Efg 149
Efg 149
Efg 149
Its Applications
The subject of wavelet analysis has recently drawn a great deal of attention
from mathematical scientists in various disciplines. It is creating a common
link between mathematicians, physicists, and electrical engineers. This
book series will consist of both monographs and edited volumes on the
theory and applications of this rapidly developing subject. Its objective is to
meet the needs of academic, industrial, and governmental researchers, as
well as to provide instructional material for teaching at both the undergrad-
uate and graduate levels.
This fourth volume of the series is a compilation of twelve papers de-
voted to wavelet analysis of geophysical processes. In addition to an intro-
ductory review article written by the editors, this volume covers such
important areas as atmospheric turbulence, seismic data analysis, detec-
tion of signals in noisy environments, multifractal analysis, and analysis of
long memory geophysical processes. The series editor is grateful to Profes-
sor Foufoula-Georgiou and Dr. Kumar for their effort in compiling and ed-
iting this excellent volume, and would like to thank the authors for their
very fine contributions.
This is a volume in
WAVELET ANALYSIS A N D I T S A P P L I C A T I O N S
Praveen Kumar
Universities Space Research Association
Hydrological Sciences Branch
NASA-Goddard Space Flight Center
Greenbelt, Maryland
®
ACADEMIC PRESS
San Diego New York Boston
London Sydney Tokyo Toronto
This book is printed on acid-free paper, fe)
ix
x Contributors
xi
Xll Preface
Chapter 10, by Davis, Marshak, and Wiscombe, explores the use of wave-
lets for multifractal analysis of geophysical phenomena. They show the appli-
cability of wavelet transforms to compute simple yet dynamically meaningful
statistical properties of one dimensional geophysical series. Turbulent velocity
and cloud liquid water content are used as examples to demonstrate the need
for stochastic models having both additive (nonstationary) and multiplicative
(intermittent) features. Merging wavelet and multifractal analysis seems
promising for both wavelet and multifractal communities and is especially
promising for geophysics, where many signals show structures at all observa-
ble scales and are often successfully described within a multifractal frame-
work.
The wavelet transform partitions the frequency axis in a particular way:
it iteratively partitions the low-frequency components, leaving the high-fre-
quency components intact at each iteration. For some processes or applications
this partition might not achieve the best decomposition, as partition of the
high-frequency bands might also be necessary. Wavelet packets provide such a
partition and are used in Chapter 11, by Saito, for simultaneous signal
compression and noise reduction of geophysical signals. A maximum entropy
criterion is used to obtain the best basis out of the many bases t h a t the redun-
dant wavelet packet representation provides. The method is applied to syn-
thetic signals and to some geophysical data, for example, a radioactivity profile
of subsurface formation and a migrated seismic section. Finally, in Chapter 12,
Percival and Guttorp examine a particular measure of variability for long
memory processes (the Allan variance) within the wavelet framework and show
that this variance can be interpreted as a Haar wavelet coefficient variance.
This suggests an approach to assessing the variability of general wavelet
classes which will be useful in the study of power-law processes extensively
used for the description of geophysical time series. A fairly extensive bibliog-
raphy of wavelet analysis in geophysics is included at the end of this volume.
Several individuals provided invaluable help in the completion of this vol-
ume. Special t h a n k s go to the reviewers of the book chapters who volunteered
their time and expertise and provided timely and thoughtful reviews. The first
author t h a n k s Mike Jasinski, Hydrologie Sciences Branch at NASA-Goddard
Space Flight Center and the Universities Space Research Association for their
support during the completion of this project. We are grateful to Charu Gupta
Kumar, who converted most of the chapters to I^T^X format and typeset and
edited the entire volume. Without her expertise and dedication the timely
completion of this volume would not have been possible. Finally, we also t h a n k
the Academic Press Editor, Peter Renz, for his efficient help during the final
stages of this project.
§1. P r o l o g u e
T h e concept of wavelet transforms was formalized in early 1980's in a
series of p a p e r s by Morlet et al. [42, 43], Grossmann and Morlet [24], a n d
Goupillaud, Grossmann and Morlet [23]. Since this formalism a n d some
significant work by Meyer ([38, 39] a n d references therein), Mallat [30, 31],
Daubechies [13, 15], a n d Chui [6] among others, t h e wavelets have become
pervasive in several diverse areas such as m a t h e m a t i c s , physics, digital sig-
nal processing, vision, numerical analysis, and geophysics, t o n a m e a few.
Wavelet transforms are integral transforms using integration kernels called
wavelets. These wavelets are essentially used in two ways when studying
Wavelets in Geophysics 1
Efi Foufoula-Georgiou and Praveen Kumar (eds.), pp. 1-43.
Copyright 1994 by Academic Press, Inc.
All rights of reproduction in any form reserved.
ISBN 0-12-262850-0
2 P. Kumar and E. Foufoula-Georgiou
§2. T i m e - F r e q u e n c y A n a l y s i s
T h e original motive for the development of wavelet transform was (see
[23]) " . . . of devising a m e t h o d of acquisition, transformation and recording
of a seismic trace (i.e., a function of one variable, the time) so as to satisfy
the requirements listed below:
50 100 50 100
angular frequency angular frequency
Figure 1. Spectral and wavelet analysis of two signals. The first sig-
nal (a) (upper left) consists of superposition of two frequencies (sinlOt
and sin20£), and the second consists of the same two frequencies each
applied separately over half of the signal duration (b) (upper right).
Figures (c) (middle left) and (d) (middle right) show the spectra
of signals, i.e., |/(ω)| 2 vs ω, in (a) and (b) respectively, and (e) (lower
left) and (f) (lower right) show the magnitude of their wavelet trans-
forms (using Morlet wavelet) respectively.
two frequencies (sin 10t and sin 20t) superimposed for the entire duration
of the signal and the second consisting of the same frequencies, b u t each
one applied separately for half of signal duration. Figures lc,d show t h e
spectrum, i.e., |/(o;)| 2 , of these two signals, respectively. As is clearly
evident, the spectrum is quite incapable of distinguishing between the two
signals.
Time varying frequencies are quite common in music, speech, seismic
signals, non-stationary geophysical processes, etc. To study such processes,
one seeks transforms which will enable one to obtain the frequency content
of a process locally in time. There are essentially two methods t h a t have
been developed to achieve this (within the limits of the uncertainty principle
Wavelet Analysis in Geophysics: An Introduction 5
2.1. W i n d o w e d Fourier t r a n s f o r m
2.1.1. Definition
In t h e Fourier transform framework, time localization can be achieved
by windowing the d a t a at various times, say, using a windowing function
#(£), and then taking the Fourier transform. T h a t is, the windowed Fourier
transform (also called the short-time fourier transform), G/(CJ, i ) , is given
by
(2)
(3)
(4)
(5)
1
support is defined as the closure of the set over which the signal/process is non-zero.
6 P. Kumar and E. Foufoula-Georgiou
For the discrete windowed Fourier transform to be invert ible, the condition
CJO^O < 2π must hold (see [15], sections 3.4 and 4.1).
2 Λ . 2 . T i m e - f r e q u e n c y localization
In order to study the time-frequency localization property of the win-
dowed Fourier transform, we need to study the properties of |<7ω,*|2 and
\çju>,t | 2 since they determine the features of f(t) t h a t are extracted. Indeed,
using Parseval's theorem, equation (3) can be written as
1 f°° A
(7)
Wavelet Analysis in Geophysics: An Introduction 7
(8)
and
(9)
These parameters measure the spread of the function | ^ ? < | and \gUjt |, about
t and CJ, respectively (see Figure 3). Owing to the uncertainty principle,
the products of σ2, and σ | satisfy (see [31])
(10)
σ
σβ
' ΡJ
1 ^Λ 1 ^Λ
1 8, | §
° g .1 σ
*
1
σ Λ|
g)
1
1
1 ^
U
σ σ
8 . 8
1 ^Λ 1 ^Λ
1 8 | §
to
2.2.1. Definition
T h e wavelet transform of a function f(t) with finite energy is defined
as t h e integral transform with a family of functions xjj\)t{u) = ~7χΦ(Μγ^)
and is given as
(12)
for all scales λ (notice the identity φ(ί) = V>i,o(0)· We also choose the
normalization j \φ(ί)\2 dt = 1. The wavelet transform Wf(X^t) is often
denoted as t h e inner product (/, φ\$)-
Notice t h a t in contrast to the windowed Fourier transform case, t h e
number of cycles in the wavelet φ\^(η) does not change with t h e dilation
(scale) p a r a m e t e r λ but the support length does. We will see shortly t h a t
when λ is small, which corresponds to small support length, t h e wavelet
transform picks u p higher frequency components and vice-versa.
T h e choice of the wavelet φ{ί) is neither unique nor arbitrary. T h e
function φ{ί) is a function with unit energy chosen so t h a t it has:
|ψ(ω)|
where
(15)
(16)
2 . 2 . 2 . T i m e - f r e q u e n c y localization
In order to understand the behavior of the wavelet transform in the fre-
quency domain as well, it is useful to recognize t h a t the wavelet transform
W/(A,£), using Parseval's theorem, can be equivalently written as
(17)
(18)
We therefore define the standard deviation (i.e., square root of the second
central moment of the right lobe) σ τ as
(19)
Similarly in the time domain the standard deviation σφλ t can b e obtained
as
(20)
where to is given as
(21)
σ
Φχ,ι = λσ
Φι,ο- ( 22 )
ω°; = - ^ . (24)
From the above relationships one can easily see that as λ increases, i.e., as
the function dilates, both ω°7 and σ,?, decrease indicating that the center
of passing band shifts towards low frequency components and the uncer-
tainty also decreases, and vice-versa (see also figure 5). In the phase-space,
the resolution cell for the wavelet transform around the point (ίο,ω^ )
is given by [to ± λσ^ 1 0 x Alf0 ± Alf0] (see Figure 6) which has variable
dimensions depending on the scale parameter λ. However, the area of the
resolution cell [σφχ t x σ ? ] remains independent of the scale or location
parameter. In other words, the phase space is layered with resolution cells
of varying dimensions which are functions of scale such that they have a
constant area. Therefore, due to the uncertainty principle, an increased
resolution in the time domain for the time localization of high frequency
components comes with a cost: an increased uncertainty in the frequency
localization as measured by σ ? . One may also interpret the wavelet
transform as a mathematical microscope where the magnification is given
by 1/λ.
(25)
Wavelet Analysis in Geophysics: An Introduction 13
HL
λσ^
Since the mapping f(t) —> f(Xt) has the effect of contracting f(t) when
λ > 1 and magnifying it when λ < 1, the above equation indicates t h a t as
the scale grows, a contracted version of the function is seen through a fixed
size filter and vice-versa. Thus, the scale factor λ has the interpretation of
the scale in maps.
3 . 2 . S c a l o g r a m , w a v e l e t variance a n d covariance
From the isometry of wavelet transform (16) we have
(26)
In general, for two functions f(t) and g(t) (see [15], equation 2.4.2)
(27)
(28)
where
(29)
gives the energy content of a function f(t) at scale λ, i.e., it gives the
marginal density function of energy at different scales λ. The function
E(X) has been referred to as wavelet variance (see [4]) or wavelet spectrum
(see [25]). In analogy, the function
(30)
(31)
(32)
By defining
(33)
the above equation can be written as
(34)
One would therefore expect t h a t Ε'(ω), and thus JE7(A), is related to the
power spectrum S/(u>) of f(t). This indeed is the case. It can be shown
(see [25]) t h a t
(35)
where Ξψχ (ω) is the spectrum of the wavelet at scale λ. T h a t is, E(s) is
the weighted average of the power spectrum of f(t) where the weights are
Wavelet Analysis in Geophysics: An Introduction 15
where R(r) and 5(CJ) are the auto-covariance function and the power spec-
t r u m of the stochastic process X(t), respectively. If an analogous rela-
tionship could be developed for non-stationary processes using the wavelet
transform, then the properties of the wavelet transform could be harnessed
in a more useful way. It turns out t h a t , indeed, such a relationship can b e
developed.
