devoid of
extrema but contain only inflection points, then the signal can be differentiated one
or more times to reveal the extrema. Hilbert spectral analysis can then be performed
on each IMF. MATLAB 6.1 is adopted for the implementation of the HHT in this
study.
TABLE 10.1
Geometrical and Material Properties
of Aluminum Beam/Plate
Beam Plate
FIGURE 10.1 Schematic view of wave propagation for impact at 200 mm position.
of data with different boundary conditions are analyzed, namely, fixed-fixed, simply
supported, and cantilever conditions. The basic concept is shown in Figure 10.1, in
a schematic view of wave propagation due to an impact. The piezoelectric sensor
captures the wave signal, which contains timings of the direct, damage-reflected,
and boundary-reflected waves. By estimating these timings with a signal-processing
technique, the damage location can be deduced. In view of the dispersive nature of
the wave, only a wave at a particular frequency, with a corresponding wave propa-
gation velocity, is used so that accurate results can be obtained.
HHT is applied in this study on the same three sets of signals. One of the signals
obtained (and shown in Figure 10.2a) was decomposed via EMD into its IMF
components, shown in Figure 10.2b. It can be observed that the first IMF component
has higher amplitude than the other IMFs. This IMF component is then subjected
to HT, and the corresponding frequency–time and energy–time distributions plotted
in Figure 10.2c and d, respectively. From the energy–time distribution, the timings
of the energy peaks in the signal, which are due to the propagating wave passing
the sensor, are obtained. The frequencies corresponding to the energy peaks are also
noted so as to establish that they do not deviate significantly from each other. It
should be pointed out that not all IMFs are used, as the first IMF is distinct from
the others and contains only the frequency of interest, whereas other IMFs do not.
The results for the estimated damage position are summarized in Table 10.2 and
compared with those obtained with Gabor wavelet analysis with coarse (discrete
WT, scale = 12) and fine (“continuous” WT or CWT, scale = 12.2) resolution. The
wavelet results have been discussed in detail by Quek et al. [7], including the 3D
wavelet plot, and will not be repeated here. It can be observed that the HHT method
is able to identify the damage position with reasonable accuracy; the results are
better than the discrete wavelet results but worse than the continuous wavelet results.
However, using CWT for this application requires determining an appropriate scale
to process the final results. From past experience, this requires a fair amount of skill,
as the selection can be rather subjective and may lead to varied end results (see, for
10 (b) 10
8 (a)
c(1)
6 0
4
2
Input
0 −10
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
−2
−4
−6
−8 5
−10
c(2)
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5 0
−3
Time, t(s) ×10
−5
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
5
5000
(c)
Frequency (Hz)
c(3)
4000 0
3000
−5
2000 0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
1000
0 5
0 1 2 3 4 5
−3
Time, t(s) ×10
c(4)
0
−5
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
5
(d)
4 1
Amplitude
3
Residue
0
2
1 −1
0 −2
0 1 2 3 4 5 0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
−3 −3
Time, t(s) ×10 Time, t(s) ×10
FIGURE 10.2 (a) Dynamic response of a beam; (b) IMFs of response signal after EMD; (c)
frequency; (d) energy spectra of first IMF component.
TABLE 10.2
Comparison of Results Based on HHT Method and Wavelet Analysis
Wavelet Analysis
HHT Analysis Scale 12 Scale 12.2
Estimated Estimated Estimated
Damage Error Damage Error Damage *Error
End Condition Position (%) Position (%) Position (%)
D C
600
500
Sensor
400 Actuator
50 mm
300
75 mm
y
100
0 X
A 0 100 200 300 400 500 600 B
example, Table 3 of Quek et al. [7]). This is due to the shortcomings of WT, which
produces false harmonics depending on the scale adopted, although WT also has
the ability to achieve high resolution in both frequency and time domain. The
advantage of HHT over WT in this case is that the HHT procedure is more direct
and can be performed to obtain the desired accuracy for the results without prior
knowledge of the excitation frequency (which in this case of impact load is broad-
band). It should be noted that there are other applications where CWT may be more
appropriate.
( )
X (t ) = cos 2πω t e − a (t −t0 )
2
(10.1)
20
15
10
Amplitude (V)
5
−5
−10
−15
−20
1.5 2 2.5 3 3.5 4
−5
Time (s) ×10
FIGURE 10.4 Actuation pulse using Equation 10.1 at frequency 600 kHz.
where ω is the actuation frequency (Hz), and a and t0 are constants. This was done
with a function generator (Yokogawa FG300 15 MHz Synthesized FG) to feed a
PZT actuator (8.0 mm × 8.0 mm × 0.5 mm). The responses were captured by a PZT
sensor placed at a distance away and channeled to an oscilloscope (Yokogawa DL716
16-Channel) for monitoring and analysis. Figure 10.3 shows the positions of the
actuator and sensor. By estimating the flight time due to the propagation of either
a S0 or an A0 Lamb wave, the distance traveled by the wave from the actuator to
the sensor can be computed as
li = ∆ ti cg, (10.2)
where ti is the time of flight referenced from the actuation time, and cg is the group
velocity of either the S0 or A0 mode, determined either theoretically or experimen-
tally.
The response signal captured by the PZT sensor (Figure 10.5a) was processed
by both HHT and Gabor WT. The resulting energy spectra are given in Figure 10.5d
and Figure 10.6. In the HHT analysis, the peaks in the energy–time spectrum for
the IMF component containing the highest energy (in this case the first IMF) give
the wave arrival times of interest. Based on these energy peaks, the distances traveled
by the wave from the actuator to the sensor can be computed; these are summarized
in Table 10.3. It can also be observed from the frequency spectrum (see Figure
10.5b) that the frequency corresponding to the energy peaks is very close to the
actuation frequency, which indicates the accurate representation of localization
events by HHT. On the other hand, Gabor WT analysis on the same signal response
was performed based on the prior knowledge that the actuation frequency is 600
kHz, which corresponds to a scale of 20.2. Similar energy peaks were obtained,
which accounted for the wave pulses on the different paths received by the sensor.
The results obtained by WT analysis are also given in Table 10.3. These indicate
1 (b)
0.8 (a) 1
c(1)
0.6 0
0.4 −1
0 1 2
Amplitude
0.2 0.05
c(2)
0 0
−0.2 −0.05
−0.4 0 1 2
0.02
−0.6
c(3)
0
−0.8 −0.02
−1 0 1 2
0 1 2 0.02
c(4)
−4 0
Time, t(s) ×10
−0.02
0 1 2
0.01
c(5)
0
5 −0.01
×10 0 1 2
15 0.01
c(6)
(c) 0
Frequency, (Hz)
10 −0.01
0 ×10−3 1 2
5
c(7)
5 0
−5
0 0 1 2
0.01
c(8)
−5 0
0 1 2 −0.01
−4 0 1 2
Time, t(s) ×10 0.2
c(9)
0
−0.2
0 1 2
0.02
c(10)
1 A0/S0 0
(d) −0.02
0.8 (incident) 0 1 2
Amplitude
A0 0.1
c(11)
0.6 0
(crack) A0
0.4 (boundary, CD) −0.1
0 1 2
0.2 0.1
Res
0
0 −0.1
0 1 2 0 1 2
−4 −4
Time, t(s) ×10 Time, t(s) ×10
FIGURE 10.5 (a) Response signal captured by sensor on aluminum plate; (b) IMFs of
response signal after EMD; (c) frequency; (d) energy spectra of first IMF component.
1.2E–03
A0/S0
(incident)
1.0E–03
8.0E–04
Amplitude
A0
6.0E–04 (crack)
A0
4.0E–04 (boundary, CD)
2.0E–04
0.0E+00
0.0E+00 5.0E–05 1.0E–04 1.5E–04 2.0E–04
Time (s)
FIGURE 10.6 Gabor wavelet analysis of response signal given in Figure 10.5a.
TABLE 10.3
Results of Wave Propagation in Aluminum Plate
Distance Traveled Distance Traveled
Theoretical Distance Traveled by Reflected Wave by Reflected Wave
Velocity by Incident Wave from Line Crack from Boundary
Signal (m/s) (mm)a (mm)b (CD; mm)c
Analysis Error Error Error
Technique A0 A0 (%) A0 (%) A0 (%)
that HHT and WT give similar results in terms of precision and accuracy. HHT does
not require prior knowledge of the activation frequency.
(β nl )2 EI
fn = , n = 1, 2, 3, ... (10.3)
2πl 2 ρA
where l, E, I, ρ, and A are the length, Young’s modulus, moment of inertia, density,
and cross-sectional area of the beam, respectively; n corresponds to the vibration
modes; and βn is solved from the following equations, respectively, for cantilever,
simply supported, and fixed-fixed conditions.
sinβ n l = 0 (10.4b)
For the experiment, the free vibration responses of the beam under the three different
support conditions (Figure 10.7) were monitored and collected by an oscilloscope
(Yokogawa DL716 16-Channel; see Figure 10.8). FFT was first performed directly on
Aluminum
Piezoelectric beam 6 mm
x Sensor
(a)
32 mm
y
920 mm
0 mm 150 mm
Aluminum
Piezoelectric beam 6 mm
x Sensor
(b) 32 mm
y
800 mm
0 mm 150 mm
Aluminum
Piezoelectric beam 6 mm
x Sensor
(c)
32 mm
y
800 mm
0 mm 150 mm
FIGURE 10.7 Schematic setup of aluminum beam under different support conditions: (a)
cantilever; (b) simply supported; (c) fixed-fixed for modal frequency analysis.
the response signal. For example, Figure 10.9a and b shows the linear and logarithmic
plots of the Fourier spectrum, respectively, for the cantilevered beam. In the linear
plot of the Fourier spectrum (Figure 10.9a), the first two modal frequencies of the
beam can be easily identified (5.99 Hz and 34.93 Hz respectively). However, the
third mode may be easily overlooked due to the low energy of the displayed peak.
On the other hand, the logarithmic scale plot of the spectrum (Figure 10.9b) exhibits
the first three modal frequencies clearly. This is also true for the fixed-fixed support
condition (see Figure 10.10). However, for the simply supported condition, it is not
easy to distinguish the modes from the logarithmic scale plot of the Fourier spectrum.
It can be observed in Figure 10.11b that besides the three fundamental modal
frequencies, there are also a number of energy peaks at other frequency ranges
having amplitudes of order close to the modal frequencies. Hence for beams with
closely spaced frequencies, the Fourier spectrum may not be appropriate.
EMD was performed on the responses given in Figure 10.8. Consider the results
for the cantilevered beam: three decomposed IMF components and one residual
component are obtained, as depicted in Figure 10.12a. The individual IMFs are
subjected to FT, and the resulting spectra are shown in Figure 10.12b through d.
The residual component is not subjected to FT as it represents the mean or residual
trend of the captured response signal. Figure 10.12b through d clearly identifies the
TABLE 10.4
Analysis of Modal Frequencies of Aluminum Beam under Different Support Conditions Obtained
by FFT, EMD Followed by FT, HHT, and WT
FFT EMD + FT HHT WT
Theoretical Freq Error Freq Error Freq Error Freq Error
Support Conditions Mode Frequency (Hz) (Hz) (%) (Hz) (%) (Hz) (%) (Hz) (%)
25
(a)
20
15
10
Amplitude
5
0
−5
−10
−15
−20
−25
0 0.2 0.4 0.6 0.8 1
Time, t(s)
15
(b)
10
5
Amplitude
−5
−10
−15
0 0.2 0.4 0.6 0.8 1
Time, t(s)
25
(c)
20
15
10
Amplitude
−5
−10
−15
−20
0 0.1 0.2 0.3 0.4 0.5
Time, t(s)
FIGURE 10.8 Free vibration response of aluminum under (a) cantilever; (b) simply sup-
ported; (c) fixed-fixed conditions.
4
×10
2.5
1st Mode (a)
= 5.99 Hz
2
1.5
Amplitude
0.5 nd
2 Mode
rd
= 34.93 Hz 3 Mode
= 96.8 Hz
0
0 50 100 150 200 250
Frequency (Hz)
102
101
0
10
10−1
0 50 100 150 200 250
Frequency (Hz)
FIGURE 10.9 FFT analysis of response of cantilever beam given in Figure 10.8a, plotted on
(a) linear scale and (b) log-scale.
first three modal frequencies of the beam (5.76 Hz, 35.23 Hz, and 96.8 Hz respec-
tively). Even for the third mode, the appearance of the model frequencies is very
distinctive because the EMD process has effectively decomposed the response signal
into its respective components, which are narrow band. Hence, the energy peak of
the third mode is not obscured by the first and second modes, which contain relatively
higher energies. Similar analyses are done for the same beam under the two other
support conditions, and the results (see Figure 10.13 and Figure 10.14) obtained are
comparable to those obtained by the conventional direct FT. However, for the case
of the fixed-fixed condition, the third modal frequency could not be identified by
the FT method (Figure 10.14). Instead, the first and second modal frequencies are
1200 st
1 Mode (a)
= 22.89 Hz
1000
800
Amplitude
600
400
200 nd
rd
3 Mode
2 Mode
= 165.8 Hz
= 81.8 Hz
0
0 50 100 150 200 250
Frequency (Hz)
st
1 Mode
rd
3 Mode (b)
2 nd
10 = 22.89 Hz 2 Mode = 165.8 Hz
= 81.8 Hz
100
Amplitude
−2
10
−4
10
FIGURE 10.10 FFT analysis of response of fixed-fixed beam given in Figure 10.8c, plotted
on (a) linear scale and (b) log-scale.
each contained in two IMF components (fourth and fifth IMFs and second and third
IMFs respectively; see Figure 10.14). On closer observation, the third and fourth
modal frequencies can be spotted in the Fourier spectrum of the second IMF com-
ponent, but they are obscured by the stronger presence of the second modal fre-
quency. This setback of the proposed method can be explained by the fact that IMFs
may contain mix modes. This may occur when the frequency of interest in an IMF
is only contained along a particular segment of the signal and not throughout the
entire time domain. This is the so-called intermittency problem discussed by Huang
et al [17].
The original HHT was also performed on the decomposed IMF components of
the response signals for the beam under the different support conditions. However,
as mentioned earlier, the frequency of interest in an IMF may be contained only
1400
2nd Mode (a)
1200
st
1 Mode = 138 Hz
= 48 Hz
1000
Amplitude
800
600
400
rd
200 3 Mode
= 268 Hz
0
0 100 200 300 400 500
Frequency (Hz)
4
10 st nd
1 Mode 2 Mode (b)
3 = 48 Hz = 138 Hz
10
3rd Mode
2
10 = 268 Hz
th
4 Mode
Amplitude
1
10 = 446 Hz
0
10
−1
10
−2
10
−3
10
0 100 200 300 400 500
Frequency (Hz)
FIGURE 10.11 FFT analysis of response of simply supported beam given in Figure 10.8b,
plotted on (a) linear scale and (b) log-scale.
along a particular segment of the signal; hence, a weighted average using the
amplitude is performed on segments of the frequency spectrum from HT for each
IMF to obtain the modal frequencies of the beam. To illustrate, spectra for the beam
under the three support conditions will be presented in detail individually.
First, consider the HT of the IMFs for the cantilever beam (Figure 10.15). From
the Hilbert spectra of the three IMF components, it can be observed that the EMD
has effectively decomposed the original response signal into its different compo-
nents, which are well contained within a narrow band, as the fluctuation of the
frequency along the time axis is rather small. For this case, weighted averaging was
performed over the entire time domain for each of the IMF components. The modal
frequencies obtained are 101.56 Hz, 35.86 Hz, and 5.64 Hz from the first, second,
and third IMFs, respectively, which corresponds well with the theoretical values.
c(1)
0
−5
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
10
c(2)
−10
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 (a)
20
c(3)
−20
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
10
Residue
5
0
−5
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
Time, t(s)
180 400
rd (b) nd (c)
160 3 Mode 2 Mode
350
= 96.8 Hz = 35.23 Hz
140 300
120
Amplitude
Amplitude
250
100
200
80
150
60
40 100
20 50
0 0
0 50 100 150 200 250 0 50 100 150 200 250
Frequency (Hz) Frequency (Hz)
4
×10
3.5
st (d)
1 Mode
3 = 5.76 Hz
2.5
Amplitude
1.5
0.5
0
0 50 100 150 200 250
Frequency (Hz)
FIGURE 10.12 (a) IMF components of response of cantilever beam given in Figure 10.8a;
FT analysis of (b) first; (c) second; (d) third IMF component.
Next, for the simply supported beam, it can be observed in the HT spectra shown
in Figure 10.16 that the frequency is stable only over a certain time range for each
IMF. For example, for the first IMF, the frequency has a small fluctuation between
c(1)
0
−5
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
5
c(2)
−5
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 (a)
10
c(3)
−10
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
1
Residue
0
−1
−2
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
Time, t(s)
45 4.5
rd (b) nd (c)
40 3 Mode 4 2 Mode
= 167.5 Hz = 82.85 Hz
35 3.5
30 3
Amplitude
Amplitude
25 2.5
20 2
15 1.5
10 1
5 0.5
0 0
0 50 100 150 200 250 0 50 100 150 200 250
Frequency (Hz) Frequency (Hz)
900
rd (d)
800 1 Mode
= 22.95 Hz
700
600
Amplitude
500
400
300
200
100
0
0 50 100 150 200 250
Frequency (Hz)
FIGURE 10.13 (a) IMF components of response of simply supported beam given in Figure
10.8b; FT analysis of (b) first; (c) second; (d) third IMF component.
0.1 sec and 0.3 sec; hence, a weighted average on the frequency is done over this
time range. The frequency obtained is 159.18 Hz, which is closer to the third
vibration mode of the beam. Similar analysis is done for the second IMF (between
10
c(1) 0
−10
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5
5
c(2)
0
−5
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5
10
c(3)
0
−10
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5 (a)
5
c(4)
0
−5
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5
2
c(5)
0
−2
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5
0.5
Residue
−0.5
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5
Time, t(s)
0.7 90
(b) nd (c)
80 2 Mode
0.6 = 138 Hz
70
0.5
60
Amplitude
Amplitude
rd
0.4 50 3 Mode
(268 Hz)
0.3 40
30 th
0.2 4 Mode
20 (460 Hz)
0.1 10
0 0
0 100 200 300 400 500 600 0 100 200 300 400 500 600
Frequency (Hz) Frequency (Hz)
600 350
nd (d) st (e)
2 Mode 1 Mode
500 = 138 Hz 300 = 48 Hz
250
400
Amplitude
Amplitude
200
300
150
200
100
100 50
0 0
0 100 200 300 400 500 600 0 100 200 300 400 500 600
Frequency (Hz) Frequency (Hz)
FIGURE 10.14 (a) IMF components of response of fixed-fixed beam given in Figure 10.8c;
and FT analysis of (b) first; (c) second; (d) third; (e) fourth IMF; (f) fifth IMF component.
45 st
1 Mode (f )
40 = 48 Hz
35
30
Amplitude
25
20
15
10
0
0 100 200 300 400 500 600
Frequency (Hz)
0.12 sec and 0.28 sec), and the result (77.88 Hz) gives the second vibration mode.
For the third IMF, a rather stable fluctuation of the frequency is observed over the
entire time domain; hence, a weighted average is taken over the entire data range. This
obviously gave the result corresponding to the first fundamental frequency (22.80 Hz).
The Hilbert spectra for the IMF components of the fixed-fixed condition are
shown in Figure 10.17. Considering the first IMF, a large fluctuation of the frequency
over the entire time domain is observed. This can be explained by the fact that the
filtered component effectively contains the noise in the original response signal;
hence, meaningful modal parameters of the beam can be extracted from it. The
second IMF reveals a stable frequency region with high amplitude over the time
domain between 0.05 sec and 0.12 sec. Taking the weighted average over this time
range gives the second modal frequency (135.34 Hz). For the third IMF, two stable
frequency fluctuation regions with high amplitude can be identified, namely, from
0 to 0.16 sec and from 0.16 to 0.46 sec. By performing a weighted average on the
two regions, two fundamental frequencies are obtained (138 Hz and 263.19 Hz). A
similar stable region is observed for the fourth IMF, from 0.04 to 0.15 sec; this
corresponds to an averaged frequency of 46.96 Hz, which gives the first modal
frequency of the beam. In the fifth IMF, the two stable regions can be pinpointed,
from 0 to 0.15 sec and from 0.24 to 0.48 sec. These two regions give a weighted
average frequency of 42 Hz and 43.3 Hz respectively.
Last, wavelet analysis using the Morlet function was also performed on the three
signals of the beam under the three support conditions. WT was first done on the
original response signals for the beam over a large scale range. By observing the
dominant energy contents in this overall spectrum, different scales were zoomed in
to obtain the dominant frequency contents. For example, WT was first performed
for the vibration response of the cantilevered beam given in Figure 10.8a over a
wide scale range of 1 to 256. This spectrum plot (see Figure 10.18a) shows three
regions with a consistent existence of a corresponding frequency throughout the
500 2.0
450
400
350
Frequency, (Hz)
300
250 1.0 (a)
200
150
100
50
0 0
0 0.2 0.4 0.6 0.8 1
Time, t(s)
100 8.0
90
80
70
Frequency, (Hz)
60
50 4.0 (b)
40
30
20
10
0 0
0 0.2 0.4 0.6 0.8 1
Time, t(s)
60 20
50
Frequency, (Hz)
40
30 10 (c)
20
10
0 0
0 0.2 0.4 0.6 0.8 1
Time, t(s)
FIGURE 10.15 HT spectrum for (a) first, (b) second, and (c) third IMF component for
response of cantilever beam as given in Figure 10.12a.
entire time domain, namely, in scales 105 to 183, 14 to 27, and 4 to 12. By focusing
toward these regions, a more refined dominant frequency can be deduced, namely
at scales of 144, 22.66, and 8, which corresponds to frequencies of 5.64 Hz, 35.86
500 4.0
450
400
350
Frequency, (Hz)
300
250 2.0 (a)
200
150
100
50
0 0
0 0.2 0.4 0.6 0.8 1
Time, t(s)
300 3.0
250
Frequency, (Hz)
200
100
50
0 0
0 0.2 0.4 0.6 0.8 1
Time, t(s)
350 8
300
250
Frequency, (Hz)
200
4 (c)
150
100
50
0 0
0 0.2 0.4 0.6 0.8 1
Time, t(s)
FIGURE 10.16 HT spectrum for (a) first, (b) second, and (c) third IMF component for
response of simply supported beam as given in Figure 10.13a.
Hz, and 101.56 Hz respectively. These results give the three fundamental frequencies
of the beam. The wavelet analyses for the other two support conditions are presented
in Figure 10.19 and Figure 10.20.
5000 10
4500
4000
3500
Frequency, (Hz)
3000
2500 5 (a)
2000
1500
1000
500
0 0
0 0.1 0.2 0.3 0.4 0.5
Time, t(s)
2500 5
2000
Frequency, (Hz)
1500
2.5 (b)
1000
500
0 0
0 0.1 0.2 0.3 0.4 0.5
Time, t(s)
2000 10
1800
1600
1400
Frequency, (Hz)
1200
1000 5 (c)
800
600
400
200
0 0
0 0.1 0.2 0.3 0.4 0.5
Time, t(s)
FIGURE 10.17 HT spectrum for (a) first, (b) second, (c) third, (d) fourth, and (e) fifth IMF
component for response of fixed-fixed beam as given in Figure 10.14a.
The results obtained by the different signal analysis methods are compared in
Table 10.4. It can be observed that all the methods produce results of approximately
the same order of error. The methods are equally feasible in determining the modal
800 6
700
600
Frequency, (Hz)
500
400 3 (d)
300
200
100
0 0
0 0.1 0.2 0.3 0.4 0.5
Time, t(s)
300 2
250
Frequency, (Hz)
200
150 1 (e)
100
50
0 0
0 0.1 0.2 0.3 0.4 0.5
Time, t(s)
248
235
222
209
196
183
170
157
144
Scale
131 (a)
118
105
92
79
66
53
40
27
14
1
Scale of colors from MIN to MAX
Time
Scale
144
100
0
−100
−200
−300
100 200 300 400 500 600 700 800 900 1000
Time
FIGURE 10.18 Morlet WT analysis for response of cantilever beam given in Figure 10.8a
at scales (a) 1–256, (b) 144, (c) 22.66, and (d) 8 with sampling rate 1000 samples/sec.
frequency is narrowband and known a priori. For the last case, the combination of the
EMD process from HHT with FT on the decomposed IMF components has proven to
be capable of revealing modal frequencies of the aluminum beam having relatively
lower energy contents. These modal frequencies may be overlooked by the traditional
direct FFT, unless a logarithmic plot is done for the energy spectrum. However, the
problem of mixed modes within individual IMFs needs to be addressed. The original
Scale
22.66
−50
100 200 300 400 500 600 700 800 900 1000
Time
Scale
5
0
−5
−10
100 200 300 400 500 600 700 800 900 1000
Time
HHT with knowledge of the problem does provide good results when it is used with
an averaging concept. These examples show that HHT is a suitable tool for processing
nonstationary wave propagation signals encountered in damage detection studies. It
provides a good representation of localized events in both the frequency and energy of
any transient signal collected. Indeed, it is a simple signal-processing tool to use and
provides reasonably good results.
248
235
222
209
196
183
170
157
144
Scale
131
118
105 (a)
92
79
66
53
40
27
14
1
Scale of colors from MIN to MAX
Time
Scale
35
20
Amplitude
−20
−40
100 200 300 400 500 600 700 800 900 1000
Time
FIGURE 10.19 Morlet WT analysis for response of simply supported beam given in Figure
10.8b at scales (a) 1–256, (b) 33.5, (c) 10, and (d) 5 with sampling rate 1000 samples/sec.
ACKNOWLEDGMENTS
The authors wish to thank the National University of Singapore for providing the
support to perform this study.