T h e wavelet spectrum E(X) or Ε'(ω) discussed in the previous sec-
tion although interesting in its own right, takes us away from t h e non-
stationarity of the process since it is obtained by integrating over t. We,
therefore, need something else. This is provided by the Wigner-Ville spec-
t r u m . Let us define a general (non-stationary) covariance function R(t, s)
as
R(t,s)=S[X(t)X(s)].
T h e n t h e Wigner-Ville spectrum (WVS) is defined as (see [11] for a dis-
cussion of WVS and other time-frequency distributions)
oo
■R(t+^,t--)e-iurdT. (36)
/
The WVSx(t, ω) is an energy density function as
oo
WVSx(t,w)duj (37)
/
-co
i.e., we get the instantaneous energy by integrating over all frequencies,
and t h e total energy can be obtained as
(38)
16 P. Kumar and E. Foufoula-Georgiern
(39)
We, therefore, see t h a t there is an inherent link between the study of non-
stationary processes and wavelet transforms akin to the link between sta-
tionary processes and Fourier transforms.
§4. E x a m p l e s of O n e - D i m e n s i o n a l W a v e l e t s
Due to the flexibility in choosing a wavelet, several functions have been
used as wavelets and it would be difficult to provide an exhaustive list. We
present here some commonly used wavelets (Haar wavelet, Mexican h a t
wavelet, and Morlet wavelet) in one-dimensional applications.
(40)
4 . 2 . M e x i c a n hat w a v e l e t
T h e Mexican hat wavelet is the second derivative of t h e Gaussian e~l ' 2
given as (see Figure 7)
(41)
Wavelet Analysis in Geophysics: An Introduction 17
0.5l·
Oh
\y : . .\y
3
(a)
Spectrum of Morlet wavelet
0.8
0.6
0.2
-2 0 2 10
frequency
(b)
"Oiax *Ίηιη
(46)
§5. D i s c r e t e W a v e l e t Transforms
W h e n the parameters λ and t in the wavelet transform (/, φχ^) take on
continuous values, it is called continuous wavelet transform. For practical
applications the scale parameter λ and location parameter t need t o b e
discretized. One can choose λ = λ™ where m is an integer and λ 0 is a
fixed dilation step greater t h a n 1. Since σψχ = \σφχ , we can choose
t = ntoX™ where to > 0 and depends upon ip(t), and n is an integer.
T h e essential idea of this discretization can be understood by an analogy
with a microscope. We choose a magnification, i.e., A^"m, and study t h e
process at a particular location and then move to another location. If t h e
magnification is large, i.e., small scale, we move in small steps and vice-
versa. This can be easily accomplished by choosing t h e incremental step
inversely proportional to the magnification (i.e., proportional to t h e scale
20 P. Kumar and E. Foufoula-Georgiou
Modulus
Phase
For t h e purpose of this subsection, let ipm,n{t) denote the above discretiza-
tion rather t h a n the general discretization given by equation (47). We will
also use t h e identity ^oo(^) = Ψ(ί)· It is possible to construct a certain
class of wavelets ip(t) such t h a t ipm,n(t) are orthonormal, i.e.,
Φπι,η(^)Φτη',η'(ή dt = J m m / £ n n / (50)
/ ·
where 6{j is t h e Kronecker delta function given as
%J
\0 otherwise. ^ '
0m,n(i) = 2 - m / 2 0 ( 2 - m i - n ) (55)
and satisfying certain properties enumerated in appendix A, such t h a t t h e
first sum on the RHS of equation (54) can be written as a linear combination
of φηΐο,η (see [30]), i.e.,
oo oo oo
Consequently,
oo m0 oo
The scaling functions and wavelets play a profound role in the analysis of
processes using orthogonal wavelets. This analysis framework is known as
the wavelet multiresolution analysis framework and is discussed below. Ap-
pendix B describes a class of orthogonal wavelets developed by Daubechies
[13] and Appendix C briefly discusses the implementation algorithm by
Mallat [30].
(59)
(60)
(61)
(62)
5.2. N o n - o r t h o g o n a l w a v e l e t t r a n s f o r m s
5.2.1. Frames
We saw in section 5.1 t h a t it is possible to find λο, to and ipmin(t)
as defined in equation (47) such t h a t ψγη,η{ί) are orthogonal. This allows
a function f(t) to be written as a series expansion as given in equation
(53). However, even if ^ m , n ( 0 are not orthogonal, the function f(t) can b e
represented completely as a series expansion under certain broad conditions
on the wavelet r/>(i), t0 and λο· These discrete wavelets which provide
complete representation of the function f(t) are called wavelet frames and
will be the subject of the next sub-section. We will see t h a t orthogonal
wavelets are a special case of this general framework. Let us first define
frames.
m a
A sequence of functions {ψη}η£Ζ Hilbert space H (see footnote
on page 24 for definition of Hilbert Space) is called a frame if there exist
two constants A > 0, B < oo, called frame bounds, so t h a t for all functions
f(t) in the Hilbert space H the following holds:
A||/||2<£|</,v>„>l2<B|l/ll2· (66)
n
T h e constant B < oo guarantees t h a t the transformation / —> { ( / , φ η ) }
is continuous and the constant A > 0 guarantees t h a t this transformation
is invertible and has continuous a inverse. This enables one to: (1) com-
pletely characterize the function, and (2) reconstruct t h e function from its
26 P. Kumar and E. Foufoula-Georgiou
decomposition.
In general, a frame is not an orthonormal basis. It provides a redundant
representation of the function f(t). This is analogous, for example, t o
representing a vector in the Euclidean plane using more t h a n two basis
vectors. T h e ratio A/B is called the redundancy ratio or redundancy factor.
Redundant representations are more robust t o noise and therefore useful
when noise reduction is an issue.
W h e n A = B, the frame is called a tight frame. In this case there is a
simple expansion formula given as
Notice t h a t this formula is very similar t o the one obtained for a n or-
thonormal set {φη}· In this case, however, {φη} niay not even be linearly
independent, i.e., there is a large degree of redundancy in the represen-
tation. Orthonormal bases arise as a special case. For a tight frame, if
A = B = 1 and if || φη \\= 1, then {φη} form an orthonormal basis and we
get t h e usual expansion formula. When {^ m , n } constitute a tight frame
then A = B = C ^ / i o l o g ^ o where Οψ is defined in equation (15) (see [15],
equation 3.3.8). However, in practice it is difficult t o get A exactly equal
to £?, b u t easier t o get A close t o B, i.e., e = - j — 1 <£ 1. Daubechies (see
[14], pg. 971) calls such frames snug frames. The expansion formula in this
case is given as
/(ί) =
ΐΓβΣ(/^>«+^ (68)
n
where t h e error 7 is of the order of 2x7 || / ||· T h e general case of A 96 B
is more involved and beyond the scope of this introduction (see [15], for
details).
5 . 2 . 2 . W a v e l e t frames
Now let L denote the transformation L : f(t) -l· {(/,ψγη,η)}, where
tpm,n(t) is defined by equation (47). We can characterize the function
f(t) through the wavelet coefficients { ( / , ipm,n)} provided the transform L
satisfies the condition (66), i.e.,
when {ψπι,η} is a snug frame. Such frames can be constructed for certain
choices of λο and £o, provided ip(t) satisfies the admissibility condition,
i.e., J φ(ί) dt — 0, and has compact support or sufficiently fast decay. T h e
conditions for the choice of λο and to are described in Daubechies (see [15],
chapter 3). Here it suffices to say t h a t these conditions are fairly broad
and admit a very flexible range. For example, for the Mexican h a t wavelet
(as given in equation (41)), for λο = 2 and to = 1, the frame bounds are
A = 3.223 and B = 3.596 giving B/A = 1.116.
One can obtain B/A closer to 1 by choosing λο < 2. Grossmann et al.
[24] suggested decomposing each octave into several voices (as in music) by
choosing λο — 2 1 / / M where M indicates the number of voices per octave.
W i t h such a decomposition we get
ψ™η(ή = 2 - m / 2 A V ( 2 " m / M * - nto). (72)
For the Mexican h a t wavelet, by choosing M = 4 and to = 1 we can obtain
A = 13.586 and B = 13.690 giving B/A = 1.007. Such a decomposition
using such a multivoice frame enables us to cover the range of scales in
smaller steps giving a more "continuous" picture. For example, with M = 4
we get discrete scales at {λ = . . . , 1 , 2 1 / 4 , 2 1 / 2 , 2 3 / 4 , 2 , 2 5 / 4 , 2 3 / 2 , 2 7 / 4 , 4 , . . . }
as against {λ = . . . , 1 , 2 , 4 , . . . } for usual M = 1. Figure 9 was created using
Morlet wavelet with M = 4 and t0 = 1. For this decomposition A = 6.918,
B = 6.923 giving B/A = 1.0008. It should be noted t h a t Morlet wavelet,
which is not orthogonal, gives a good reconstruction under t h e framework
of equation (71). Multivoice frames are discussed extensively in Daubechies
([15], chapter 3) where more details on the values of A and B for different
choices of M and to are given for the mexican hat and the Morlet wavelet.
R e d u n d a n t representations such as the one presented above, in addition
to their noise reduction capability, are useful when representations t h a t are
close to the continuous case are sought (see for example [3, 32, 35, 5, 33]
and [49]).
5.3. B i o r t h o g o n a l w a v e l e t s
Under the wavelet multiresolution framework, the decomposition and
reconstruction of a function is done using the same wavelet, i.e.,
(see [15], theorem 8.1.4) t h a t the only real and compactly supported sym-
metric or antisymmetric wavelet under a multiresolution framework is t h e
Haar wavelet. In certain applications however, real symmetric wavelets
which are smoother and have better frequency localization t h a n t h e Haar
wavelet may b e needed. In such situations, biorthogonal wavelets come
to t h e rescue. It is possible t o construct two sets of wavelets {ipm^n} and
{Ψτη,η} such t h a t
= EE</^™.»>^.«w· (75)
m n
T h a t is, one can accomplish decomposition using one set of wavelets and
reconstruction using another. T h e wavelets ißmyn(t) = ^ 7 7 ^ ( 2 ^ — n ) a n ( ^
ψηι,η^) = 2^72^(2^ ~ n ) n e e d t 0 Satis
fy
• · · c v2 c Vi c v0 c y_i c K-2 c · · ·
• · · C V2 C Vi C Vo C VLi C V-2 C ■ ■ ■
with Vm — s p a n { 0 m j „ } and Vm = s p a n { 0 m > n } and t h e complementary
spaces Om = s p a n { ^ m > n } and O m = s p a n { ^ m > n } . T h e spaces Vm a n d
Om (Vm and O m , respectively) are not orthogonal complements in general.