Scale
10
1
0
−1
−2
−3
−4
100 200 300 400 500 600 700 800 900 1000
Time
Scale
2
0
−2
−4
−6
−8
100 200 300 400 500 600 700 800 900 1000
Time
1958
1855
1752
1649
1546
1443
1340
1237
1134
Scale
1031 (a)
928
825
722
619
516
413
310
207
104
1
Scale of colors from MIN to MAX
Time
Scale
163
(b)
Coefficients Line - Ca, b for scale a = 163 (frequency = 49.847)
80
60
40
Amplitude
20
0
−20
−40
−60
−80
500 1000 1500 2000 2500 3000 3500 4000 4500 5000
Time
FIGURE 10.20 Morlet WT analysis for response of fixed-fixed beam given in Figure 10.8c
at scales (a) 1–2042, (b) 163, (c) 60, (d) 28, and (e) 18 with sampling rate at 10000 samples/sec.
Scale
60
20
0
−20
−40
−60
−80
500 1000 1500 2000 2500 3000 3500 4000 4500 5000
Time
Scale
30
5
0
−5
−10
−15
500 1000 1500 2000 2500 3000 3500 4000 4500 5000
Time
Scale
163
20
0
−20
−40
−60
−80
500 1000 1500 2000 2500 3000 3500 4000 4500 5000
Time
REFERENCES
1. Crema L B, Peroni I and Castellani A (1985). Modal Tests on Composite Material
Structures Application in Damage Detection. Proc. Int. Modal Anal. Conf. Exhibit
Vol. 2, pp. 708–713.
2. Cook D A and Berthelot Y H (2001). Detection of Small Surface-Breaking Fatigue
Cracks in Steel Using Scattering of Rayleigh Waves. NDT & E Int. 34:483–492.
3. Hurlebaus S, Niethammer, Jacobs L J and Valle C (2001). Automated Methodology
to Locate Notches with Lamb Waves. Acous. Soc. Am. ARLO 2(4):97–102.
4. Staszewski W J (1998) Structural and Mechanical Damage Detection Using Wavelets.
Shock Vibr. Digest 30(6):457–472.
5. Kishimoto K, Inoue H, Hamada M and Shibuya T (1995). Time Frequency Analysis
of Dispersive Waves by Means of Wavelet Transform. ASME J. Appl. Mech. 62:
841–846.
6. Lee Y Y and Liew K M (2001). Detection of Damage Locations in a Beam Using
the Wavelet Analysis. Int. J. Struct. Stab. Dyn. 1(3):455–465.
7. Quek S T, Wang Q, Zhang L and Ong K H (2001). Practical Issues in Detection of
Damage in Beams Using Wavelets. Smart Mater. Struct. 10(5):1009–1017.
8. Huang H, Norden E, Zheng S, Long S R, Wu M C, Shih H H, Zheng Q, Yen N C,
Tung C C and Liu H H (1998). The Empirical Mode Decomposition and Hilbert
Spectrum for NonLinear and Nonstationary Time Series Analysis. Proc. R. Soc. Lond.
A 454:903–995.
9. Huang H, Norden E, Zheng S and Long S R (1999). A New View of Nonlinear Water
Waves: The Hilbert Spectrum. Annu. Rev. Fluid Mech. 31:417–457.
11 The Analysis of
Molecular Dynamics
Simulations by the
Hilbert-Huang Transform
Adrian P. Wiley, Robert J. Gledhill,
Stephen C. Phillips, Martin T. Swain,
Colin M. Edge, and Jonathan W. Essex
CONTENTS
ABSTRACT
Proteins are an integral component of all living systems, and obtaining a detailed
understanding of their structure and dynamics is of considerable importance. Molecular
dynamics computer simulations are an important tool in this process. However, the
length of simulation required to probe the full range of motions available to proteins
245
11.1 INTRODUCTION
Proteins are polymer chains made up from 20 possible monomer units (amino acids),
each sharing a common backbone structure with differing functional side chains (1).
YPGDV, a simple protein used in later discussion, consists of five amino acids:
tyrosine, proline, glycine, aspartic acid, and valine (Figure 11.1).
It is convenient to consider the amino acid chains (the primary structure) in
terms of relatively rigid and stable secondary structure units. These are classified as
either α-helices, where a section of coiled amino acids forms a rod-like helix, or
β-sheets, where amino acid strands form layered structures. These units of secondary
structure are connected by flexible loop regions, and secondary structure units can
themselves associate to form compact globular units called domains that move in a
correlated fashion.
Understanding the functions and mechanisms of biological molecules requires
detailed knowledge of their structure and internal motions. The arrangement, or
conformation, of a protein is of considerable importance to its biological function
(2, 3). For example, Creutzfeldt-Jakob Disease (CJD) is a fatal neurodegenerative
computer power now allow routine simulations of large protein systems of around
10 ns. Over this time scale, conformational events such as the opening of the two
large flaps in HIV-1 protease can be observed (18).
A classical description of atoms is generally used in MD; the energy of the
system is assumed to be independent of the location of electrons (BornOppenheimer
approximation), so atoms are not polarizable and are described by a point location
only. MD effectively solves Newton’s Second Law,F = ma, for each atom, using a
finite difference integrator to overcome the many body problem. One such integrator,
the Verlet algorithm (19), is shown in Equation 11.1. The time step, δt, must be
sufficiently small to keep the system energy stable, yet as large as possible for
efficient time evolution of the system (a 2 fs time step is commonly employed).
Forces are obtained from the gradient of a potential energy function. This function
has separate terms for bond stretching, angle deformations, dihedral motion, and
non-bonded interactions. The parameters used for the potential energy function are
optimized with data from vibrational spectra, experimental geometry information,
and ab initio quantum mechanical calculations (20).
( ) () ( ) ()
r t + δt = 2r t − r t − δt + δt 2a t . (11.1)
Simulations can be run with constant energy, although various thermostats (21,
22, 23, 24) and barostats (22, 23) can be applied to reproduce more biologically
relevant constant temperature and constant pressure ensembles.
Frequency analysis of molecular dynamics trajectories yields spectra that can
be compared to infra-red experimental data. The majority of intramolecular motions
are wavelike in nature and exist either as localized oscillations in single bonds or
angles, or as collective motions (or modes) characterized by coupled vibrations. For
example, the amide I mode in proteins that originates from the carbonyl stretching
vibration also includes a contribution from the backbone carbon–nitrogen bond (25).
Postsimulation digital filtering has been applied to molecular dynamics simula-
tions to remove high frequency motions that are irrelevant to conformational changes.
One simple approach Fourier transforms the atomic trajectories, applies a frequency-
domain filter, and inverse Fourier transforms the result (26). Alternatively the filtering
can be done in the time-domain by using non-recursive digital filters (27). These
methods have proved useful for the visualization of low frequency motions that are
important for conformational change.
Although MD has been used to study protein conformational change, the time
scales over which these motions occur are generally longer than those accessible by
conventional simulation routes. Methods to overcome this problem are therefore
required. Reversible Digitally Filtered Molecular Dynamics (RDFMD) is a method
that applies digital filters to the velocities in an evolving simulation to promote low
frequency motions. RDFMD has been successfully applied to a range of systems
(28, 29). The analysis of an RDFMD simulation cannot be achieved with sufficient
time resolution by using traditional Fourier-based methods, especially given the
discovery of nonstationary frequency targets (30). The Hilbert-Huang transform
provides the necessary analysis tools, as presented here.
∫ π (t − u) du
x (u)
h(t ) = Ρ (11.2)
−∞
In practice, the Hilbert transform can be computed by taking the Fourier trans-
form of the data, setting the negative frequency components to zero, doubling the
positive frequency components, and back-transforming (35). This method relies on
the infinite replication of the data set required by the Fourier transform, yielding
unreliable regions at the start and end of the transformed data due to the disconti-
nuities introduced. These regions are excluded from our analysis.
It can be shown that the signal magnitude is unchanged by HT, but the phase is
adjusted by π/2 (35). The original signal, x(t), and its Hilbert transform, h(t), may
be considered part of a complex signal, z(t):
() ()
z t = x t + ih t () (11.3)
()
z t =A t e ()
iφ t
() (11.4)
where
() ()
A t = x 2 t + h2 t () (11.5)
and
()
h t
φ t = tan −1 .
()
()
(11.6)
x t
The rate of change of the phase angle with time is the frequency of motion
occurring. The instantaneous frequency, f(t), is thus defined as:
()
f t =
1 dφ t()
. (11.7)
2π dt
The minimum frequency that can be reliably sampled by this method is 1/T (31),
where T is the time length of the data set. For an instantaneous frequency to be
physically meaningful, the signal must carry no riding waves and must be locally
symmetrical about its mean point, as defined by the envelope of the local maxima
and minima. Real-world phenomena that are described by signals matching these
criteria are not common.
EMD is a recently developed signal-processing method. It decomposes a signal,
which may be nonstationary, into a set of intrinsic mode functions (IMFs) suitable
for the Hilbert transform (31). An IMF is defined as a wave in which the number
of extrema and the number of zero-crossings differ by a maximum of one and where,
at any point, the mean of the envelope defined by the local maxima and minima is
zero.
To generate an IMF, the local mean is repeatedly determined and subtracted
from the data set until the number of extrema and zerocrossings in the residual data
differ by at most one (this process is termed “sifting”). The local mean is found by
taking the mean of a curve through all of the maxima and a curve through all of the
minima. These curves are defined, in this work, by cubic spline functions. The
algorithm proceeds by subtracting each generated IMF from the remaining signal
until either the recovered IMF or the residual data is small, or the residual has
become a trend component with only a single maximum and minimum. The final
residual of the data is similar in concept to the DC component of a Fourier transform,
except that it contains the overall trend of the data.
Once a signal has been decomposed into a set of IMFs, each IMF is separately
Hilbert transformed, giving the amplitude and instantaneous frequency at discrete
points in time. If the IMF is a harmonic oscillator of variable amplitude and fre-
quency, then its signal energy at time t can be written in terms of the signal amplitude,
A, and frequency, v:
( ) ( ) A (t ) .
2 2
E t ∝v t (11.8)
In the HHT spectra presented in this work, the energy in each IMF is calculated
and presented using Equation 11.8.
This paper presents the application of the HHT method to the analysis of MD
and RDFMD simulation trajectories, with a view to obtaining a better understanding
of system dynamics, validating the assumptions underlying the RDFMD methods,
and deriving some of the parameters associated with the RDFMD approach.
FIGURE 11.2 The bond length (C=O) and angle (CA-C-N) used in the frequency analysis.
FIGURE 11.3 Fourier transform of backbone degrees of freedom using a Hanning window
and a 50-point running average.
events that occur simultaneously across several dihedrals. Figure 11.7 plots the eight
main-chain dihedral angles of YPGDV that are most important for conformational
change (see Figure 11.1) as a function of time in a region where a spontaneous
conformational transition occurred. The energy in the 0 to 10 cm–1 region obtained
by integrating the HHT spectrum is also shown, and the three highest energy peaks
occur from 15 to 22 ps, corresponding to a rearrangement of dihedrals ψ2, φ3, and
ψ3. Several dihedrals are often seen moving at once, as this reduces the need for
significant solvent reorganization.
It is necessary to show that the EMD method produces physically relevant IMFs.
Results for the φ3 data of the spontaneous transition are shown in Figure 11.8. The
low frequency IMFs follow physical motions from the signal, and there is an obvious
separation between the high frequency (IMFs 1 to 5) and low frequency motions
(IMFs 6 to 10). The application of EMD as a signal-smoothing technique is also clear.
It is a requirement of the HHT method that frequencies of motion be separable.
Without this, artifacts may enter the data (31). This is inherent in the method because,
in each sifting iteration, the extracted IMF tries to follow the highest frequency
motion in the remaining signal. Meaningful separation is not always guaranteed for
dihedral data due to the flexible nature of the simulation model, and so EMD results
must always be interpreted with caution. However, we have shown the physical
relevance of IMFs and, as Figure 11.8 shows, we find that a single IMF describes
2000
1800
Frequency/wavenumbers
1600
1400
0 10 20 30
Time/ps
2000
1800
Frequency/wavenumbers
1600
1400
0 10 20 30
Time/ps
FIGURE 11.6 The first IMFs of the coupled angle, Gly / Asp CA-C-N, and bond motion,
Gly C=0.
300
Angle / degrees
200
100
0 10 20 30
Energy in 0 -10 cm-1 region
0
0 10 20 30
Time / ps
FIGURE 11.7 Top: The eight dihedral angle trajectories around a spontaneous conforma-
tional transition. Bottom: The energy in the 0 to 10 cm–1 region extracted using HHT.
the filter on the input signal. For our purposes, the desired frequency response is an
amplification of low frequencies (typically those lower than 100 cm–1) with no
change in the amplitude of higher frequencies. The coefficients of the digital filter
are chosen such that the actual frequency response of the filter matches the desired
response as closely as possible (see Figure 11.9). The longer the list of coefficients,
the closer the actual frequency response is to the desired frequency response. How-
ever, short filters are desirable as they require fewer data points and hence less
simulation. We have found filters of 1001 coefficients (i.e., m = 500) to be sufficient.
v′t = ∑c v
i =− m
i t −i . (11.9)
1.5
1
Frequency Response
0.5
−0.5
0 25 50 75 100
Frequency/wavenumbers
FIGURE 11.9 The desired and actual frequency responses for a 1001 coefficient filter.
saved as the midpoint and then restored after the application of the filter — ensuring
that the new velocities are in phase with the atomic positions.
After applying a filter, we allow the system to evolve for a number of steps d.
Another filter may then be applied in order to build up gradually the kinetic energy
in the low frequency modes. We have been using a value of d = 20 time steps (29),
and as a safeguard against unrealistic, high energy events, the filter series is termi-
nated early if the kinetic energy of the system exceeds that equivalent to a temper-
ature of 2000 K. Since new velocities are applied at the midpoint of a buffer, if d
is less than m, molecular dynamics must be run backward in time to fill the buffer,
as shown in Figure 11.10. This is possible with the velocity Verlet MD integrator (39).
This flexible method allows many filters to be applied, with subsequent filter
applications positioned anywhere in time relative to the first. It is, however, possible
that the system is not approaching a conformational transition, and amplifying low
frequency motions will have little effect. It is for this reason that the number of filter
applications in a series is limited to ten, after which the simulation proceeds as
normal from the end of the final filter buffer. Once filter applications have been
terminated (either by reaching the maximum number of filters allowed or by heating
to the temperature cut-off), 20 ps of MD are performed, giving a new starting
structure from which the filtering process can be repeated.
Prior to the RDFMD simulation, a desired frequency target must be found. A
short MD simulation of YPGDV has been used for this purpose. Since conforma-
tional motions are complex with non-stationary frequencies, Fourier methods yield
FIGURE 11.10 RDFMD sequence showing the delay parameter, d, and the filter length, 2m + 1.
200
Signal Energy/Arbitrary Units
Dihedral Angle/degrees
150
Signal
0–10
0–25
0–50 100
50
0
0 10 20 30
Time/ps
160
150
Displacement/degrees
170
160
150
450 500 550
Timesteps through buffer
FIGURE 11.12 YPGDV ψ2 data. Top: Signal before and after the application of a digital
filter. Bottom: Sum of the residual and the low frequency IMFs showing the trend in the data.
30
Filter 1
Filter 2
Filter 3
Filter 4
Filter 5
Dihedral angle change (RMSD)
20
10
0
Energy
FIGURE 11.13 Change in YPGDV backbone dihedral angles as low frequency kinetic energy
is successively amplified.
region. The dihedral motion is clearly amplified either side of the central point in
the direction of the low frequency trend.
The amount of additional motion in the backbone dihedral angles caused by the
application of a filter to the system should correlate with the amount of energy in
the low frequency region at the point of filter application. To test this hypothesis,
we define a measure of the additional dihedral motion going from one buffer to the
next as the average of the root-mean-square deviations (RMSDs) of the eight dihedral
angles in the central region of the buffer (the midpoint ± 100 steps). The amount of
energy in the low frequency region is calculated by taking the instantaneous fre-
quencies and amplitudes of the dihedral IMFs at the midpoint of the buffer and
calculating the energy at each frequency, as described previously. To estimate the
amount of energy in the filter amplification region, these energies are multiplied by
the filter’s frequency response, scaled from unity at 0 cm–1 to zero at 100 cm–1.
Finally, the energies are summed to provide our measure of the kinetic energy
available for amplification.
Figure 11.13 plots the dihedral angle RMSDs against the energy amplified for
five successive applications of a ×5 amplifying filter. The energy and the dihedral
motion are clearly correlated, though the data is quite spread. We also see a steady
increase in the energy as the digital filter is applied to successive filter buffers. Fewer
points are displayed for the later buffers due to simulations terminating as they
exceed the temperature cut-off. A trend line has been fitted to the entire data set by
a least-squares procedure, plotted to the extremes of the data set only.
FIGURE 11.14 Change in YPGDV backbone dihedral angles as low frequency kinetic
energy is successively amplified using different digital filters.
digital filter and the resulting RMSD in dihedral motion. As expected, using larger
amplifying filters increases the energy in the critical degrees of freedom more rapidly,
although there will inevitably be a compromise between the need for efficiency by
amplifying as rapidly as possible and the requirement that the conformational tran-
sitions be physically sensible without significantly disrupting the protein structure,
for which gentle amplification will be required. Finally, more detailed analysis of
spontaneous conformational transitions has confirmed that, although such transitions
are indeed associated with very low frequency motions (< 10 cm–1), they are also
associated with higher frequencies (25 to 50 cm–1). Analysis of the remainder of the
simulations trajectory has identified other occasions where significant energy resides
in this frequency region, but without associated conformational change. We suspect
that these conditions are associated with possible conformational change events.
Amplification of frequencies of up to 50 cm–1 may therefore be more efficient in
terms of driving conformational change.
11.5 LIMITATIONS
The limitations of the HHT method as applied to molecular dynamics simulations
are essentially threefold. First, analysis of short (1001 time step) RDFMD filter
buffers results in a low frequency limit in the Hilbert transform of 16.7 cm–1. Given
that this limit lies in the frequency range over which amplification is taking place,
it restricts our ability to analyze the simulation behavior over the content of the filter
buffer. Second, conformational transitions often leave large residuals, giving dihedral
angle plots similar to step functions. EMD produces physically unrealistic IMFs in
an attempt to recreate the unusual signal given by the data less the residual, as shown
in Figure 11.15. Although the data could be windowed to reduce the size of the
affected regions, low frequency resolution would be lost. This limitation is due to
the nature of the data and suggests that the HHT must be used with care, with
constant checks of the physical relevance of the IMFs. Third, the HHT requirement
of separable scales is not necessarily met by signals from molecular dynamics
simulations. Great care must therefore be taken when performing analyses by, for
example, careful examination of the IMFs to confirm their physical relevance,
repeated analysis over many different simulations, averaging over a number of
simulation signals, and integration of the HHT spectra over frequency and time.
11.6 CONCLUSIONS
In this paper the application of the HHT method to the analysis of molecular
dynamics computer simulations has been reported. The non-stationary data associ-
ated with these simulations are unsuitable for analysis with Fourier methods. How-
ever, it has been shown that the HHT method does yield useful insight into the
dynamics. In particular, the assumptions underpinning the RDFMD method of
enhancing conformational change, namely that low frequency vibrations are asso-
ciated with conformational change and that non-recursive digital filters may be used
to amplify these vibrations, have been confirmed. HHT analysis has also suggested
FIGURE 11.15 YPGDV ψ2 data. Top: Signal and residual (dotted line). Middle: Signal after
the residual is removed (equivalent to the sum of the IMFs). Bottom: IMFs generated by EMD.
that frequencies greater than those earlier identified with spontaneous conformational
change may in fact be more appropriate targets for RDFMD. However, the presence
of large residuals and an incomplete separation of scales indicate that HHT cannot
be applied to these simulations as a “black box” and that care must be taken.
REFERENCES
1. Creighton, T. E. (1993). Proteins structures and molecular properties. 2nd ed. New
York: W. H. Freeman and Company, pp. 192–193.
2. Gerstein, M., Lesk, A. M., Chothia, C. (1994). Structural mechanisms for domain
movements in proteins. Biochemistry. 33(22):6739–6749.
3. Källblade, P., Dean, P. M. (2003). Efficient conformational sampling of local sidechain
flexibility.J. Mol. Biol. 326:1651–1665.
4. Cardone, F., Liu, Q. G., Petraroli, R., Ladogana, A., D’Alessandro, M., Arpino, C.,
Di Bari, M., Macchi, G., Pocchiari, M. (1999). Prion protein glycotype analysis in
familial and sporadic Creutzfeldt-Jakob disease patients. Brain Res. Bull. 49(6):
429–433.
5. Caliskan, G., Kisliuk, A., Tsai, A. M., Soles, C. L., Sokolov, A. P. (2003). Protein
dynamics in viscous solvents. J. Chem. Phys. 118(9):4230–4236.
6. Rosca, F., Kumar, A. T. N., Ionascu, D., Ye, X., Demidov, A. A., Sjodin, T., Wharton,
D., Barrick, D., Sligar, S. G., Yonetani, T., Champion, P. M. (2002). Investigations
of anharmonic low-frequency oscillations in heme proteins. J. Phys. Chem A. 106(14):
3540–3552.
7. Rhodes, G. (1993). Crystallography Made Crystal Clear. 2nd ed. San Diego, Calif.:
Academic Press.
8. Zhang, X., Wozniak, J. A., Matthews, B. W. (1995). Protein flexibility and adaptability
seen in 25 crystal forms of T4 lysozyme. J. Mol. Biol. 250:527–522.
9. Hayward, S. (1999). Structural principles governing domain motions in proteins.
Proteins: Struct. Funct. Genet. 36:425–435.
10. Gogo, N. K., Skrynnikov, N. R., Dahlquist, F. W., Kay, L. E. (2001). What is the
average conformation of bacteriophage T4 lysozyme in solution? A domain orienta-
tion study using dipolar couplings measured by solution NMR. J. Mol. Biol. 308:
745–764.
11. Amadei, A., Linssen, A. B. M., Berendsen, H. J. C. (1993). Essential dynamics of
proteins. Proteins: Struct. Funct. Genet. 17(4):412–425.
12. Buchner, M., Ladanyi, B. M., Stratt, R. M. (1992). The short-time dynamics of
molecular liquids. Instantaneous-normal-mode theory. J. Chem. Phys. 97(11):8522–
8535.
13. de Groot, B. L., Hayward, S., van Aalten, D. M. F., Amadei, A., Berendsen, H. J. C.
(1998). Domain motions in bacteriophage T4 lysozyme: a comparison between
molecular dynamics and crystallographic data. Proteins: Struct. Funct. Genet. 31:
116–127.
14. Sanders, J. K. M., Hunter, B. K. (1993). Modern NMR Spectroscopy. 2nd ed. Oxford:
Oxford University Press.
15. Johnson, W. C. (1990). Protein secondary structure and circular dichroism — a
practical guide. Proteins: Struct. Funct. Genet. 7(3):205–214.
16. Fischer, M. W. F., Majumdar, A., Dahlquist, F. W., Zuiderweg, E. R. P. (1995). 15N,
13C, and 1H NMR assignments and secondary structure for T4-lysozyme. J. Magn.
12 Decomposition of Wave
Groups with EMD
Method
Wei Wang
CONTENTS
ABSTRACT
By adding a proper real temporary time series that can induce a series of additional
extrema, intermittent small fluctuations can be accurately decomposed from the large
waves they ride on. We can thus effectively solve the mode mixing problem that arises
when the amplitudes of the components of a time series are very different. Another
kind of mode mixing arises when the frequencies of the wave components are close
to each other, such as in wave groups. Here a temporary complex time series is
introduced to shift down the frequencies of the components. This will greatly enlarge
the ratio of the frequencies in a wave group, and the wave groups can then easily be
decomposed with the empirical mode decomposition method.
12.1 INTRODUCTION
Being among the most energetic events in the ocean wave field, wave groups play
an important role in ocean wave research. The significant characteristic of a wave
267
group is that the group velocity Cg is equal to the energy propagation speed of the
wave field; this is widely used in ocean wave models such as WAM. The formation
of wave groups is assumed analytically to be the result of the superposition of several
cosine waves with almost the same frequency. In practice, this kind of analysis often
cannot be achieved since no method up to now has been available to decompose a
wave group efficiently into several simple components. This problem has been the
main obstacle in wave group studies.
Feasible methods that might be used to decompose wave groups are Fourier
transform, wavelet transform (Chan 1995), and empirical mode decomposition
(EMD) analysis (Huang et al. 1998, 1999). For a continuous uniform wave group,
in which the wave group arises from the superposition of continuous constant-ampli-
tude cosine waves, Fourier transform can distinguish the frequencies of all the
components accurately. For an intermittent wave group, in which the wave group
arises from the superposition of intermittent cosine waves, Fourier transform can
yield only a collection of frequencies that will not represent the actual components.
The reason is that Fourier transform is a high-resolution frequency detector with no
resolution in the time domain; as a result, the intermittency of wave groups will
show as additional frequencies in the spectrum, and the actual frequencies of the
wave components will be altered to new locations.
Compared with Fourier transform, wavelet transform is more suitable for dealing
with intermittent signals. It has higher resolution in the time domain and lower
resolution in the frequency domain. Theoretically, intermittent wave groups can be
decomposed by wavelet transform when the frequencies of the wave components
are widely different, but this method would fail when the frequencies are only slightly
different. In practice, the lower resolution of wavelet transform in the frequency
domain will also induce a rich distribution of harmonics around the real components
and will ultimately result in difficulties in the distinction of the actual wave group
components.