Equation (78), however, implies t h a t
Vm JL Öm and Vm J- O m . (80)
Wavelet Analysis in Geophysics: An Introduction 29
§6. T w o - D i m e n s i o n a l W a v e l e t s
6.1. Continuous wavelets
T h e continuous analogue of wavelet transform (12) is obtained by treat-
ing u — (u 1,1*2) and t = (£1,^2) as vectors. Therefore for the two dimen-
sional case
λ>0
(81)
(82)
2. fft{t)dt = 0.
Two examples of two-dimensional wavelets are discussed in the follow-
ing subsection.
(83)
(84)
1
^ t a n - ^ . (85)
V^ J — OO J — OO
Vm = t£®V£ (89)
where 0 represents a tensor product. It, therefore, follows (by expanding
y m + i as in (89) and using property M l ) t h a t the orthogonal complement
0m of Vm in Vm+\ consists of the direct sum of three subspaces, i.e.,
Om = {VL®Oln)®{Ol®V^)®{0]n®0]n). (90)
i V ^ m n / î ) ^ Trink") ^ m n k ) ( n k)£%2 J
Ä 7 = {(/,*L,*)(n,fc)<Ez2}> (93)
2
Ä / = {(/,*L,*)(„,fc)€z*K (94)
and
Q £ / = {(/,*3m„*)(n,fc)€Z'}· (95)
T h e corresponding continuous approximation will be denoted by Qmf(t),
Qmf(t) and Qmf(t) respectively. For implementation to discrete d a t a see
[31]·
T h e decomposition of Om into the sum of three subspaces (see equa-
tion (90)) acts like spatially oriented frequency channels. Assume t h a t we
have a discrete process at some resolution m + 1 whose frequency domain
is shown in Figure 11 as the domain of Pm+if. W h e n t h e same process
Wavelet Analysis in Geophysics: An Introduction 33
Frequency support of Q mf
Λ
Λ
3 1 Λ
3
Ψ (ω)
Λ Λ
2 2
Ψζ(ω)
Frequency support of P f
Λ Λ
3 3
Ψ3(ω) Ψ (ω)
Frequency support of P m f
.Frequency support of Q mf
ω = ( œt , cot )
1 2
§7. C o n c l u s i o n s
Wavelet theory involves representing general functions in terms of sim-
pler building blocks at different scales and positions. Wavelets offer a ver-
satile and sophisticated tool yet simple to implement, and have already
found several applications in a wide range of scientific fields. T h e list of
34 P. Kumar and E. Foufoula-Georgiou
Table I
Anonymous ftp site information for wavelet related software.
1 FTP site Directory References/Notes f
cs.nyu.edu pub/wave [35, 33], C programs 1
playfair.stanford.edu pub/software/wavelets MATLAB Scripts
simplicity.stanford.edu /pub/taswell MATLAB scripts
gdr.bath.ac.uk /pub/masgpn S software
pascal.math.yale.edu /pub/software/xwpl Wavelet Packet
Laboratory
[email protected] e-mail request for KHOROS
cml.rice.edu /pub/dsp/software MATLAB scripts
1 wuarchive.wustl.edu edu/math/msdos/modelling C programs
wavelet applications increases at a fast pace and includes to date signal pro-
cessing, coding, fractals, statistics, image processing, astrophysics, physics,
turbulence, mathematics, numerical analysis, economics, medical research,
target detection, industrial applications, quantum mechanics, geophysics
etc. (see, for example, the edited volumes by Chui [7], Ruskai et al. [48],
Benedetto and Frazier [3], Farge et al. [20], Meyer and Roques [41], Combes
et al. [9], and Beylkin et al. [2], among others). A general literature survey
on wavelets can be found in [46]. In geophysics, significant progress has
already been m a d e in studying and unraveling structure of several geophys-
ical processes using wavelets (see the extended bibliography of wavelets in
geophysics at t h e end of this volume). We have no doubt t h a t the study of
geophysical phenomena, which are by nature complex, and take place and
interact at a range of scales of interest, will continue to benefit from t h e
use of the powerful and versatile tools t h a t wavelet analysis has to offer.
There are a number of useful sources of information about wavelet pub-
lications and computer software for the implementation of wavelet analysis.
An electronic information service called "wavelet digest" exists (at the time
of this writing) on the Internet with the address waveletumath.scarolina.edu.
Several anonymous ftp sites exist on the Internet from where software for
wavelet analysis can be obtained. In Table I we provide a brief list (known
to t h e authors at the time this was written) solely for the purpose of infor-
mation to readers, without any recommendations, or reference to suitability
or correctness of these codes.
A . P r o p e r t i e s of Scaling F u n c t i o n
T h e scaling function satisfies the following properties:
5. (j)(t) = Ση ϊι(η)φ(2ί — η), i.e., the scaling function at some scale can
b e obtained as a linear combination of itself at the next scale (h(n)
are some coefficients called the scaling coefficients). This is a two-
scale difference equation (see [16] and [17] for a detailed t r e a t m e n t
of such equations).
For the particular case of Haar wavelet (see equation (40)) and t h e corre-
sponding scaling function (equation (58)) h(0) = h(l) = 1 and h{n) = 0
for all other n, and g(l) = —g(0) = 1 and g(n) = 0 for all other n.
B. Daubechies' Wavelets
Daubechies [13] developed a class of compactly supported scaling func-
tions and wavelets denoted as ( ^ , ρ/ψ). They were obtained through t h e
solution of the following two-scale difference equations:
2N-1
φ(ή = y/2 Σ Κη)Φ{2ί - n) (Β·ΐ)
2ΛΓ-1
3. T h e constraints
27V-1
10 20 30 40
Frequency (Radians)
10 20 30 40
Frequency (Radians)
C . I m p l e m e n t a t i o n A l g o r i t h m for O r t h o g o n a l W a v e l e t s
T h e implementation algorithm for wavelet multiresolution transforms
is simple. From a d a t a sequence {c^} (say at resolution level m = 0)
corresponding to a function f(t) we construct
ci = 5>(»-2*)<$. (C.2)
38 P. Kumar and E. Foufoula-Georgiou
10 20 30 40
Frequency (Radians)
10 20 30 40
Frequency (Radians)
= 1S
T h e detail sequence { ^ } f c G z {(/'^i^)}fceZ obtained as
d\ = Y,g{n-2k)cl (C.3)
Equivalently the above two equations can be written in the matrix notation
The algorithm given in equation (C.2) and (C.3) can be recursively im-
plemented and the d a t a and details at lower and lower resolutions can be
obtained. This also highlights another important feature. Once we have
the coefficients h{n), we never need to construct the scale function 0(f) or
the wavelet φ(ί) for implementation on a discrete d a t a set. T h e assump-
tion involved, however, is t h a t c^ = f f(t)(f)(t — n)dt, i.e., the integration
kernel is φ(ί) corresponding to the chosen /i(n)'s. T h e reconstruction of
the original sequence can be achieved by using
and is exact.
References
1. Arneodo, A., G. Grasseau and M. Holschneider, Wavelet transform of
multifractals, Physical Rev. Let, 61(20), 2281-2284, 1988.
2. Beylkin, G., R. R. Coifman, I. Daubechies, S. G. Mallat, Y. Meyer,
L. A. Raphael, M. B. Ruskai (eds.), Wavelets and Their Applications,
Jones and Bartlett Publishers, Boston, 1991.
3. Benedetto J. J., and M. W. Frazier (eds.), Wavelets: Mathematics and
Applications, CRC Press, Boca Raton, Florida, 1993.
4. Brunet, Y. and S. Collineau, Wavelet analysis of diurnal and nocturnal
turbulence above a maize crop, This Volume, 1994.
5. Burt, P . J., Fast filter transforms for image processing, Computer
Graphics and Image Processing, Vol. 16, 20-51, 1981.
6. Chui, C. K., An Introduction to Wavelets, Wavelet Analysis and its
Applications, Vol. 1, Academic Press, New York, 1992.
7. Chui, C. K., Wavelets - a Tutorial in Theory and Applications, Wavelet
Analysis and its Applications, Vol. 2, Academic Press, New York, 1992.
8. Coifman, R. R., Y. Meyer and V. Wickerhauser, Size properties of
wavelet-packets, in Ruskai et al., 453-470, 1992.
9. Combes, J . M., A. Grossman, P h . Tchamitchian (eds.), Wavelets:
Time-Frequency Methods and Phase-Space, Proc. of the Int. Conf.,
Marseille, Frace, Decemberl4-18, 1987, Springer-Verlag, 1989.
10. Cohen, A., Non-separable bidimensional wavelet bases, Revista
Mathematica Iberoamericana, 9(1), 51-137, 1993.
11. Cohen, L., Time-frequency distributions-a review, Proc. of the IEEE,
77(7), 941-981, 1989.
40 P. Kumar and E. Foufoula-Georgiou
The first author would like to thank Mike Jasinski and the Hydrologie
Sciences Branch at NASA-Goddard Space Flight Center for the support
during the course of this work. The second author would like to acknowl-
edge the support of National Science Foundation grants BSC-8957469 and
EAR-9117866, and NASA grant NAG 5-2108. We would also like to ac-
knowledge the thoughtful comments of Don Per ci val, Naoki Saito and Paul
Liu on this chapter.
Praveen Kumar
Universities Space Research Association/
Hydrological Sciences Branch (Code 974)
NASA-Goddard Space Flight Center
Greenbelt, MD 20771
[email protected]
Wavelet Analysis in Geophysics: An Introduction 43
Efi Foufoula-Georgiou
St. Anthony Falls Hydraulic Laboratory
Department of Civil and Mineral Engineering
University of Minnesota
Minneapolis, Minnesota 55414
e-mail: efiUmykonos.safhl.umn.edu
Applications of Structure Preserving Wavelet Decompositions
to Intermittent Turbulence: A Case Study
Wavelets in Geophysics 45
Efi Foufoula-Georgiou and Praveen Kumar (eds.), pp. 45-80.
Copyright 1994 by Academic Press, Inc.
All rights of reproduction in any form reserved.
ISBN 0-12-262850-0
46 C. Hagelberg and N. Gamage
§1. Introduction
Various physical processes in the atmospheric boundary layer such as
the vertical transport of momentum and heat are associated with intermit-
tent, coherent events, such as convective updrafts and plumes occurring
due to localized heating and the resulting convection [18, 20]. Character-
izing the physics of these events through observations and their behavior
under varying conditions is still a topic of considerable continued study
[1, 3]. There is a need to verify existing models (conceptual, mathemati-
cal or numerical) with observational data, and the need to provide models
with parameter estimates for realistic simulation and continued insight into
the physics to further develop conceptual models. Additionally, there is a
need to determine from observation a set of parameters which best de-
scribe the processes associated with intermittent events. Identification of
such parameters will suggest new approaches to modeling.
Conditional sampling techniques have been extensively reported in
the literature as a means of extracting coherent events from data records
[1, 3, 17, 20, 24]. Typically, conditional sampling techniques extract co-
herent events from a record to form an ensemble. The ensemble is then
analyzed to characterize the events. This process loses important inter-
mittency information since the global placement of the events is neglected.
Various approaches to the sampling method have been used including "di-
rect" thresholding on the values of scalars [18], thresholding on computed
variance over short record patches [24, 25], visual identification of time do-
main characteristics [3], and thresholding time domain gradients [17]. In
the latter case, an indicator record is constructed applying a threshold on,
for example, the absolute value of the time derivative of velocity. This al-
lows the definition of intermittency as a function of threshold which is 0 if
no events are selected and 1 if the whole record is selected. This technique
does not attempt to address the issue of coherence. More recently, con-
ditional sampling methods have been developed based on variants of the
wavelet transform [4, 9, 20]. The techniques proposed and demonstrated
here, based on the wavelet transform, provide a general and objective means
for achieving the desirable characteristics of conditional sampling methods
and also provide a measure of intermittency which is directly related to
coherent events.