In consideration of the resolution in the time–frequency domain, the most accept-
able method is EMD. Compared to the Fourier transform and the wavelet transform,
the EMD method has higher resolution in both the time and frequency domains. As
an example, Huang et al. (1998) decomposed a wave group that came from the linear
sum of two cosine waves into eight components with up to 3000 sifting processes
and an extremely stringent criterion. Although the first two components contained
most of the energy and were similar to the original components, the results are not
acceptable for the following reasons:
2 2
C1
1 0
−2
0 0 500 1000 1500 2000 2500 3000
2
−1
C2
0
−2 −2
0 500 1000 1500 2000 2500 3000 0 500 1000 1500 2000 2500 3000
FIGURE 12.1 The original time series and the first two IMF components decomposed with
the EMD method. The solid line in left panel is the original signal, the dashed line in the left
panel is the envelope, and the lines in the right panel are the IMF components.
the high frequency fluctuations in the original signal, and it can easily be separated
by subtracting the temp-signal from the IMFs. But an optionally selected temp-signal
will not match the original signal. Hence, the selection of the temp-signal is very
important.
For the case of a large wave with single frequency, containing an intermittent
signal with high frequency and small amplitude, we can write:
It is easily found that in order to restrict ε to smaller than 5%, the suitable choice
of the temp-signal satisfies 4T2/5 < T3 T1, in which T1, T2, and T3, are the periods
of the signal of the dominant wave, the fluctuations, and the temp-signal, respectively.
In addition, the amplitude of the temp-signal should be large enough to form a series
of extrema in the original signal.
For the signal of Figure 12.1, the mathematical description is:
2π
cos t , as t < 1264 or t > 1296
640
X (t ) = (12.4)
cos 2π t − 0.02 sin 2π t , as 1264 ≤ t ≤ 1296
640 32
1.5
−0.5
−1
−1.5
0 500 1000 1500 2000 2500 3000
t
FIGURE 12.2 The shape of the original signal with the superposition of the temp-signal.
0.03 0.03
(a) (b)
0.02 0.02
0.01 0.01
0 0
−0.01 −0.01
−0.02 −0.02
−0.03 −0.03
1200 1250 1300 1350 1400 1200 1250 1300 1350 1400
FIGURE 12.3 Comparison of the original small-scale fluctuation and the one decomposed
with the EMD method. The solid line represents the decomposed signal, and the dashed line
represents the original signal. In the left panel, the temp-signal is defined by atmp = 0.2 and
Ttmp = 60. In the right panel, the temp-signal is defined by atmp = 50 and Ttmp = 120.
The signal includes a small-scale fluctuation with frequency 50 times larger and
amplitude 20 times smaller than those of the dominant wave. Based on the discussion
above, the temp-signal
2π
X tmp (t ) = atmp cos t
Ttmp
should be selected in the range 26 < Ttmp < 640. Here we chose atmp = 0.2 and Ttmp
= 60. The composed signal is shown in Figure 12.2.
After processing with the EMD method and Deng et al.’s (2001) method of
dealing with the boundary condition, the intermittent small-scale fluctuation is
decomposed successfully, as shown in Figure 12.3a. As a comparison, different
choices of atmp and Ttmp were used, and the result was shown in Figure 12.3b.
The above example represents a case where the small-scale intermittent fluctu-
ation is located on the crest of the dominant wave. For a case where the small-scale
1 0.03
(a) (b)
A 0.02
0.5
0.01
0 0
−0.01
−0.5
−0.02
−1 −0.03
0 500 1000 1500 2000 2500 3000 1950 2000 2050 2100 2150
FIGURE 12.4 Same as Figure 12.3, but the intermittent small-scale fluctuation is changed
to the position marked by A in the left panel. The temp-signal is selected as atmp = 50 and
Ttmp = 120.
2 1
(a) (b)
1.5
1 0.5
0.5
0 0
−0.5
−1 −0.5
0 500 1000 1500 2000 2500 3000 1100 1200 1300 1400 1500 1600
in which ω2 > ω1. The ratio of the two frequencies is α = ω2/ω1. Typically α is
smaller than 1.1, but it can be greatly enlarged when we are subtracting a temporary
frequency, ω0, in the numerator and denominator simultaneously. For a real signal,
such a downshifting is hard to achieve. Theoretically, Equation 12.5 can be rewritten
in a complex form as
where Y(t) is the complex form of y(t), which can easily be constructed by using
the Hilbert transform. For an arbitrary time series, y(t), we can always have its
Hilbert transform, ỹ (t), as
∞
y( τ )
∫
1
y (t ) = P d τ, (12.7)
π −∞ t − τ
where P indicates the Cauchy principal value. This transform exists for all functions
of class Lp (Huang et al. 1998). Therefore, the complex function Y(t) can be formed as
Z (t ) = Zr (t ) + iZi (t )
= Y (t ) ⋅ exp(−iω 0t )
(12.9)
= exp(iω1t ) + exp(i(ω 2t + φ)) ⋅ exp(−iω 0t )
The ratio of the frequencies for signal Z(t) is α = (ω2 – ω0)/(ω1 – ω0).
For a well-selected ω0, α can be much larger than 1.5, the limitation in the EMD
method. Zr(t) and Zi(t) can easily be decomposed, either with the EMD method or
the method described in Section 12.2, and represented by their IMF components as:
Zr (t ) = ∑C
k =1
rk (t ) + rnr ,
n
(12.10)
Zi (t ) = ∑ C (t) + r .
k =1
ik ni
The combination of the corresponding IMF components of Zr(t) and Zi(t) will
reconstruct Z (t) as a sum of complex components Ck(t):
Z (t ) = Zr (t ) + iZi (t )
n
= ∑ (C (t) + iC (t)) + (r
k =1
r i k nr + irni ) (12.11)
= ∑ C (t) + R .
k =1
k n
Y (t ) = Z (t ) ⋅ exp(iω 0t )
n (12.12)
= ∑
k =1
Ck (t ) ⋅ exp(iω 0t ) + Rn ⋅ exp(iω 0t ).
y(t ) = Re Y (t )
n
= Re
k =1
∑
Ck (t ) ⋅ exp(iω 0t ) + Rn ⋅ exp(iω 0t )
(12.13)
= ∑ c (t) + r ,
k =1
k n
where ck(t) are IMF components of the original signal, and rn is the trend or mean
value.
12.4 VALIDATIONS
We offer two examples to verify the performance of the present method. The first
is a continuous uniform wave group with a phase shift in one of the components.
For ω1 = 2π/30, ω2 = 2π/32, and φ = π/6, the wave profile represented by Equation
12.5 is given in Figure 12.6. In this case α = 1.067. With ω0 = 2π/28, the wave
profile of Z(t), governed by Equation 12.9, is shown in Figure 12.7. Here, a and b
represent the real and imaginary components of Z(t), respectively.
Under such conditions, the ratio of the frequencies of the components is approx-
imately 2. Thus, Zr(t) and Zi(t) can easily be decomposed into a series of IMF
1.5
0.5
y (t)
−0.5
−1
−1.5
−2
0 200 400 600 800 1000
t
FIGURE 12.6 Wave profile represented by Equation 12.9 with ω1 = 2π/30, ω2 = 2π/32, and
φ = π/6.
2
Zr (t)
−2
2
Zi (t)
−2
0 200 400 600 800 1000
t
FIGURE 12.7 Wave profile represented by Equation 12.10 with ω0 = 2π/28. (a) The real
component of Z(t); (b) the imaginary component of Z(t).
components. Figure 12.8 shows the wave profiles of all the IMF components of the
real and imaginary parts.
Results calculated from Equation 12.11, Equation 12.12, and Equation 12.13
are shown in Figure 12.9 and Figure 12.10. The solid lines in Figure 12.9 and Figure
12.10 correspond to the first and second modes in Equation 12.5, respectively. For
comparison, the actual modes are also plotted as dotted lines in the same figures.
The agreement is excellent.
2 2
1 1
C1r
C1i
0 0
−1 −1
−2 −2
0 200 400 600 800 1000 0 200 400 600 800 1000
t t
2 2
1 1
C2r
C2i
0 0
−1 −1
−2 −2
0 200 400 600 800 1000 0 200 400 600 800 1000
t t
FIGURE 12.8 The IMF components of Zr(t) and Zi(t). (a,b) IMF components of Zr(t); (c,d)
IMF components of Zi(t).
1.5
0.5
−0.5
−1
−1.5
0 200 400 600 800 1000
t
FIGURE 12.9 Comparison of the first IMF component calculated from Equation 12.11, Equa-
tion 12.11, and Equation 12.13 (solid line) and the first mode of Equation 12.5 (dotted line).
Figure 12.11 illustrates the differences between the IMF components and the
actual modes. It is clear that the differences are negligible.
The second example is an intermittent wave group. The intermittent wave group
is formed as
1.5
0.5
−0.5
−1
−1.5
0 200 400 600 800 1000
t
FIGURE 12.10 Comparison of the second IMF component calculated from Equation 12.11,
Equation 12.11, and Equation 12.13 (solid line) and the second mode of Equation 12.5 (dotted
line).
1
0.5
0
−0.5
−1
0 200 400 600 800 1000
1
0.5
0
−0.5
−1
0 200 400 600 800 1000
t
FIGURE 12.11 Differences between IMF components and actual modes. The upper panel
shows the differences between the first IMF component and the first mode in Equation 12.5.
The lower panel shows the differences between the second IMF component and the second
mode in Equation 12.5.
For ω1 = 2π/30, ω2 = 2π/32, φ = π/6, t1 = 1740, and t2 = 3150, the wave profile
represented by Equation 12.14 is shown in Figure 12.12. With ω0 = 2π/28, the wave
profile of Z(t), governed by Equation 12.10, is shown in Figure 12.13, where a and
b represent the real and imaginary components of Z(t), respectively.
In the first example (shown in Figure 12.9, Figure 12.10, and Figure 12.11),
only negligible errors exist at the ends because we have confined end effects with
a method of adding characteristic waves at the ends. In the second example, however,
1.5
0.5
y (t)
−0.5
−1
−1.5
−2
0 500 1000 1500 2000 2500 3000 3500 4000 4500 5000
t
FIGURE 12.12 Wave profile represented by Equation 12.14 with ω1 = 2π/30, ω2 = 2π/32,
φ = π/6, t1 = 1740, and t2 = 3150.
2
(a)
1
Zr (t)
−1
−2
2
(b)
1
Zi (t)
−1
−2
0 500 1000 1500 2000 2500 3000 3500 4000 4500 5000
t
FIGURE 12.13 Wave profile represented by Equation 12.10 with ω0 = 2π/28. (a) The real
component of Z(t); (b) the imaginary component of Z(t).
end effects are hardly confined because the ends of the wave groups are not at the
ends of the data set. If the ends of the wave groups are left unattended, wide swings
caused by spline fitting will exist at the ends and will propagate to the interior and
ulterior as the number of sifting iterations increases. Figure 12.14 and Figure 12.15
show the final results.
2 2
(a) (c)
1 1
Cr1
Ci1
0 0
−1 −1
−2 −2
2 2
(b) (d)
1 1
Cr2
Ci2
0 0
−1 −1
−2 −2
0 1000 2000 3000 4000 5000 0 1000 2000 3000 4000 5000
t t
FIGURE 12.14 The IMF components of Zr(t) and Zi(t): (a,b) IMF components of Zr(t); (c,d)
IMF components of Zi(t).
2
(a)
1
−1
−2
2
(b)
1
−1
−2
0 500 1000 1500 2000 2500 3000 3500 4000 4500 5000
t
FIGURE 12.15 Comparison of the IMF components calculated from Equation 12.11, Equa-
tion 12.12, and Equation 12.13 (solid line) and the two actual modes of Equation 12.14 (dotted
line).
12.5 CONCLUSIONS
The present paper introduces two methods for decomposing a wave group into a
series of components of relatively single scale. The first method was used not only
in solving the problem of decomposition of wave groups but also as an improved
method of EMD for a variety of uses. The second method can decompose wave
groups of small α, but the sensitivity to the selection of the temporary frequency is
still an open question.
REFERENCES
Chan, Y.T. (1995). Wavelet basics. Boston: Kluwer.
Deng, Y.J., W. Wang, C.C. Qian, Z. Wang and D.J. Dai. (2001). Boundary-processing tech-
nique in EMD method and Hilbert transform. Chinese Science Bulletin, 46(11),
954–960.
Huang, N.E., Z. Shen, S.R. Long, et al. (1998). The empirical mode decomposition and the
Hilbert spectrum for nonlinear and non-stationary time series analysis. Proc. R. Soc.
Lond. A, 454, 899–995.
Huang, N.E., Z. Shen and S.R. Long. (1999). A new view of nonlinear water waves: the
Hilbert spectrum. Annu. Rev. Fluid Mech., 31, 417–457.