The wavelet transform has evolved in various disciplines which require
analysis of signals exhibiting intermittency in various forms. We attempt
to illustrate some properties of the wavelet transform and propose certain
new ideas useful to the study of atmospheric turbulence and the planetary
boundary layer.
Sharp edges contain significant energy at high wave number. A signal
consisting of intermittent sharp edges can have varying spectral charac-
Intermittent Turbulence 47
temperature, and (derived) buoyancy flux density fields obtained from the
atmospheric boundary layer are studied for two cases corresponding to late
morning and late afternoon on the same day. We compute the intermit-
tency index for each of the records. We then construct partition of the
signals and compute total buoyancy flux for various combinations of the
components. Additionally, spectra for the velocities and their decomposi-
tions are studied. A summary is provided in Section 5.
ll/l|2 = ( / , / ) 1 / 2 .
Similarly, the convolution of / and g is
/-(«):=/(-*)· (6)
2
A basic w a v e l e t is an L (H) function, xp{t), which satisfies the fol-
lowing admissibility criteria [5, 7]:
Equivalently,
W.f(b):=f*t/>;{b). (8)
2
T h e continuous wavelet transform of an L ( R ) function is a function
of scale, s φ 0, and location, 6, and as such forms a surface over t h e (5, b)-
plane called the phase plane of the wavelet transform.
Note t h a t we have chosen a particular scaling for our definition of t h e
wavelet transform in equation (7) through equation (4). Other choices are
possible [2, 10], however, it has been argued t h a t the choice provided by
equation (7) facilitates the interpretation of the wavelet transform as a
measure of coherence of structures with respect to scale [10]. We note t h a t
[10] a t t e m p t e d to distinguish between scaling choices of yfs and s in t h e
definition of the wavelet transform by calling scaling by yfs the wavelet
transform and scaling by s the covariance transform. It appears t h a t t h e
emerging convention in the literature is to disregard this distinction by
using t h e n a m e "wavelet transform" for any choice of scaling, and we adopt
this point of view, making distinctions where necessary. Using the scaling
yfs leads to an energy preserving transform, t h a t is, a type of Parseval's
50 C. Hagelberg and N. Gamage
relation for wavelets. While this is attractive for interpreting the wavelet
transform as a decomposition of variance, it obscures the interpretation for
detecting coherent events.
In the notation of equation (7) the wavelet transform may b e inter-
preted as a covariance between the wavelet at a given dilation, φ8 and t h e
signal, / , at a specified location in the signal [10]. Using the notation of
equation (8) the wavelet transform may also be interpreted as t h e signal,
/ , filtered by a particular (non-uniform, band pass) filter, φ8 [6, 21]. We
will discuss the application of the wavelet transform to the construction of
filters and the filtering properties of some wavelets in Section 3.
T h e continuous wavelet transform may be inverted to recover the signal
as follows.
(9)
2 . 2 . D y a d i c s u b s a m p l i n g in scale
We restrict the scale parameter, s to powers of 2. This allows the
implementation of the fast wavelet transform algorithm given by [22].
Thus, let s = 2·7 for j any integer. For this restriction to yield a
usefully invertible transform, the regularity condition (ii) must be modified
as follows. T h e function φ G £ 2 ( R ) (which satisfies the zero mean condition
(i)) is called a dyadic wavelet if there exist positive constants 0 < A < B <
oo such t h a t
(10)
which is called the stability condition for the dyadic wavelet [5, 22]. It has
been shown t h a t a dyadic wavelet also satisfies the regularity condition (ii)
[5].
(11)
Σ φ·@ω)χ@ω) = 1 . (15)
j=—oo
= Σ Σ i^wi2, (is)
j= — oo k= — oo
EJ= Σ i ^ i 2 · (19)
k= — oo
2 . 3 . F i n i t e r e s o l u t i o n in d i s c r e t e a p p l i c a t i o n s
Since discrete d a t a are limited to finite resolution (smallest scale) de-
termined by a sampling rate it is not possible to perform an analysis at a
resolution finer t h a n some fixed scale. Choosing t h e smallest scale t o b e
unity (s = 23\ j = 0) means we cannot represent d a t a on scales between
grid points. Actual d a t a are also limited to some finite record length, say
s = V for j = J > 1. Furthermore, any definitions of functions on t h e finite
grid may fail to follow the continuous properties in the conditions given by
equations (10), (14), and (15). This is discussed in detail in [22]. T h e
solution is to introduce a smoothing function, φ, whose frequency content
represents the fractional loss of energy when performing a wavelet trans-
form followed by the inverse transform (under the restrictions imposed by
finite resolution). T h a t is,
oo
φ(ω) = Σψ(2*ω)χ(νω).
5i/(i):=/*^(*), (21)
3 3
where φ^(ί) = 2~ φ^2~ ). T h e information content of the signal smoothed
at some coarse scale, Sjf{t), and the wavelet coefficients for all scales up to
the coarse scale, {Wj/(£)}i<j<./5 is sufficient to recover the original discrete
signal. We accomplish this using the fast wavelet transform algorithm
described in [22].
As noted, the orthonormal wavelet coefficients may be obtained from
the continuous dyadic wavelet coefficients through appropriate subsampling
and scaling. For real applications the d a t a record length is finite and an
orthogonal subsampling depends on which d a t a point is assigned as t h e
end of record. Since the wavelet transform is not shift invariant, a shift
in the location at which the wavelet coefficients are computed results in a
change in t h e wavelet coefficients, and hence a change in the estimate of
t h e spectrum resulting from the coefficients. By averaging over all possible
orthogonal subsamplings of the translation parameter we approach a shift
invariant estimate of the spectrum. An equivalent approach would be to
average over several globally translated orthogonal wavelet analyses.
54 C. Hagelberg and N. Gamage
§3. Filtering P r o p e r t i e s of W a v e l e t s
3 . 1 . P a r t i t i o n s as filters
We wish to partition a signal into two components, one containing
significant structures and the other containing the remaining portion of
the signal. Once a partition is obtained, further analysis of each compo-
nent can proceed. For example, an intermittency index can be assigned
to the structure-containing component, and spectral characteristics of t h e
components can be analyzed. In certain communication theory applica-
tions t h e significant structures may correspond to a signal prior to being
transmitted, while the remaining portion of the signal corresponds to t h e
noise introduced during transmission. In the case of atmospheric turbu-
lence, structures may be defined as regions of sharp transition of significant
amplitude [10, 20], while the remaining portion corresponds to a different
physical flow characteristic (not necessarily regarded as "noise").
To create a partition of a signal one must begin with assumptions
regarding t h e n a t u r e of the signal. For example, Fourier (band-pass) filter-
ing partitions a signal into components based on frequency content. Some
criterion is used to determine the components, such as the location of a fre-
quency band or the location of multiple bands determined by a frequency
amplitude cutoff. T h e partition then consists of the portion of the signal
with frequency content inside the bands and the portion with frequency
content outside the bands. A filter consists of retaining only one element
of the partition. If the significant structures are band limited and t h e
remaining portion of the signal is outside the structure containing band,
then the signal will exhibit spectral gaps. T h e b a n d - p a s s filters can then
be prescribed according to the location of the spectral gaps.
However, if the structure of interest in the signal is a sharp transition
it contains significant high frequency content. T h e high frequency content
can not b e distinguished from the frequency contribution of several smaller
amplitude sharp transitions, or from colored or white noise. In this case
the component of interest in the signal is not usefully band limited and t h e
spectrum will not exhibit a clear spectral gap. A Fourier filtering technique
is not well suited to creating a partition of the signal in such a case. T h e
wavelet transform can help in this situation since it effectively limits t h e
Intermittent Turbulence 55
different frequencies will produce two local maxima in the wavelet variance
[4]. In such a situation, each local maximum corresponds to a wavelength
and some choice as to their significance must be based on knowledge of t h e
n a t u r e of the signal.
Once the dominant event scale or scales are found, one of the obvious
methods leading to a partition of the signal is to limit the reconstruction
from the phase plane to the scales above and below some scale near t h e
dominant scale, or between two dominant scales. One approach to deter-
mining a partition if only one dominant scale is present is to reconstruct
from t h e phase plane information using the large scales down to some small
scale which is determined from the dominant scale by taking an appropri-
ate fraction of t h e dominant scale. The partition then consists of the large
scale reconstruction and the small scale reconstruction (that is, the remain-
ing portion of t h e signal). If there is a clear "scale gap" between peaks in
the wavelet variance, the scale for the partition could be chosen to be in
the gap [4]. We refer to this as scale threshold partitioning. Scale thresh-
old partitioning has the problem of neglecting the fact t h a t structures of
interest are multiscaled and may have significant information content even
at small scales. T h a t is, portions of the information content of a single
structure may fall into b o t h elements of the partition. This is an inherent
problem in partitioning signals in which multiscale structures are impor-
t a n t b u t are found in the presence of broad band noise. Another inherent
problem is determining the appropriate fraction of the dominant scale for
the threshold. Examples are given later in this section and in Section 4.
Even with these difficulties, a scale threshold partition works very well for
separating out white noise at the smallest resolvable scale.
In [23] the rate of growth or decay of significant "ridges" in the phase
plane is used as a means of quantifying what is meant by noise. This is
a measure of the regularity of a multiscale feature. For highly singular
features the wavelet transform will undergo a rapid increase in value at
small scales as the scale is decreased. Assuming t h a t the "noise" is less
regular t h a n the structures of interest and limited to small scales, a ridge
reconstruction ([23, 22]) is performed using only the portions of ridges
which do not exhibit an increase at small scales, or by extending ridges to
smaller scales keeping the amplitude fixed.
In [23] the use of ridge threshold filtering described above is motivated
by the interest in signal compression. For this study we have no need for
signal compression and therefore have the information of the entire phase
plane at our disposal. We can therefore employ a simpler approach to
partitioning which still allows for the multiscale nature of coherent events.
Furthermore, for applications to atmospheric turbulence data, we are inter-
ested in partitioning the signal rather t h a n an actual filtering (that is, we
retain b o t h components of the partition). T h e method we examine is t h e
Intermittent Turbulence 57
3 . 2 . D e t e c t i o n characteristics of anti—symmetric a n d s y m m e t r i c
wavelets
In this section we demonstrate certain general characteristics of t h e
differences between using symmetric versus anti-symmetric wavelets for
signal analysis. Figure l a is the vertical velocity record from an aircraft
measurement during the F I F E 87 field experiment (described in Section 4).
T h e signal contains structures resembling square or r a m p e d pulses on t h e
order of 10 seconds wide (40 meters). There are approximately eight or nine
such events. In order to illustrate the response of the wavelet transform to
such structures, consider the simulated signal of Figure 2. This consists of
four "pulses" of similar shape, unevenly spaced, with a resolution of 256
points. T h e shape was chosen to have some characteristics typical of t h e
structures observed in the F I F E 87 vertical velocity record.
T h e subsequent analyses are performed using the anti-symmetric quad-
ratic spline wavelet and the symmetric cubic spline wavelet shown in Fig-
ure 3. T h e wavelet transform of the simulated signal based on a symmetric
wavelet and an anti-symmetric wavelet at a single scale are shown in Fig-
ure 4. Note t h e magnitude of the symmetric transform is large at t h e
boundaries of t h e transitions, while the magnitude of t h e anti-symmetric
58 C. Hagelberg and N. Gamage
600
400
out white noise, since it has a decorrelation length scale of two points. Scale
partitioning has been suggested as a method of filtering in [4].