13 Perspectives on the
Theory and Practices
of the Hilbert-Huang
Transform
Nii O. Attoh-Okine
CONTENTS
13.1 Introduction ..................................................................................................282
13.2 Basic Ideas ...................................................................................................282
13.2.1 The Hilbert-Huang Transform .........................................................282
13.2.2 Starting Point....................................................................................288
13.3 Current Applications ....................................................................................289
13.3.1 Biomedical Applications ..................................................................289
13.3.2 Chemistry and Chemical Engineering.............................................289
13.3.3 Financial Applications......................................................................290
13.3.4 Meteorological and Atmospheric Applications ...............................290
13.3.5 Ocean Engineering...........................................................................291
13.3.6 Seismic Studies ................................................................................292
13.3.7 Structural Applications.....................................................................295
13.3.8 Health Monitoring............................................................................296
13.3.9 System Identification........................................................................296
13.4 Some Limitations .........................................................................................297
13.5 Potential Future Research ............................................................................298
References..............................................................................................................299
Addendum: Perspectives on the Theory and Practices
of the Hilbert-Huang Transform............................................................................302
13A.1 Analytical .........................................................................................302
13A.2 Hybrid Method.................................................................................302
13A.3 Bidimensional EMD ........................................................................303
13A.4 Some Observations...........................................................................304
References..............................................................................................................304
281
ABSTRACT
The theory and practice of the Hilbert-Huang transform (HHT) provide another way
of analyzing nonlinear and nonstationary time series. The central idea of HHT is the
empirical mode decomposition (EMD), which is then decomposed into a set of intrinsic
mode functions (IMFs) and analyzed by using the Hilbert spectrum. This chapter is
primarily a summary of current applications of HHT and suggests areas for future
research based on my personal views.
13.1 INTRODUCTION
The Hilbert-Huang transform (HHT) provides a new method of analyzing nonsta-
tionary and nonlinear time series data. It allows the exploration of intermittent and
amplitude-varying motions. The classical signal analysis theory and practice have
been dominated by the Fourier transform, which represents the signal system in the
frequency domain. The Fourier works well for strictly periodic or stationary random
functions of time. To address the issues of nonstationarity and nonperiodic functions,
various methods have been introduced in the literature. These include wavelet anal-
ysis, the Wigner-Ville distribution, and the evolutionary spectrum. Unfortunately,
these methods failed in one way or another [Huang et al., 1998].
HHT is emerging as a powerful signal-processing tool. During the past five
years, a great deal has been learned about the computation, interpretation, and
implementation of the HHT technique. This work has been scattered in a wide variety
of fields, including signal processing, structural health monitoring, earthquake engi-
neering, nondestructive testing, transportation engineering, ocean engineering, and
financial and biomedical applications. No one has drawn the new work together in
a comprehensive manner.
This chapter is primarily a summary of my own current views. My failure to
mention a particular contribution should not be taken as an indication of disinterest
or disagreement. After this general introduction, basic ideas of HHT are presented
in a manner that should be accessible to readers with no previous familiarity with
HHT. The paper then discusses various application areas, outlines some perceived
“difficulties,” and finally highlights a future research direction of HHT.
history of the signal is decomposed into a finite and often small number of IMFs
such that they may be Hilbert transformed. The EMD is based on local characteristics
in the time scale of the data, making the approach direct, posterior, and adaptive.
The IMFs admit well-behaved Hilbert transforms.
An IMF is defined by the following two criteria:
• Identify all the local extrema, then connect all the local maxima by cubic
spline as upper envelope.
• Repeat the procedure for the minima to produce the lower envelope. The
upper and lower envelope should cover all the data. If the mean is desig-
nated as m1 and the difference between the data and m1 is the first com-
ponent h1, then
x (t ) − m1 = h1 (13.1)
where h1 is an IMF.
The mean m1 is given by the sum of local extrema connected by cubic spline:
L +U
m1 =
2
After repeating the sifting process up to k times, h1k becomes an IMF. That is:
Let h1k = c1, the first IMF from the data. c1 should contain the finest scale or the
shortest period component of the data/profile. Now c1 can be separated from the rest
of the data by
x (t ) − c1 = r1 . (13.4)
x (t ) = ∑ c +r ,
j =1
j n (13.6)
where cj is the jth IMF and n is the number of sifted IMFs. rn can be interpreted as
a trend in the signal/profile. The cj has zero mean. Due to the iterative procedure,
none of the sifted IMFs derived is closed analytical form [Schlurmann, 2002]. The
IMFs can be linear or nonlinear based on the characteristics of the data. The IMFs
are almost orthogonal and form a complete basis. Their sum equals the original data
[Salisbury and Wimbush, 2002]. The EMD then picks out the highest-frequency
oscillation that remains in the signal.
Flandrin et al. [2003] established how EMD can be used as a filterbank. They
represented the signal as follows:
x (t ) = mk (t ) + ∑ d (t)
k =1
k (13.7)
where mk(t) stands for a residual “trend,” and the modes {dk(t), k = 1, …, K} and
dk(t) stand for details.
The orthogonality of the IMF components can be checked as follows, if Equation
13.6 is expressed as
n +1
x (t ) = ∑ c (t).
j =1
j (13.8)
That is, one has included the last residual or data trend as an additional element:
n +1 n +1 n +1
x (t ) =
2
∑ c (t) + 2∑ ∑ c (t)c (t).
j =1
2
j
j =1 k =1
j k (13.9)
If the decomposition is orthogonal, then the cross terms given in the second part
on the right hand side should be zero when they are integrated over time. The overall
index of orthogonality OI can be defined as follows:
IO = ∑ ∑ ∑
T n +1 n +1
c (t )c (t )
.
j =1 k =1
j k
(13.10)
t =o
2
x (t )
To ensure that the IMF components retain enough physical sense of both ampli-
tude and frequency modulations, we can limit the size of the standard deviation
(SD), computed from two consecutive sifting results as
SD = ∑ |h
t =0
k −1 (t ) − hk (t ) |2
hk2−1 (t )
x (t ) = Re j (13.11)
j =1
Equation 13.7 gives both the amplitude and frequency of each component as a
function of time. The Fourier representation would be as follows:
x (t ) = Re ∑ a exp
j =1
j
iω j t
(13.12)
the classical Hilbert transform of the original data might possess unrealistic features
[Huang et al., 1998]. This implies that the IMF represents a generalized Fourier
expansion. The variable amplitude and instantaneous frequency enable the expansion
to accommodate nonstationary data. Huang et al. [2001] demonstrated that the
Fourier- and wavelet-based components and spectra may not have a clear physical
meaning as in HHT. For example, the wavelet-based interpretation of a pavement
profile is meaningful relative to selected mother wavelets [Attoh-Okine, 1999].
Furthermore, Hilbert analysis [Long et al., 1995] is based on almost noncausal
singular information. Therefore, at any given time, the data or signal has only one
amplitude and one frequency, both of which can be determined locally. These
represent the best information at that particular time. Physically, the definition of
instantaneous frequency has true meaning for monocomponent signals, where there
is one frequency, or at least a narrow range of frequencies, varying as a function of
time. Since most data do not show these necessary characteristics, sometimes the
Hilbert transform makes little physical sense in practical applications. In the present
method, the EMD is used to decompose the signal into a series of monocomponent
signals. Furthermore, to extract significant information, the time–frequency–ampli-
tude joint distribution [a(t), (t), t] can be developed. This joint distribution in 3D
space can be replaced by [H(,t)], where x(t) = H(,t). This final representation is
referred to as the Hilbert spectrum.
With the Hilbert spectrum defined, we can define a marginal spectrum:
∫
t
h(ω ) = H (ω, t )dt . (13.13)
0
The marginal spectrum offers a measure of the total amplitude (or energy) contri-
bution from each of the frequency values [Huang et al., 1999a]. The degree of
stationarity is defined as follows:
H (ω, t )
∫
T
1
D(ω ) = (1 − )dt . (13.14)
T 0 h(ω ) / T
The closer the above equation is to zero, the more stationary the system is and the
instantaneous energy:
IE (t ) =
∫ ω
H 2 (ω, t )d ω (13.15)
n n
H (ω, t ) = ∑
j =1
H j (ω, t ) ≈
_
∑ A (t)
j =1
j (13.16)
n n
∑ ∑∫
T
h(ω ) = (ω ) ≈
hj _ a j (t )dt (13.17)
0
j =1 j =1
provides a measure of the total amplitude contribution for each frequency value
[Zhang et al., 2003].
By using the Hilbert spectrum, first and second order time-moments and the
frequency-dependent order can be derived. The first and second order of time-depen-
dent moments are as follows:
mω ,1 (t ) =
∫ −∞
∞
ωH (ω, t )d ω
(13.18)
∫ −∞
H (ω, t )d ω
mω ,2 (t ) =
∫ ω H (ω, t)dω − m
−∞
∞
2
ω ,1 (t ) (13.19)
∫ H (ω, t)dω
−∞
mt,1 (ω ) =
∫ −∞
∞
tH (ω, t )dt
(13.20)
∫ −∞
H (ω, t )dt
mt ,2 (ω ) =
∫ −∞
∞
t 2 H (ω, t )dt
− mt ,1 (ω ) (13.21)
∫ −∞
H (ω, t )dt
The marginal in the time and frequency domains can be calculated as follows:
∞
mω ,0 (t )
∫ ∞
H (ω, t )d ω (13.22)
∞
mt,0 (ω ) =
∫∞
H (ω, t )dt (13.23)
The information content of the marginal mt,0() is much reduced [Huang et al., 2001].
The time marginal corresponds to the instantaneous power |x(t)|2 of the signal, and
the frequency marginal corresponds to the spectral energy density |X()|2.
Huang et al. [2003a] developed a confidence limit for HHT. The ensemble mean
and standard deviation of IMF sample sets obtained with different stopping criteria
are calculated and form a simple random set. The confidence limit for EMD/HSA
is then defined as a range of standard deviations from the assembled mean. The
authors attempted to establish a confidence limit for the EMD/HSA approach as a
statistical measure of the validity of the results. The method of establishing a
confidence limit depends on factors like stopping criteria, the maximum number of
siftings, intermittence, sifting methods, and the nomenclature needed for the confi-
dence limit. The method does not invoke the ergodic assumption. Qu and Jarzynski
[2001] did a comparative study of different time–frequency representations of Lamb
waves. The authors noted that the time–frequency accuracy of the Hilbert spectrum
is dependent on the accuracy of the EMD. That is, if the decomposition into IMFs
does not capture the signal’s behavior, the resulting Hilbert spectrum will not give
precise time–frequency results.
∞
f (τ)
∫
1
fˆ = dτ (13.24)
π −∞ t−τ
Y (t ) = f (t ) + i ∧ f (t ) (13.25)
or
which is a local time-varying wave with amplitude a(t) and phase (t). The Hilbert
transform can be considered to be a filter that simply shifts the phase of its input
by –/2 radians. Unfortunately, most signals are not band limited. One of Huang’s
notable contributions was to devise a method — the sifting method — that transforms
a wide class of signals into set of band-limited functions. The instantaneous fre-
quency may possess negative values. Lai (1998) developed a proper method to
implement positive frequencies. The Hilbert spectrum is similar to the Fourier
spectrum in a physical sense, but the Fourier approach is more global. In the Hilbert
spectrum, at any given time, the signal has only one amplitude and one frequency,
both determined locally [Long et al., 1995].
− t −t 0 − t −t 0
M k (t ) = A + B(t − t0 )e T1
+ C (1 − e T2
); t0 ≤ t t1 (13.28)
where t0 is the instant in time when oxygen concentration drops suddenly, t1 is the
instant when oxygen concentration increases suddenly, A is the mean value of Mk(t)
before time t0, and B, C, T1, and T2 are constants.
Montesinos et al. [2002] applied HHT to signals obtained from BWR neuron
stability. The authors used data from a nuclear power plant that were recorded during
startup of the plant. They show a sharp peak produced by sudden activity insertion
due to the change of the recirculation pump velocity from low to high speed. The
authors attempted to recognize the IMF that is related to the stability of the BWR.
They further used an autoregressive model to analyze the fourth IMF and showed
that the IMFs are mutually orthogonal.
Sh (t )
V (t; T ) = (13.29)
S (t )
where T corresponds to the period at the Hilbert spectrum peak of the high-passed
signal up to h terms, and
Sh (t ) = ∑ c (t).
j =1
j (13.30)
The unit of this variability parameter is the fraction of the market value. This is
a simple and direct measure of the market volatility.
events. The authors further developed a dynamical model using selected IMFs and
SOI data that can make a predictive estimate.
Pan et al. [2002] used HHT to analyze satellite scatterometer wind data over the
northwestern Pacific and compared the results to vector empirical orthogonal func-
tion (VEOF) results. Duffy [2004] used HHT to analyze three different data files:
The HHT persistently detected periodic features such as tides, snowmelt, and
heavy precipitation events in terms of sea level heights. Using the radiation data,
Duffy [2004] observed anomalies during ENSO and its effect on the aerosol con-
centration in the troposphere. Some of the IMFs of the observation data could be
paired with meteorological events, such as the passage of a cold front, a warm front,
and a trough. Generally, the HHT is capable of discerning mesoscale and synoptic
events.
With a five-day time series of temperature, Lundquist [2003] demonstrated how
the HHT allows simultaneous identification of the stationary diurnal temperature
cycle as well as intermittent and nonstationary cooling events such as frontal pas-
sages and density currents. One of the IMFs obtained by the authors appears to be
associated with inertia gravity waves.
and nonlinearity. They also found that the energy flux computed using the HHT
spectrum is higher than that from wavelet and FFT techniques.
Larsen et al. [2004] used HHT to characterize the underwater electromagnetic
environment and identify transient manmade electromagnetic disturbances. The
authors collected field data based on in-water magnetic data. The HHT was able to
act as a filter, effectively discriminating different dipole components. Therefore, the
method helps in detecting moving electric dipoles in the ocean. The authors further
developed an empirical mode–matched filter. The main idea is to apply a matched
filter to a signal generated by using IMFs as basis function, rather than to the target
signal. In the empirical mode–matched filter, EMD is initially applied to the signal
(time series) to generate the IMFs, i(t), i = 1, …, N, where N is the number of IMFs.
The second step involves determining the linear combination of the IMFs containing
the target signal s(t) that is best approximated to the known target signature in the
least-square sense:
min s(t ) −
ci ≥ 0 ∑ c α .
i =1
i i (13.31)
A matched filter based on the target signature s(t) is applied to the modeled
signal, which is composed of weighted linear combination of the IMFs. The authors
attempted to use receiver operator characteristics (ROC) curves, HHT, and the
matched filter to address the probability of detection. Zeris and Prinos [2004]
compared the results of using HHT and wavelet transform for turbulent open channel
flow data. Laser Doppler and hot films techniques were used to obtain signals in
turbulent open channel flow, with and without the imposed bed suction. The authors
were fairly successful in the application. Unfortunately, it is not clear how the authors
did the comparative analysis.
Chen et al. [2002a] used HHT to determine the dispersion curves of seismic
surface waves and compared their results to Fourier-based time–frequency analysis.
The surface-wave dispersion measurement is essential for the study of the crust and
upper mantle structures, earthquake source mechanisms, and inelastic properties of
the earth. The authors applied EMD to the seismogram and obtained the IMFs from
which the Hilbert spectrum is calculated. Then, for each frequency, the time of the
maximum amplitude corresponds to the group arrival at which the group velocity
can be obtained. The result of group velocity and phase velocity determined by using
HHT is more accurate than by Fourier-based methods.
Shen et al. [2003] applied HHT to ground motion and compared the HHT result
with the Fourier spectrum. The HHT is capable of identifying the occurrence and
properties of destruction near fault ground motions. The authors used the method
as a demonstration tool in the estimation of seismic loading in bridge design.
Magrin-Chagnolleau and Baraniuk [undated] applied HHT to decompose a seismic
trace into different oscillatory modes and were able to extract instantaneous frequen-
cies of these modes. Chun-xiang and Qi-feng [2003] made a comparative study of
wavelet and HHT using earthquake record. The IMF directly decomposed from the
original data, whereas wavelet components are decomposed according to mother
wavelets, which means the results can be influenced by the selected mother wavelets.
The Hilbert spectrum clearly presents the energy distribution with time and fre-
quency.
Zhang et al. [2003] examined the applicability of HHT to dynamic and earth-
quake loading motion recordings. The IMF was capable of capturing low and high
frequency components. The authors commented how EMD can act like a filter by
grouping the IMFs. The authors defined EMD-based high frequency motions as the
summation of the first few IMF components, whereas the EMD-based low frequency
motion is the summation of the remaining components. Unfortunately, the authors
failed to give the exact number of IMFs needed to be summed. This choice, high-
lighted by the authors, is very subjective.
Wen and Gu [2004; Gu and Wen, 2004] extended the HHT and applied their
methods to earthquake ground motions. The authors express x(t) as follows:
x (t ) = Re ∑ a (t) exp
j =1
j
iθ j ( t )
+ rn (t ) (13.32)
x (t ) = Re ∑ a (t) exp
j =1
j
i [ θ j ( t )+ φ j ]
+ rn (t ) . (13.33)
This makes x(t) a random process. The process then has the following mean, cova-
riance, and variance functions:
n
µ x (t ) = Re ∑ iθ iφ
a j (t )e j (t ) E (e j ) + rn (t ) = rn (t ) , (13.34)
j =1
K XX (t1, t2 ) =
1
2 ∑ a (t )a (t ) cos[θ (t ) − θ (t ) ,
j =1
j 1 j 2 j 1 j 2 (13.35)
n
σ x (t ) = E [ x (t ) − µx (t )] =
2
1
2
2
∑ a (t).
j =1
2
j (13.36)
The authors extended the approach to a vector process and developed the
cross-correlation covariance as follows:
x k (t ) = Re ∑a j =1
jk (t ) exp
i [ θ jk ( t )+ φ jk ]
+ rnk (t ) . (13.37)
With the modified approach, the spectra of the simulated samples compare well
with those of the record.
Zhang [2005] used a modified HHT to estimate the damping factor of nonlinear
soil and its role in seismic wave responses at soil sites, using data from earthquake
recordings. Zhang presented the following equation:
n n
x (t ) = Re ∑
j =1
a j (t ) exp
iθ j ( t )
= x (t ) = Re ∑ Λ (t) exp
j =1
j
[ − ϕ j ( t )+ iθ j ( t )]
(13.38)
where j(t) can be interpreted as the source-related intensity, j(t) is exponential factor
characterizing the time-dependent decay of waves in the jth IMF component due to
damping, and
a j (t ) = Re ∑ Λ (t) exp
j =1
j
− ϕ j (t )
. (13.39)
d ϕ j (t )
η j (t ) = . (13.40)
dt
n
.a j (t ) .Λ j (t )
D(ω, t ) = ∑ − a (t) + Λ (t) dt
j =1 j j
(13.41)
n
.a j (t ) .Λ j (t )
∫ ∑∫
T T
Λ
d (ω ) = D(ω, t )dt = − + dt = d (ω ) + d (ω ) .
a
(13.42)
0
j =1
0 a j (t ) Λ j (t )
It appears that the authors were successful in all cases. Quek et al. [2004] did
a comparative analysis of HHT, wavelet transform, and FFT in damage detection.
Although the HHT was found to be a more direct method compared to the wavelet
transform, the authors provide similar results in terms of precision and accuracy.
The authors advocated the combination of EMD with Fourier transform on the
decomposed IMF components. They ascertained that this combination is capable of
revealing modal frequencies with relatively low energy content.
Using HHT, Li et al. [2003] analyzed the results of a pseudodynamic test of two
rectangular reinforced concrete bridge columns. The authors conducted extensive
laboratory investigation to model a near-fault ground acceleration of the Chi-Chi
Earthquake. The HHT was used to analyze the signal output from various columns
tested. The aim of these analyses was to understand the structural behaviors of the
and random decrement technique (RDT) are much better than those resulting from
the fast Fourier approach. In their studies, to identify instantaneous frequency and
damping ratio, the measured response time history of the MDOF system is decom-
posed into IMFs. The authors checked the validity of model by applying it to a real
civil engineering structure. The results show that the FFT-based method fails to
identify modal damping ratios when two modal frequencies are too close.
Xu et al. [2003] compared the modal frequencies and damping ratios in various
time increments and different winds for one of the tallest composite buildings in the
world. The results were then compared to FFT. Once again, the HHT method
outperformed the FFT. The FFT-based results significantly overestimate the first
lateral and longitudinal modal damping ratios of the building.
Yang et al. [2002] developed a damage identification–based HHT. The authors
applied the technique to the ASCE structural health monitoring benchmark structure.
The approach was quite successful in identifying damage sections. Yang et al.
[2003a] used HHT to identify multi-degree-of-freedom linear systems by using
measured free vibration time histories. The authors proposed band filter and EMD
cases with polluted signals and high modal frequencies. Failure to introduce the
band-pass filtering approach will result in a large number of siftings. The Fourier
spectrum of the acceleration signal helps to identify the range of natural frequencies.
The acceleration signal is processed through band-pass filters. The time history
obtained from each band-pass proceeds through EMD. A linear least-square fit
procedure is proposed to identify natural frequencies, damping ratios, mode shapes,
mass matrix, damping matrix, and stiffness matrix. Yang et al. [2003b] extended
their work [Yang et al., 2003a] to identify general linear structures with complex
modes by using free vibration response data polluted by noise. Simulated results
show that the approach is capable of identifying the complex mode shapes as well
as the mass, stiffness, and damping matrices. This can help with the complete
dynamic characteristics of general linear structure.
• The approximate envelopes obtained using the local maxima and minima
do not always encompass all the data.
• The use of discrete signals instead of continuous introduces numerical
uncertainty in the true extrema.
The spline at the beginning and the end of the envelope can result in large
variation, which is very critical in a low frequency signal.
Deng et al. [2001] discussed the use of a cubic spline in calculating the lower
and upper envelopes. The authors mentioned the data extension method as key in
addressing this issue. They used a two-step extension method based on a neural
network approach. The original signal was extended forward and backward, and a
neural network was used to learn the envelopes.
Chen and Feng [undated] proposed a technique to improve the HHT procedure.
The authors noted that the EMD has its limitation in distinguishing different
components in narrow-band signals. The narrow band may contain either (a) com-
ponents that have adjacent frequencies or (b) components where the frequencies
may not be adjacent, but one of the components has a dominant energy intensity
much higher than the other components. The improved technique is based on beat-
ing-phenomenon waves.
Olhede and Walden [2004] compared HHT to a maximal-overlap discrete wave-
let transform (MODWT) and wavelet packet transforms (MODWPT). The authors
replaced EMD projections by a wavelet-based one and developed a Hilbert spectrum
via wavelet packets. The authors claimed the MODWPT and MODWT to be superior
to the HHT using a particular class of Fejer-Korovkin wavelet filters. It appears that
their assertion may be true only for nonlinear data. Attoh-Okine [2004] did a
comparative analysis of HHT, MODWPT, and MODWT on pavement profile data
and deduced that the HHT outperforms the MODWPT and MODWT in the case of
nonlinear and nonstationary data.
Datig and Schlurmann [2004] did the most comprehensive studies on the per-
formance and limitations of HHT with particular applications to irregular waves.
The authors did extensive investigation into the spline interpolation. The authors
discussed using additional points, both forward and backward, to determine better
envelopes. They also did a parametric study on the proposed improvement and
showed significant improvement in the overall EMD computations. The authors
noted that HHT is capable of differentiating between time-variant components from
any given data. Their study also showed that HHT was able to distinguish between
riding and carrier waves.
REFERENCES
Attoh-Okine, N. O. (1999). Application of Wavelets in Pavement Profile Evaluation and
Assessment. Proc. Estonian Acad. Science, 5, 53–63.
Attoh-Okine, N. O. (2004). Comparative Analysis of Hilbert-Huang Transform and Maximal
Overlap Discrete Wavelet Packet Transform. Working paper, University of Delaware
Engineering Department.
Chen, C. H., Li, C. P., and Teng, T. L. (2002a). Surface-Wave Dispersion Measurements
Using Hilbert-Huang Transform. TAO 13, 2:171–184.
Chen, J., and Xu, Y. L. (2002). Identification of Modal Damping Ratios of Structures with
Closely Spaced Modal Frequencies. Struct. Eng. Mech. 14, 4:417–434.
Chen, Y., and Feng, M. Q. (Undated). A Technique to Improve the Empirical Mode of
Decomposition in the Hilbert-Huang Transform. p. 24.
Chun-xiang, S., and Qi-feng, L. (2003). Hilbert-Huang Transform and Wavelet Analysis of
Time History Signal. Acta Seismological Sinica 166, 4:422–429.
Datig, M., and Schlurmann, T. (2004). Performance and Limitation of the Hilbert-Huang
Transformation with an Application to Irregular Water Waves. Ocean Eng. 31, 14–15:
1783–1834.
Deng, Y., et al. (2001). Boundary-Processing Technique in EMD Method and Hilbert Trans-
form. Chinese Science Bull. 46, 11:954–961.
Duffy, D. (2004). The Application of Hilbert-Huang Transform to Meteorological Datasets.
J. Atmospheric Oceanic Technol. 21, 599–611.
Flandrin, P., Rilling, G., and Goncalves, P. (2003). Empirical Mode of Decomposition as a
Filter Bank. IEEE Signal Process. Lett.
Gabor, D. (1946). Theory of Communication. Proc. IEE 93, 429–457.
Gu, P., and Wen, Y. K. (2005). Simulation of Nonstationary Random Processes Using Instan-
taneous Frequency and Amplitude from Hilbert-Huang Transform. In this Volume.
Huang, N. E., et al. (1999a). A New View of Nonlinear Water Waves: The Hilbert Spectrum.
Annu. Rev. Fluid Mech. 31, 417–57.
Huang N. E., et al. (2003a). A Confidence Limit for the Empirical Mode Decomposition and
Hilbert Spectrum Analysis. Proc. R. Soc. London A, 459, 2317–2344.
Huang, N. E., Chern, C. C., Huang, K., Salvino, L. W., Long, S. R., and Fan, K. L. (2001).
A New Spectral Representation of Earthquake Data: Hilbert Spectral Analysis of
Station TCU129, Chi-Chi Taiwan, 21 September 1999. Bull. Seismological Soc. Am.
91, 5:1310–1338.