Finally, the phase plane threshold and the scale threshold may be com-
bined to create a partition such as t h a t shown in Figure 11. We have
combined a phase plane threshold of a factor of 0.2 (also emperically deter-
mined as above) of the largest of the local extremes with a scale threshold
of 2 1 .
-i r
§4. A p p l i c a t i o n s t o A t m o s p h e r i c D a t a
4 . 1 . D e s c r i p t i o n of t h e d a t a
We now apply the techniques described in Sections 2 and 3 to two
portions of d a t a collected during the First International Satellite Land
Surface Climatology Project (ISLSCP) Field Experiment 87 (FIFE87). A
description of t h e experiment and analysis of d a t a using other techniques
including conditional sampling may be found in the F I F E special issue of
J G R [15]. T h e purpose of this section is to demonstrate t h e utility of t h e
techniques presented in the previous sections by providing new insight into
t h e interpretation of specific data. This is an extension of t h e results shown
in [12] where a different segment of the d a t a was analyzed.
An objective of FIFE87 was to gather in situ boundary layer measure-
68 C. Hagelberg and N. Gamage
4 . 2 . A n a l y s i s of t h e d a t a
Table 1 collects some length scale statistics computed using an a n t i -
symmetric wavelet transform for vertical velocity components and the vir-
tual potential t e m p e r a t u r e for the morning and afternoon legs. Listed are
the dominant scale lengths (as derived from the largest local maximum in
the wavelet variance), the number of occurrences of events of this size, and
t h e intermittency index as defined in Section 3c. Due to large amplitude
low frequency components in the temperature records, the largest variance
for an anti-symmetric wavelet will occur at the largest scale. We therefore
Intermittent Turbulence 69
A combination phase plane threshold and scale partition has been cre-
ated for the velocity components (Figure 12a and c) and the buoyancy flux
densities (Figure 12b and d) for the morning and afternoon cases. Combi-
nation partitions were also created for the temperature records but are not
shown. An anti-symmetric wavelet analysis was used in each case. The
sharp vertical edges have been maintained in the larger scale partition in
each case.
Total buoyancy flux over the signal was computed by summing over
the flux density for each of five possibilities - from the original data, νϋ'θ'υ,
which will be the reference flux; from the structure partition of the derived
buoyancy flux density record, (u/0(,) str ; from the structure component of
the velocity with the original temperature data, u)'stx9'v; from the structure
component of the temperature with the original velocity data, t i / ^ s t r ) ' ;
and from the structure component of velocity with the structure component
of temperature, (wstTY(6v stTy. This was done for the morning case and
the afternoon case resulting in the ten numbers shown in Table 3 where the
total fluxes are shown as a percentage of the reference flux. The morning
reference flux was about 40% of the afternoon reference flux.
The flux transport computed using the structure component of velocity,
(u;str )'#(,, accounts for 69% of the reference flux in the morning and 58% of
the reference flux in the afternoon. The flux transport computed using the
structure component of temperature, u)f(6VstJ.y, is relatively higher in both
the morning (82%) and the afternoon (75%). Additionally, the total flux
computed using the structure component of the velocity with the structure
component of the temperature, (wstTy(6V8tT)\ is 58% in the morning and
42% in the afternoon indicating that the temperature structures are not
highly correlated with the velocity structures.
2 h
> o h
1 I — i 1 — i 1 — I 1 r η — i 1 — i 1 1 1 — i — i 1 p
Figure 12. Morning case: Combination phase plane and scale parti-
tions for (a) (top) vertical velocity and (b) (bottom) buoyancy flux
density, Vertical shifts have been added to offset the components.
72 C. Hagelberg and N. Gamage
~\—i—|—i—i—i—i—i—i—i—i—i—i—i—i—i—i—r
_J I i i i i I i i i i L
Time (s)
Figure 12. Afternoon case: Combination phase plane and scale par-
titions for (c) (top) vertical velocity and (d) (bottom) buoyancy flux
density, Vertical shifts have been added to offset the components.
Intermittent Turbulence 73
§5. S u m m a r y a n d C o n c l u s i o n s
In this paper we have utilized the wavelet transform as a time (or space)
series analysis tool for the analysis of d a t a containing intermittent coherent
events. We use a non-orthogonal formulation of the wavelet transform, and
choose scaling to emphasize the edge and singularity detection capabilities
of the transform. T h e fast algorithm of [22] is implemented for this purpose.
We use local maxima of the wavelet variance (defined in Section 2) as
a means for identifying important scales in the data. This approach has
been used in other studies as well [10, 4]. It is important to note t h a t
this is a reduction of the wavelet phase plane and is subject to the bias of
non-characteristic events. For example, a single event with a large ampli-
t u d e may determine a local maximum in t h e wavelet variance when several
events of a differing scale may contain the desired structural information.
Using the wavelet transform for detecting coherent events is consid-
ered in Section 3. T h e locations of local extrema of an anti-symmetric
wavelet transform correspond t o locations of centers of large gradients in
the signal. Similarly, the locations of local extrema of a symmetric wavelet
transform correspond to locations of large curvature in the signal. Alterna-
tively, t h e zero crossings of the symmetric wavelet transform correspond to
locations of centers of gradients in the signal, b u t t h e magnitude (sharpness
of the gradient) is then only determined by the slope of t h e zero crossing.
Each type of detection is necessarily scale dependent which must b e consid-
ered when a t t e m p t i n g to identify structures. A crucial choice for structure
detection algorithms is between anti-symmetric and symmetric analyzing
wavelets. T h e specific type of wavelet, and its regularity are less critical in
applications to real data. This is due in part to the limitations imposed by
Intermittent Turbulence 75
i 1 i i i i
103 101
Time (s)
| I I I I—I 1 Γ I i i i i—i—i r
M i l I I I L
10j 10z
wavelength (m)
Figure 13. Spectral variance density derived from the vertical velocity
component for (a) (top) the morning case, and (b) (bottom) the
afternoon case. The spectra have been multiplied by k5'3 so that a
—5/3 region will appear horizontal in the figure.
76 C. Hagelberg and N. Gamage
lo- 1 L
"Π—I 1 1 Γ
I I IΊ I L· J I 1 L·
10J 10z
wavelength(m)
Figure 14. Spectral variance density of structure (thick line) and non-
structure (thin line) components of w-velocity shown in Figure 12a and
c. (a) (top) Morning case, and (b) (bottom) afternoon case.
Intermittent Turbulence 77
References
1. Antonia, R. A., A. J. Chambers, C. A. Friehe, and C. W. Van A t t a ,
1979: Temperature ramps in the atmospheric surface layer. Journal of
the Atmospheric Sciences, 36, N o . l , 99-108.
2. Arneodo, A., G. Grasseau, and M. Holschneider, 1989: Wavelet Trans-
form Analysis of Invariant Measures of Some Dynamical Systems. In
Wavelets (Combes, J. M., A. Grossman, and P h . Tchamitchian, Eds.),
Springer-Verlag, 315 pg.
3. Bergstrom, H. and U. Hogstrom, 1989: Turbulent exchange above a
pine forest II. Organized structures. Boundary-Layer Meteorology 4 9
231-263.
4. Collineau, S., and Y. Brunet, 1993: Detection of turbulent coherent
motions in a forest canopy; P a r t I: wavelet analysis. Boundary-Layer
Meteorology 6 5 357-379.
5. Chui, C. K., 1992: An introduction to wavelets. Academic Press, Inc.
266 pg.
6. Daubechies, I., 1988: Othonormal bases of compactly supported
wavelets. Communications in Pure and Applied Mathematics, 4 1 , 9 0 9 -
996.
7. Daubechies, I., 1992: Ten lectures on wavelets. C B M S - N S F Regional
Conference Series in Applied Mathematics, Soc. for Industrial and
Appl. M a t h . , 357 pg.
8. Gamage, N. K. K., 1989: Modeling and analysis of geophysical turbu-
lence: use of optimal transforms and basis sets. P h . D . Thesis, Oregon
Intermittent Turbulence 79
§1. Introduction
According to the Kolmogorov theory [25] (hereafter referred to as K41),
the ensemble average of the n t h order velocity difference (Δΐί;) between two
points separated by spatial distance (r), in the inertial subrange, for high
Reynolds number flow is given by
<|Δ«,·η = ΛΓ η «6»*Γ* (1)
Wavelets in Geophysics 81
En* Foufoula-Georgiern and Praveen Kumar (eds.), pp. 81-105.
Copyright 1994 by Academic Press, Inc.
All rights of reproduction in any form reserved.
ISBN 0-12-262850-0
82 G. Katul et al.
and Ui are the velocity components (i — 1,2,3), n is the order of the struc-
ture function, v is the kinematic viscosity, Kn is a universal constant inde-
pendent of the flow but dependent on n, r is the separation distance that
is much smaller than the energy containing length scale or integral length
scale (L) but much larger than the Kolmogorov microscale η(= [^/3/(e)]1/4),
and (·) is the ensemble averaging operator. The scaling laws of Equation (1)
have been confirmed by many experiments for n = 2 (e.g. the existence of
the 2/3 law for the structure function or —5/3 law for the power spectrum)
as originally proposed by Kolmogorov [25] and discussed in Monin and Ya-
glom ([42], pp. 453-527). However, the scaling laws in Equation (1) are
not accurate for n > 2, as evidenced by many other experiments (see e.g.
[1]). Deviations from these scaling laws have classically been attributed
to the intermittency in e. This intermittency results in an e(x) behavior
that resembles an on-off process. That is, the dissipation of turbulent ki-
netic energy occurs only in a small fraction of the fluid volume. Hence, the
breakdown of Equation (1) is attributed to the inequality between (e n ) and
(e) n , as noted by Landau (see footnote in [29]; p. 126). As a result, many
phenomenological models for intermittency corrections to K41 have been
proposed. Example phenomenological models include the lognormal model
[26], the ß-model [17], and other multifractal models ([40], [39] and [2]).
Many atmospheric surface layer (ASL) flows exhibit an inertial sub-
range that extends over many decades so that intermittency effects on
inertial subrange scaling becomes important (see [22]; [42], Ch.8). Refined
intermittency studies in the natural environment encounter difficulties due
to: 1) the limited sampling period over which steady state mean mete-
orological conditions exist, 2) the need for instrumentation that is free
of atmospheric contamination and possible calibration drifts due to tem-
perature and humidity changes, 3) the need for instrumentation that is
field robust and capable of providing all three velocity components (since
changes in wind direction are inevitable), and 4) the need for turbulence
conditions that allow the application of Taylor's frozen hypothesis with
minimum wavenumber distortion.
The first difficulty severely limits the number of data points that can
be used to evaluate the ensemble average in Equation (1). Typically, the
ergodic hypothesis is used to evaluate the ensemble average in Equation (1)
from the measured time averages. The convergence of the time average and
the ensemble average requires a very large number of measurements that
may not be available in many field studies due to unsteadiness in the mean
meteorological conditions. The second and third difficulties limit the use
Atmospheric Surface Layer Turbulence 83
of many laboratory fast response sensors such as hot wire probes t h a t are
capable of resolving scales as small as t h e dissipation scales b u t difficult
to operate for long periods in the natural environment. We note here t h a t
some success in using such instruments (e.g. triaxial hot wire probes) for
extended periods were reported [43]. However, hot wire probes are not
suited for ASL measurements in arid and semi-arid environments since t h e
air t e m p e r a t u r e fluctuation can be very large (up to 6°C in seconds).
T h e development of analyzing tools t h a t allow the study of intermit-
tency effects in t h e ASL from limited number of field measurements is
necessary. T h e purpose of this paper is to investigate t h e usefulness of or-
thonormal wavelet transforms to quantify intermittency effects on inertial
subrange scaling from ASL velocity measurements. For this purpose, we
develop a conditional sampling scheme t h a t is capable of identifying dissi-
pation events in the space-scale wavelet domain. This conditional sampling
scheme can identify the location of large dissipation events t h a t contribute
to inertial subrange intermittency.
T h e wavelet transform is applied to 56 Hz triaxial sonic anemometer
velocity measurements in the ASL t h a t exhibit an inertial subrange for
three decades. Since intermittency investigations typically utilize Fourier
power spectra and structure functions, we first establish a relation between
the wavelet coefficients and these statistical measures. Then, we introduce
conditional wavelet statistics t h a t are developed to isolate events responsi-
ble for deviations from K41. However, before we discuss these approaches,
we offer a brief review of wavelet transforms with emphasis on applications
to turbulence measurements.
§2. A n a l y s i s of T u r b u l e n c e using W a v e l e t T r a n s f o r m s
Wavelet transforms are recent mathematical tools t h a t can unfold tur-
bulence signals into space and scale [15]. Continuous wavelet transforms
have been applied to many turbulence measurements and proved t o be
successful in identifying local scaling exponents ([4], [14] and [3]), inter-
mittency visualization [30], and coherent motion in ASL flows ([10], [19],
[18] and [33]). Orthonormal wavelets are a discrete form of t h e continu-
ous wavelets; however, they have the added feature of forming a complete
basis with t h e analyzing wavelet functions orthogonal to their translates
([41]; [12]; [9], Chapter 1). T h e application of orthonormal wavelets has
yielded i m p o r t a n t new techniques in the study of turbulence ([53], [54],
[55], [37], [38] and [23]). For completeness, a brief review of continuous and
orthonormal wavelet transforms is given.
Analogous to Fourier transforms, wavelet transforms can be classified
as either continuous or discrete. T h e continuous wavelet transform is first
introduced followed by a motivation for using discrete wavelet transform.
84 G. Katul et al.
f
'—OO
r+oo
rjj(y)dy = 0. (6)
(8)
For this basis function, the wavelet coefficients WT^m+l\k) and t h e coarse
grained signal 5 ^ m + 1 ^ (k) at scale ra + 1 can be determined from t h e signal
S^m) at scale m by using
(9)
(10)
number of samples (integer power of 2) (see [5] and [6]). For t h e Haar
wavelet, the coarse grained signal defined by Equation (10) is a low-pass
filtered function obtained by a simple block moving average, while t h e
wavelet coefficients are computed from the high-pass filter in Equation (9).
T h e wavelet coefficients and coarse grained signal may be calculated by t h e
following pyramidal algorithm:
§3. E x p e r i m e n t
T h e d a t a presented here were collected during an experiment carried
out on J u n e 27, 1993 over a uniform dry lakebed (Owens lake) in Owens
valley, California. The lakebed is contained in a large basin bounded by
the Sierra Nevada range to the east and the White and Inyo Mountains
to the west. T h e instrumentation site is located on the northeast end
of the lakebed (elevation=l,100 m ) . The site's surface is a heaved sand
soil extending uniformly 11 km in the North-South direction and 4 km in
the East-West direction. T h e three velocity components were measured
at z = 2.5 m above the surface using a triaxial ultrasonic anemometer
(Gill Instruments/1012R2) to an accuracy of ± 1 % . Sonic anemometers
Atmospheric Surface Layer Turbulence 87
Table I
Summary of energy, meteorological, turbulence, and surface roughness
conditions during the experiment. T h e net radiation was measured by
a Q6 Fritshen type net radiometer, the soil heat flux was measured by
two R E B S soil heat flux plates, the sensible heat flux was measured by
a Campbell Scientific eddy correlation system (sampling at 10 Hz), and
t h e friction velocity was measured by a triaxial sonic anemometer. T h e
m o m e n t u m roughness length was determined from 6 near-neutral runs
of (U) and it* using the logarithmic velocity profile.
Energy Conditions
Net radiation (Rn) 173 W m " 2
Soil Heat Flux (G) 78 W m " 2
Sensible Heat Flux (H) 90 W m " 2
Meteorological Conditions
Mean Horizontal Wind Speed ((U)) 2.68 m s " 1
Mean Air Temperature (Ta) 31.6°C
Turbulence Conditions
Friction Velocity (u+) 0.165 m s - 1
Root-Mean Square Velocity (au) 0.516 m s " 1
A t m o s p h e r i c Stability Conditions
Height above ground surface (z) 2.5 m
Obukhov Length (L) -3.98 m
Surface R o u g h n e s s
Momentum Roughness Length (z0) 0.13 mm
§4. W a v e l e t S t a t i s t i c s
In this section, we first develop relations between the Haar wavelet co-
efficients and the Fourier power spectrum and the structure function, then
we discuss hdw wavelet transforms can be used to investigate intermit tency
effects on the inertial subrange.
Atmospheric Surface Layer Turbulence 89
4 . 1 . R e l a t i o n b e t w e e n w a v e l e t coefficients a n d Fourier p o w e r
spectrum
In Fourier analysis, the fundamental tool used to characterize turbu-
lence is the power spectral density function E(K). T h e function E(K)
represents t h e energy density contained in each wavenumber b a n d dK, and
thus provides information regarding the importance of each scale of mo-
tion. However, important spatial information regarding location of events
become implicit in the phase angle of Fourier transform. In this section,
we relate t h e Haar wavelet coefficients t o t h e Fourier power spectrum and
show how spatial information can be expressed in an explicit manner.
T h e variance of the turbulence measurement, in terms of t h e wavelet
coefficients, can be deduced from the conservation of energy
M m
M 2 - -l
2 λ
a = Ν~ Σ Σ {WT^\i]f. (12)
m—l i=0
Km = £-. (14)
Hence, the power spectral density function E(Km) is computed by dividing
TE by t h e change in wavenumber AKm(= 2π2~ m dy~l ln[2]) so t h a t
E{Km) (15)
- 2^(2)
where (·) is averaging in space over all values of (i) for scale index (m)
(see [37] and [53]). The adequacy of Equations (14) and (15) are dis-
cussed in Katul ([24], pp. 149, 186-187). In Equation (15), the wavelet
power spectrum at wavenumber Km is directly proportional to t h e aver-
age of t h e square of the wavelet coefficients at t h a t scale. Because the
power at wavenumber Km is determined from averaging many squared
wavelet coefficients, we expect t h e wavelet power spectrum to b e smoother
t h a n its Fourier counterpart. This is apparent in Figure 2 which displays
good agreement between Fourier and wavelet power spectrum for all three
decades of inert ial subrange wavenumber s.
90 G. Katul et al.
Fourier
Wavelet
-5/3
SD K
^ ^ = d ^ 2 ) [ { W T ( m ) [ i ] 4 > - ((WT{m)w2))2}*- ( 16 )
In Meneveau ([37] and [38]), it was pointed out t h a t a plot of E(Km) and
E^m) + SDE gives a compact representation of the energy and its spatial
Atmospheric Surface Layer Turbulence 91
o
c
•22
mO
σ>
\_
4)
C
c
0)
0)
cSo
2 3 2 3 2 3 2 3
10-2 1 0- 1 1 00 101 102
Wovenumber (m- 1 )
<*««.> _i§gg=i. (1 „
An example of the variation of CVE is shown in Figure 3.
Notice in Figure 3 t h a t CVE increases as the wavenumber increases
indicating increased turbulent activity at smaller scales. T h e increase in
CVE can be related to the increased spottiness or intermittency in t h e
dissipation r a t e and is discussed next. In Tennekes and Lumley ([46] pp.
66), the dissipation rate for locally isotropic turbulence is given by
du
(e) = 15i/ (18)
dx
3. Using step (2), the Ax in the ratio (cr e /(e)) cancels out from t h e
numerator and denominator and we are left with the differencing
operations.
4 . 2 . R e l a t i o n b e t w e e n w a v e l e t coefficients a n d s t r u c t u r e f u n c t i o n
Using Equation (9) the wavelet coefficients can be related to the n t h
order structure function, for any flow variable 0, using
4 . 3 . C o n d i t i o n a l s a m p l i n g a n d i n t e r m i t t e n c y effects o n K 4 1
In general, intermittency of turbulent fluids is symbolized by an on-off
process in the dissipation r a t e so t h a t at a certain time, the turbulent energy
is only active in a small fraction of the fluid volume [36]. In a one dimen-
sional cut (namely along the x direction), the signature of intermittency is
large isolated dissipation events within an overall passive surrounding fluid
([36], pp. 104-109). For t h a t purpose, we classify the wavelet coefficients
as either "dissipative" or "passive". The dissipative wavelet coefficients are
the coefficients directly influenced by the large localized dissipation events
discussed above, while the passive coefficients are not. T h e distinction be-
tween dissipative and passive must be based on some minimum dissipation
threshold criterion. Such a criterion is difficult to construct without some
relation between the velocity and the dissipation. Before we establish t h e
conditioning criteria, let us first define the indicator function 1^ at scale
index (m) by
Arguments (1) and (2) may not be valid for turbulence t h a t is locally
anisotropic. However, based on the numerous studies of local isotropy in
ASL flows (e.g. [22]; [42], Ch.8), we adopt the working hypothesis t h a t tur-
bulence is locally isotropic in the inert ial subrange. These three arguments
94 G. Katul et al.
(22)
where e^ is t h e dissipation at position index (i) and scale ( m ) , (e)^m^ is
the mean dissipation at scale (ra), and [u^ (x + Ax) — u^m\x)]/Ax is t h e
finite difference approximation of t h e velocity gradient at scale (ra) and
position x. T h e first equality follows from arguments (1) and (2). T h e
second equality is due to t h e fact t h a t Ax required to convert gradients t o
differences (in a finite difference approximation of t h e derivative) cancels
out in t h e numerator and denominator, respectively. T h e third equality
is a direct consequence of Equation (20). For example, if F = 5, then all
squared wavelet coefficients t h a t are in excess of 5 times t h e average squared
wavelet coefficient at scale index (ra) are set to zero. This conditioning
criterion allows us to consider a conditional power spectrum Ec given by
E (Am) (23)
2^(2)
where (·) is now averaging in space over all non-zero values of [1^ WT^ (i)]2.
c
Hence, E represents t h e power spectrum of the passive fluid fraction. Also,
we can define t h e conditional n t h order structure function by
where ((·)) is averaging over all non-zero values of [1^ WT^ (i)]. These
conditional statistics can be computed by:
1) Using t h e pyramidal algorithm to calculate t h e Haar wavelet coef-
ficients at each scale index (m) and position index (i); 2) squaring these
coefficients to obtain t h e energy content at each scale index (m) and posi-
tion index (z); 3) averaging t h e squared wavelet coefficients for each scale
index (m); 4) dividing t h e squared wavelet coefficient (at space index i) by
the value computed in step (3); 5) If this ratio is larger t h a n some preset
value for F , then set this coefficient to zero, otherwise leave as is; 6) Use
Equation (23) or (24) t o determine t h e power spectrum or t h e n t h order
structure function with averaging performed over all non-zero values at
scale index (ra). Repeat t h e above steps for all values of (m) within t h e
inertial subrange. T h e adequacy of this conditional sampling criteria for
recovering K41 and characterizing intermittency is considered next.