Huang, N. E., Shen, Z., Long, S. R., Wu, M. C., Shih, E. H., Zheng, Q., Tung, C. C., and Liu,
H. H. (1998). The Empirical Mode Decomposition Method and the Hilbert Spectrum
for Non-Stationary Time Series Analysis. Proc. R. Soc. London A 454, 903–995.
Huang, W., Shen, Z., Huang N. E., and Fung, Y. C. (1999b). Engineering Analysis of Biological
Variables: An Example of Blood Pressure Over 1 Day. Proc. National Acad. Science
95, 4816–4821.
Huang, W., Shen, Z., Huang N. E. and Fung, Y. C. (1999c). Nonlinear Indicial Response of
Complex Nonstationary Oscillations as Pulmonary Hypertension Responding to Step
Hypoxia. Proc. National Acad. Science 96, 1834–1839.
Huang, N. E., Wu, M.-L., Qu, W., Long, S. R., Shen, S. S. P., and Zhang, J. E. (2003b).
Application of Hilbert-Huang Transform to Non-Stationary Financial Time Series
Analysis. Appl. Stochastic Models Bus. Industry 19, 245–268.
Hwang, P. A., Wang, D. W., and Kaihatu, J. M. (2005). A Comparison of the Energy Flux
Computation of Shoaling Waves Using Hilbert and Wavelet Spectral Anlysis Tech-
niques. In this Volume.
Lai, Y. (1998). Analytic Signals and the Transition to Chaos Deterministic Flows. Phys. Rev.
E 58, 6:R6911–6914.
Larsen, M. L., Ridgway, J., Waldman, C. H., Gabbay, M., Buntzen, R. R., and Battista, B.
(2005). Nonlinear Signal Processing of Underwater Electromagnetic Data. In this
Volume.
Li, Y. F., Chang, S. Y., Tzeng, W. C., and Huang, K. (2003). The Pseudo Dynamic Test of
RC Bridge Columns Analyzed Through the Hilbert-Huang Transform. Chin. J. Mech.
A 19, 3: 373–387.
Loh, C.-H., Wu, T. C., and Huang, N. E. (2001). Application of the EMD-Hilbert Spectrum
to Identify Near-Fault Ground Motion Characteristics and Structural Responses. Bull.
Seismological Soc. Am. 191, 5: 1339–1353.
Long, S. R., Huang, N. E., Tung, C. C., Wu, M. L., Lin, R. Q., et al. (1995). The Hilbert
Techniques: An Alternative Approach for Non-Steady Time Series Analysis. IEEE
Geoscience Remote Sensing Soc. Lett. 3, 3–11.
Lundquist, J. K. (2003). Intermittent and Elliptical Inertia Oscillation in Atmospheric Bound-
ary Layer. J. Am. Meteorological Soc. 60, 2261–2273.
Magrin-Chagnolleau, I., and Baraniuk, R. G. (undated). Empirical Mode Decomposition
Based Time Frequency Attribute. pp 4.
Montesinos, M. E., Munoz-Cobo, J. L., and Perez, C. (2002). Hilbert-Huang Analysis of
BWR Neutron Detector Signals: Application to DR Calculation and to Corrupted
Signal Analysis. Ann. Nuclear Energy 30, 715–727.
Neithammer, M., et al. (2001). Time–Frequency Representations of Lamb Waves. J. Acoustical
Soc. Am. 109, 5: Pt. 1, 1841–1847.
Olhede, S., and Walden, A. T. (2004). The Hilbert Spectrum via Wavelet Projections. Proc.
R. Society A 460, 2044: 955–975.
Osegueda, R., Kreinovich, V., Nazarian, S., and Roldan, E. (2003). Detection of Cracks at
Rivet Holes in Thin Plates Using Lamb-Wave Scanning. Proc. SPIE 5047, 55–66.
Pan, J., Yan, X. H., Zheng, Q., Liu, W. T., and Klemas, V. V. (2002). Interpretation of
Scatterometer Ocean Surface Wind Vector EOFs over the Northwestern Pacific.
Remote Sensing Environ. 84, 53–68.
Phillips, S. C., Gledhill, R. J., and Essex, J. W. (2003). Applications of the Hilbert-Huang
Transform to the Analysis of Molecular Dynamics Simulations. J. Phys. Chem. A
107, 4869–4876.
Pines, D., and Salvino, L. (2002). Health Monitoring of One-Dimensional Structures Using
Empirical Mode Decomposition and the Hilbert-Huang Transform. Proc. SPIE 4701,
127–143.
Qu, J., and Jarzynski, J. (2001). Time Frequency Representation of Lamb Waves. J. Acoust.
Soc. Am., 109(5) Part 1, 1841–1847.
Quek, S. T., Tua, P. S., and Wang, Q. (2003). Detecting Anomalies in Beams and Plate Based
on the Hilbert- Huang Transform of Real Signals. Smart Mater. Structures 12,
447–460.
Quek, S. T., Tua, P. S., and Wang, Q. (2004). Comparison of Hilbert-Huang, Wavelet, and
Fourier Transforms for Selected Applications. In this Volume.
Salisbury, J. I., and Wimbush, M. (2002). Using Modern Time Series Analysis Techniques to
Predict ENSO Events from SOI Time Series. Nonlinear Processes in Geophysics 9,
341–345.
Schlurmann, T. (2002). Spectral Analysis of Nonlinear Water Waves Based on the Hil-
bert-Huang Transformation. Trans. ASME 124, 22–27.
Schlurmann, T., and Datig, M. (2005). Carrier and Riding Wave Structure of Rogue Waves.
In this Volume.
Shen, J. J., Yen, W. P., and Fallon, J. O. (2003). Interpretation and Application of Hilbert-Huang
Transformation for Seismic Performance Analyses. Advanced Mitigation Technolo-
gies 657–666.
Veltcheva, A. (2005). An Application of HHT Method to the Nearshore Sea Waves. In this
Volume.
Veltcheva, A. D. (2002). Wave and Group Transformation by a Hilbert Spectrum. Coastal
Eng. J. 44, 4: 283–300.
Wang, W. (2005). Decomposition of Wave Groups with EMD Method. In this Volume.
Wen, Y. K., and Gu, P. (2004). Description and Simulation of Nonstationary Processes Based
in Hilbert Spectra. J. of Eng. Mech. 130, 8: 942–951.
Wiley, A. P., Gledhill, R. J., Phillips, S. C., Swain, M. T., Edge, C. M., and Essex, J. W.
(2005). The Analysis of Molecular Dynamics Simulations by the Hilbert-Huang
Transform. In this Volume.
Xiang, S. C., and Feng, L. Q. (2003). Hilbert-Huang Transform and Wavelet Analysis of Time
History Signal. Acta Seismological Sinica 16, 4: 422–429.
Xu, Y. L., Chen, S. W., and Zhang, R. C. (2003). Modal Identification of Di Wang Building
Under Typhoon York Using the Hilbert-Huang Transform Method. The Structural
Design of Tall and Special Buildings 12, 21–47.
Yang, J. N., Lei, Y., Pan, S., and Huang, N. (2003a). System Identification of Linear Structures
Based on Hilbert-Huang Spectral Analysis. Part 1: Normal Modes. Earthquake Eng.
Structural Dynamics 32: 1443–1467.
Yang, J. N., Lei, Y., Pan, S., and Huang, N. (2003b). System Identification of Linear Structures
Based on Hilbert-Huang Spectral Analysis. Part 1: Complex Mode. Earthquake Eng.
Structural Dynamics 32: 1533–1554.
Yang, J. N., Lei, Y., Lin, S., and Huang, N. (2004). Hilbert-Huang Based Approach for
Structural Damage Detection. J. Eng. Mech. 130, 1: 85–95.
Yang, J. N., Lin, S., and Pan, S. (2002). Damage Identification of Structures Using Hil-
bert-Huang Spectral Analysis. 15th ASCE Eng. Mech. Conf., New York.
Yu, D., Cheng, J., and Yang, Y. (2003). Application of EMD Method and Hilbert Spectrum
to the Fault Diagnosis of Roller Bearings. Mechanical Syst. Signal Process. 19, 2:
259–270.
Zeris, A., and Prinos, P. (2005). Coherent Structures Analysis in Turbulent Open-Channel
Flow Using Huang-Hilbert and Wavelets Transforms. In this Volume.
Zhang., R. R. (2004). An HHT Based Approach to Quantify Nonlinear Soil Amplification.
In this Volume.
Zhang, R. R., Demark, L. V., Liang, J., and Hu, Y. (2004). On Estimating Site Damping with
Soil Non-Linearity from Earthquake Recordings. Int. J. Non-Linear Mech. 39,
1501–1517.
Zhang, R. R., Ma, S., Safak, E., and Hartzell, S. (2003). Hilbert-Huang Transform Analysis
of Dynamic and Earthquake Motion Recordings. J. Eng. Mech. 129, 8: 861–875.
13A.1 ANALYTICAL
Coughlin and Tung [2004] used EMD to extract the solar cycle signal from strato-
spheric data. The authors highlighted some difficulties in using EMD — especially
the influence of the end point in the sifting process. They addressed a method of
reducing the influence of the end effects on the internal solution. Coughlin and Tung
[2004] extended both the beginning and end of the spline by introducing a wave
equation of the form:
2πt
Wave Extension = A sin + phase + local mean (13A.1)
p
The typical amplitude, A, and period p, are determined by the nearest local extrema.
() ()
Abeginning = max 1 − min 1
( )
Aend = max N − min N ( )
(13A.2)
() ()
pbeginning = 2 time max 1 − time min 1
( )
pend = 2 time max N − time min N( )
where max(1) and min(1) are the first two local extrema in the time series and
max(N) and min(N) are the last two local extrema. This approach reduces the large
swings in the spline calculation. This approach is quite different from the one
proposed by Dätig and Schlurmman [2004].
Loutridis [2004] used EMD to decompose vibration signals from gear systems.
The author developed an empirical law that relates the energy of the intrinsic modes
to the crack magnitude of the gear. Using this technique, the degree of crack can be
predicted based on energy obtained from the IMFs.
∑ λ Φ( x − x )
N
() ()
s x = pm x + i i (13A.3)
i =1
1. Find the positions and amplitudes of all local maxima and minima in the
input signal in_{lk}(m,n), where m and n denote spatial dimensions of the
image.
2. Create upper and lower envelopes by spline interpolation, denoted by
e_{max}(m,n) and e_{min}(m,n).
3. For each position, calculate the mean of the upper and lower envelopes:
emlk m, n =( ) (
emax m, n + emin m, n ) ( ) (13A.4)
2
( )
hlk m, n = inlk m, n − emlk m, n( ) ( ) (13A.5)
(
emlk m, n ≺ ε ) (13A.6)
( )
hl( k +1) m, n = hlk m, n ( ) (13A.7)
( )
rl m, n = inl1 m, n = cl m, n ( ) ( ) (13A.9)
(
in(l +1)1 m, n = rl m, n ) ( ) (13A.10)
( )
x m, n = rL m, n + ( ) ∑ c ( m, n )
i =1
j (13A.11)
REFERENCES
Coughlin, K. T. and Tung, K. K. (2004). 11-Year Solar Cycle in the Stratosphere Extracted
by the Empirical Mode Decomposition Method. Adv. Space Res. 34, 323–329.
Dätig, M. and Schlurmann, T. (2004). Performance and Limitations of the Hilbert-Huang
Transformation with an Application to Irregular Water Waves. Ocean Eng. 31, 14/15:
1783–1834.
Iyengar, R. N. and Kanth R. S. T. G (2004). Intrinsic Mode Functions and a Strategy for
Forecasting Indian Monsoon Rainfall. Meteorol. Atmospheric Phys. (online publica-
tion).
Lee, Z. K., Wu, T. H. and Loh, C. H. (2003). System Identification on Seismic Behavior of
an Isolated Bridge. Earthquake Eng. Structural Dynamics 32, 1797–1812.
Linderhed, A. (2003). 2-D Empirical Mode Decomposition: In Spirit of Image Compression.
SPIE Proc. 4738, 1–8.
Linderhed, A. (2003). Image Compression Based on Empirical Mode Decomposition. Proc.
SSBA Symp. Image Anal. 110–113.
Linderhed, A. (2004). Adaptive Image Compression with Wavelet Packets and Empirical
Mode Decomposition. Dissertation No. 909 submitted to Linkoping Studies in Science
and Technology.
Liu, Z. and Peng, S. (2005). Boundary Processing of Bidimensional EMD Using Texture
Synthesis. IEEE Signal Process. Lett. 12, 1: 33–36.
Loutridis, S. J. (2004). Damage Detection in Gear Systems Using Empirical Mode Decom-
position. Eng. Structures 26, 1833–1841.
Nunes, J. C., Bouaoune, Y., Delechelle, E., Niang, O. and Bunel P. (2003). Image Analysis
by Bidimensional Empirical Mode Decomposition. Image Vision Comput. 21,
1019–1026.
Nunes, J. C., Niang, O., Bouaoune, Y., Delechelle, E., and Bunel P. (2003a). Bidimensional
Empirical Mode Decomposition Modified for Texture Analysis. SCIA 2003, Lecture
Notes on Computer Science 2749, 171–177.