Atmospheric Surface Layer Turbulence 95
10-1
F
h
^ 0 *
^10-2
>^o *
jy^o*
l· ^yr y
l· ο ^ *'
>" '
- 9 ^ >'
3
Q s ^ * ' '
2 - ^s^ ** • Conditioned
o Unconditioned
10-3 s Slope = 0.68
: ^^o
"· Slope = 0.66
~S^ s
'"+ ! 1I , | | ! 1 1 1 11
2 3 4 2 3 4 2 3 4
10-1 1 00 10 1 102
Seporotion Distonce r ( m )
§5. R e s u l t s a n d D i s c u s s i o n
This section discusses the effects of intermittency on K41 using t h e
conditional wavelet analysis for three cases: 1) n = 2, 2) ra = 3, and 3)
n = 6. In each case, we check whether K41 is recovered when intermittency
is suppressed, and then we investigate the statistical structure of t h e events
responsible for deviations from K41 scaling. We do not present theoretical
details regarding intermittency models, b u t we focus more on the contrast
between the conditioned (intermittency suppressed) and unconditioned (in-
termittency active) statistics.
Case 1: n = 2
It is known t h a t intermittency effects are generally small and may
not be detectable for t h e structure function with n = 2 ([1] and [52]).
We test this hypothesis by comparing the unconditioned and conditioned
[F = 5) structure functions of Equation (24). T h e results are presented in
Figure 4aa.
Both conditioned and unconditioned second order structure functions
96 G. Katul et al.
exhibit scaling laws t h a t are in agreement with K41 (slope= 2/3) indicating
t h a t intermittency effects may not be significant for n = 2 (see [7] and [8]
for a possible physical explanation). We also present a summary of t h e
regression statistics for the regression model log[.D 2 (r)] = Alog[r] + B in
Table 2. Notice in Table 2 t h a t the coefficient of determination (R2) for t h e
regression model is in excess of 0.99; hence, the determination of scaling
laws from wavelet structure functions appears to be very reliable.
Table II
Summary of the regression statistics for the model
\og[(\AU\n}] = Alog[r] + B. The coefficient of
determination (R2) and the standard error of estimate
(SEE) are also shown. The conditioned statistics are
for the conditioning criterion F = 5.
n Slope Intercept #2 SEE Conditioned (C)/
(A) (B) Unconditioned (U)
2 0.680 -2.73 0.996 0.038 C
0.660 -2.55 0.996 0.036 u
3 1.010 -3.88 0.995 0.056 c
0.950 -3.53 0.995 0.055 u
6 2.008 -7.10 0.993 0.150 c
1.690 -5.76 0.987 0.170
u
One must note t h a t certain scale aliasing occurs due to the use of or-
thogonal and complete basis since the number of Haar modes characterizing
the frequency domain is relatively small (m = 15). This aliasing may in-
fluence the indicator function / ( m ) . Hence, intermittency characterization
by the indicator function can be overestimated or underestimated based
on the value of F. For t h a t purpose, we performed the same analysis for
F = 4 , 5 , 7 , and 10. The slope variation (for n = 2) did not differ by
more t h a n 0.004. This analysis clearly reveals the robustness of the pro-
posed conditional sampling scheme. Similar results were also obtained by
Yamada and Ohkitani [54].
Case 2: n = 3
As shown by Landau and Lifshitz ([29], pp. 123-128), a relation be-
tween t h e third order structure function and (r) is given by
|(AC/) 3 | = Î ( € ) r . (25)
iu-' E
:
2
10-2
^y^ x»
^y^ *
-D 2 4
y ^ \ '
S 10-3
2 *" • Conditioned
•
10-4 y^^S' o Unconditioned
Slope = 1.01
Slope = 0.95
I
^ A -X . _ ! , . . J. ι ι ι ι I
| , , —1 1—1 1 1 1 1 1 1 1 1 1 1 1 1
2 3 4 2 3 4 2 3 4
10-1 1 00 101 102
Seporotion Distance r (m)
Case 3: n = 6
T h e sixth order structure function can be related to the dissipation
correlation function from
((At/)6)
(e(x)e(x + r)) (26)
where t h e dissipation correlation function is given by
Table III
Some values of the intermit tency parameter μ from various sources.
Source Measurements μ
Mahrt [32] Atmospheric mixed layer: 0.3-0.4
Convective and Nocturnal
Anselmet et ai [1] Turbulent and Duct flow 0.2 ± 0.05
Kuznetsov et al. [27] Wind-tunnel boundary layer 0.15-0.25
(Depends on external intermittency).
Monin and Yaglom [42] See Ch.8 for an extensive review 0.2-0.5
10-1
10-2
10-3
10-4
**·
10-6
10-7 • Conditioned
o Unconditioned
- - Slope = 2.008
10-8
Slope = 1.69
10-9
2 3 4 3 4 3 4
io- 100 " ~ ' 10* 102
Seporotion Distance r (m)
{(WT^\i])4)
FF(Rm) = (29)
((WTW\i})2)2
• Conditioned
o Unconditioned
§6. S u m m a r y a n d C o n c l u s i o n s
Triaxial sonic anemometer velocity measurements at 2.5 m above a uni-
form dry lakebed (11 km by 4 km) were used to investigate intermittency in
t h e inertial subrange. T h e power spectrum of the horizontal velocity mea-
surements exhibited a —5/3 power law for three decades allowing detailed
investigation of scaling laws in the inertial subrange. In order to describe
space-scale relations in the inertial subrange, we utilized the orthonormal
wavelet representation. T h e orthonormal wavelet representation was well
suited for this investigation since the basis function are orthogonal and
m u t u a l independence of the expansion coefficients is guaranteed. In addi-
tion, it was shown t h a t the expansion coefficients can be related directly
to quantities commonly used in conventional turbulence analysis. Rela-
tions between t h e orthonormal wavelet coefficients and the Fourier power
spectrum, as well as relations with the nth order structure function were
established. A comparison between Fourier and wavelet power spectra was
also carried out. Good agreement between the two spectra was noted even-
though the Haar wavelet has poor localization in the frequency domain.
Since intermittency build up in the inertial subrange was due to local-
ized dissipative events, a conditional wavelet scheme was developed. T h e
conditional wavelet scheme relied on an indicator function t h a t identified
the wavelet coefficients directly influenced by large and localized dissipation
events. T h e conditional wavelet scheme efficiently suppressed intermittency
within the inertial subrange by removing these wavelet coefficients. K41
statistics, up to sixth order, were recovered when intermittency in t h e dis-
sipation was suppressed from the wavelet coefficients. It was also found
t h a t intermittency did not significantly influence second and third order
statistics in agreement with many other studies. T h e robustness of t h e
conditional wavelet scheme was also verified.
Finally, we demonstrated t h a t intermittency was directly responsible
for non-Gaussian statistics in the inertial subrange, while K41 was associ-
ated with near Gaussian statistics. T h e wavelet transform produced inter-
mittency factors comparable to values obtained from laboratory and other
field experiments.
References
1. Anseimet, F., Y. Gagne, E. J. Hopfinger, and R. A. Antonia, High-
order velocity structure functions in turbulent shear flows, J. Fluid
Mech., 140, 63-89, 1984.
2. Aurell, E., U. Frisch, J. Lutsko, and M. Vergassola, On t h e multifractal
properties of the energy dissipation derived from turbulence data, J.
Fluid Mech., 238, 467-486, 1992.
3. Argoul, F., A. Arneodo, G. Grasseau, Y. Gagne, E. J. Hopfinger, U.
102 G. Katul et al.
1990.
54. Yamada, M., and K. Ohkitani, Orthonormal wavelet analysis of turbu-
lence, Fluid Dynamics Research, 8, 101-115, 1991a.
55. Yamada, M., and K. Ohkitani, An identification of energy cascade in
turbulence by orthonormal wavelet analysis, Prog. Theor. Phys., 86,
799-815, 1991b.
The authors would like to thank Scott Tyler for his assistance and support
at Owen's lake, and Teresa Ortenburger and Mike Mata for their help in the
data collection. We are grateful for the funding support from the National
Science Foundation (NSF) grant (EAR-93-04331), United States Geological
Survey (USGS), Water Resources Center (WRC) grant (W-812), Kearney
Foundation, and UCDAVIS superfund grant (5 P42ES04699-07).
Gabriel G. Katul
School of the Environment
Duke University
Durham, NC 27708-0328
USA
e-mail: [email protected]
John D. Albertson
Hydrologie Science
University of California at Davis
Davis, CA 95616
USA
Chia R. Chu
Department of Civil Engineering
National Central University
Chungli, Taiwan
Marc B. Parlange
Hydrologie Science &
Department of Agricultural and Biological Engineering
University of California at Davis
Davis, CA 95616
USA
An Adaptive Decomposition:
Application to Turbulence
§1. I n t r o d u c t i o n
A common issue in time series analysis is sorting out the different modes
of variation. If such modes overlap in Fourier space or are primarily local
or event-like, then traditional Fourier and eigenvector decompositions are
generally less effective in separating the different modes. As an alternative,
nine hours of turbulence d a t a are orthogonally decomposed into piece-wise
constants. Based on the decomposition statistics, four modes of variation
are defined. One of these modes includes a majority of the m o m e n t u m
flux and is described by a subrange of scales which depends on the record
position.
T h e Haar wavelet is the underlying basis in the decomposition applied
in this study. Decomposing the turbulence in terms of higher order wavelets
which are more compact in Fourier space yield similar results. Higher order
wavelets however, are slightly less efficient in capturing the sharp gradients
associated with the transport physics which is the primary goal of this
study. On the other hand, higher order wavelets do appear b e t t e r suited
for representing the larger scale, smoother variations. Additional reasons
for specifically using the Haar wavelet are provided at the beginning of
Section 2.1, and results of applying alternative orthogonal wavelets are
discussed at the end of Section 2.3.
In general, wavelets decompose global variance or energy in terms of
scale and position within the record. Wavelets are used extensively in
seismic analysis and are beginning to find their way into other geophys-
ical disciplines [3, 11, 12, 13, 18, 21]. Related to this study, Farge [9]
and Meneveau [23] discuss a variety of wavelet applications to turbulence.
Ways of constructing, describing, and implementing wavelet tools are many.
Daubechies [8] constructs wavelets which are compact in physical space,
and yet still provide a complete basis for decomposing the total sampled
variance. Viewing wavelet bases from a linear algebra perspective [24] can
be useful in solving numerical equations [1]. Spline wavelets are gener-
ally effective for interpolating and filtering d a t a [2, 7, 25]. Another com-
mon view is t h a t a wavelet basis set describes a multiresolution analysis
[10, 19, 22].
A multiresolution analysis is a convenient setting for locally describing
the different scales of variation in the data. Multiresolution techniques are
used in image analysis, for example, to store and transmit images compactly
[5,6]. T h e current development could be posed in terms of a multiresolution
analysis [16] or alternatively in terms of the wavelet analysis referred to
above. However, the methods in this study (Section 2) are kept simplified
such t h a t it is unnecessary to explicitly appeal to these closely related
topics.
T h e strategy in this study is to first decompose the turbulence time
An Adaptive Decomposition: Application to Turbulence 109
series into distinct modes. Because small scale variations in the turbulence
record correspond to two types of motions, an adaptive technique is used
to separate t h e m (Section 2.2). For a simple demonstration of t h e adaptive
technique, only t h e scale separating these two small scale modes is allowed
to spatially vary. Generally, the adaptive technique allows for t h e separa-
tion of physically distinct modes with overlapping scales. This separation
is not possible in conventional filtering with a specified response function or
distribution of weights. After decomposing the turbulence, the spatial dis-
tribution of m o m e n t u m flux will be reconstructed for the different modes.
Additional statistics are then computed for the different modes (Section
2.3). In Section 3 a brief physical interpretation of the results is provided.
§2. P a r t i t i o n i n g t h e T i m e Series
T h e d a t a used in this study consist of 9.1 hours of the three velocity
components measured 45 meters above flat terrain in near neutral condi-
tions [17]. T h e wind speed fluctuates about a mean value of ü = 12.8 m / s
throughout the 9.1 hours, and most of the energy is concentrated in t h e
1 minute or 1 km eddy motions (Section 2.3). For a turbulence d e p t h of
roughly 500 m the Reynolds number is more t h a n 10 8 . Additional statistics
describing this d a t a set can be found in [17] and [21].
T h e value of the longitudinal wind component at the i t h record position
is denoted as
«,·; i = l,2,...,2M (1)
where δ = ^ s is the width of a sampling interval and in t h e current analysis
M = 19, corresponding to 524,288 d a t a points. There are equivalent time
series of t h e cross stream component v and the vertical wind component
w.
i i
π=
^Σ>- (2)
Variations in the longitudinal wind are described by difference terms de-
fined as
i 2 m -i
i — 1 i— 1
n = 1 + int(——) ; £ = 1 + int(- r)
The total covariance between the time series of the longitudinal and vertical
wind components is obtained by multiplying the u reconstruction (4) by an
equivalent expression for w, and averaging over all record positions. The
resulting covariance quantity is written as
2
l
-JM Ys(Ui * U)(Wi ~W) = (6)
i=l
M Λ 2 M-m
Σ ^Μ^ Σ ^u(am;n)Aw(am;n).
2
7Ti=l n=l
. /l'■ — 1 x „ . . j' - 1N
n = 1 + int(——)
v ; t = 1 + int(- r)
2m 2m
where raf = M, ra^ = 1 , and ra+ = m~_1 — 1. A given mode c at the
i t h record position includes all the difference terms (3) associated with the
range of dilation scales a - through a + · The small scale cutoff for the
112 J. Howell and L. Mahrt
2.2.1. Definitions
T h e particular modal definitions, detailed below, are based on an inter-
pretation of t h e variance and covariance spectra (Section 2.3). T h e modal
reconstructions (Section 2.2.2) substantiate the partitioning. T h e largest
scale (mesoscale) mode leads to very little flux and consist of predominantly
horizontal motions. T h e large eddy mode leads to some flux, and for this
114 J. Howell and L. Mahrt
1
I ' ' ' ' I ' ' ' ' I ' ' ' ' I ' ' ' ' I ' ' ' ' I ' ' ' ' I ' ' ' '
U=19
m o d e of v a r i a t i o n
100km
15 = 1 mesoscale
l
c = 2 large eddy
n° 10 1 km
c = 3 transporting eddy
10 m
c = A fine scale
m= 1 I . . . . I . . . . I . . . . I . . . . I
Figure 1. The small scale cutoffs for the four different physical modes
of variation
where n and £ are defined in (10). This mode includes scales between
αιο « 800 m and a\2 « 3.3 km.
(iii) T h e transport mode, c = 3, at the ith record position is defined as
9
u
^i= Σ (-l)'Ati(a m ;n) (12)
m=mä i
where 1 < ra«^· < 7 is determined for each interval of length αγ « 100
m according to t h e above discussion. This mode includes scales between
a\ « 1.6 m and a 9 « 400 m.
(iv) T h e fine scale mode, c = 4, at the z th record position is defined as
m
i,i
u
*,i = ] £ (~l/Au(a m ; n) (13)
2.2.2. R e c o n s t r u c t i o n s
T h e sum of the different modes is equal to (4), so it follows t h a t
c
Ui -ü =^2uCii (14)
c=l
§3. V a r i a n c e a n d Covariance S p e c t r a
A global estimate of the variance on the scale a m is the square of
the Haar transform (3) on the scale am averaged over all the translation
positions, written as
ηΙΥΊ — TO
1
var[ (15)
O Q I I I I I | I I I I | I I I I | I I I I | I I I I | I I I I | I I I I | I I I I
5 I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I
Figure 3. The sum of the four different modes which equals the orig-
inal data.
scales, and the difference (transform) intervals are allowed to overlap. This
leads to a difference t e r m defined as
(a)
4
2
s
, , | „ M | , , U | , U , | , , : H | , , , , | , , I , | H , , | , _
w I i i i iI i i I i Ii I i iI i i i i I i i i i ii i i i i i i i i_
fcl
0 -u. /•iJ" ^'· 1
0 M||v|W"' *|H^Yl||^lj
-2
50km X
-4 IIIIMIMNII ■ I M I I I I M M I I M I I M MI
*- F M \ 2
£ °
-4 500m
X
s -2 50^m
/I ■ I ■ ■ ■ ■ I ■ ■ ■ ■ I ■ ■ t ■ I ■
3 -4
198 200 202 199.8 200 200.2
5-7). T h e numbers for each mode in the lower portion of the figures are
the relative contributions to the total covariances. These numerical values,
in addition to other statistics, are summarized in Table I.
Included in these statistics are gradients in the longitudinal wind com-
ponents. Gradients in a piece-wise constant signal are readily defined as
the difference between adjacent constant values divided by the distance
between the center positions of the associated segments. Consequently, for
t h e t r a n s p o r t mode the gradients are computed over variable distances,
since t h e piece-wise constants are of variable width (Figure 2, c = 3).
T h e skewness values of the gradients in the longitudinal wind compo-
nents for t h e different modes are listed in Table I. A positive (negative)
value of the skewness for the u gradients indicates the wind speed increases
(decreases) more rapidly in the downstream direction. Thus, t h e results
120 J. Howell and L. Mahrt
J=, 0.15
CD
Ü
I °· 10
CO
>
£ 0.05
0.03
0.00
1000 100 10 1 0.1 0.01 0.001
dilation a (km)
Table I
Haar wavelet decomposition statistics. Variances (var and cov) are in
m2/s2. T h e values in the parenthesis are the results if the adaptive step is
not taken to include additional small scale motions in the transport mode.
Mode Mesoscale Large Eddy Transport Fine Scale Total J
u var 1.043 1.474 1.456 0.411 4.3841
(1.263) (0.604)
v var 0.779 0.407 1.019 0.524 2.7294
(0.852) (0.691)
w var 0.027 0.170 0.731 0.457 1.3845
(0.526) (0.662)
u-w cov -0.049 -0.300 -0.438 -0.068 -0.8552
(-0.368) (-0.138)
u-w corr -0.29 -0.60 -0.42 -0.16 -0.35
(-0.45) (-0.22)
isotropy 0.03 0.18 0.59 0.98 0.39
(0.50) (1.02)
skewness of 0.048 -0.227 -0.846 -0.006 -0.220
u gradient (-0.230) (-0.019)
An Adaptive Decomposition: Application to Turbulence 121
Table II
As in Table I except the numbers are D4 wavelet decomposition statistics.
Mode Mesoscale Large Eddy Transport Fine Scale Total
u var 1.052 1.556 1.364 0.372 4.3441
(1.207) (0.529)
v var 0.783 0.432 1.023 0.490 2.7282
(0.891) (0.622)
w var 0.029 0.176 0.737 0.442 1.3838
(0.563) (0.616)
u-w cov -0.069 -0.311 -0.421 -0.059 -0.8605
(-0.363) (-0.117)
u-w corr -0.40 -0.59 -0.42 -0.15 -0.35
(-0.44) (-0.20)
isotropy 0.03 0.18 0.62 1.02 0.39
(0.54) (1.07)
§4. P h y s i c a l I n t e r p r e t a t i o n
T h e turbulence d a t a analyzed in this study has been decomposed into
deviations of small scale averages from larger scale averages. T h e turbu-
lence measurements include 9.1 hours of wind tower d a t a which translates
to a b o u t 420 km using Taylor's hypothesis and a mean wind speed u = 12.8
m / s . T h e flow is partitioned primarily according to t h e scale dependence
of the u-w covariance (Figure 7). Deviations of 3.2 km averages from the
entire record average (variations on scales > 6.4 km) define the mesoscale
mode. Deviations from the 3.2 km averages represent t h e turbulence, which
in t u r n are partitioned into three different modes of variation (Section 2).
Based on the u-w covariance (see shoulder in Figure 7 around a = 2
km), t h e large eddy mode is defined by deviations of 400 m averages from
the 3.2 km averages, t h a t is variations on the scales between 800 m and
3.2 km. T h e exact scales defining t h e large eddy m o d e are somewhat
arbitrary, since the physics changes between the small scale and large scale
ends of this regime. At the small scale end of the large eddies, t h e vertical
motions become more significant. T h e large eddy mode corresponds to
motions which are more horizontal at the 45 m observation level compared
to the smaller scale motions as indicated by the small value of t h e isotropy
coefficient (Table I).
T h e large eddies may be the lower part of boundary layer scale motions
(on t h e order of 1 km deep) where t h e lower part of the eddies observed at
the tower level are forced by the ground to be more horizontal. In this case,
t h e t r a n s p o r t by the large eddies would increase with height. Roll vortices
are an example of such eddies [4]. W h e n observed from time series measured
from towers, such larger scale motions are sometimes referred to as inactive
eddies because they contribute significantly to the horizontal variance b u t
contribute little to t h e m o m e n t u m flux at levels closer to t h e ground. In this
case t h e motions are not considered to be traditional turbulence and must
be removed from t h e signal before traditional similarity arguments can b e
applied [15]. This concept may be most descriptive of t h e larger scale p a r t
124 J. Ho well and L. Mahrt
of the large eddies in the present partitioning. The weak vertical motions
t h a t occur on the large eddy scales, however, are well correlated with t h e
variations of the longitudinal wind component leading to significant (35%
of total) m o m e n t u m flux. While the mean shear seems to exert a greater
influence on the large eddy mode compared to the mesoscale mode (greater
gradient skewness, Table I), the shear effect on the large eddies is still small
compared to t h a t of the transporting eddies.
Deviations of the raw time series from the 400 m averages define t h e
two smallest scale turbulence modes. In order to distinguish the two types
of small scale motions, the scale separating the two modes varies depending
on the local behavior of the transport as discussed in Section 2. Specifically,
averages are computed over a sufficiently small scale ( < 400 m) in order
to resolve the local vertical transport of momentum. Deviations of these
smaller scale averages from the 400 m averages are included in the t r a n s p o r t
mode to capture a majority of the momentum flux, including local extremes
(Figure 4, panel c).
As a result, the fine scale structure is also determined by the variable
cutoff scale. This is in contrast to conventional high pass filtering where
the cutoff scale is constant throughout the record. A constant cutoff scale
is not used in this case because transporting eddies intermittently occur
on scales which are traditionally assigned to the fine scale structure. From
another point of view, the fine scale mode includes relatively larger scale
motions only at locations where small scale transport is absent. This means
t h a t motions occurring on a range of scales between 1.6 m and 50 m are of
the transporting eddy type at some positions while at other positions these
motions make u p the fine scale structure.
T h e t r a n s p o r t mode is characterized by strong gradient skewness (Ta-
ble I) reflecting t h e strong influence of the mean shear on the transporting
eddies. According to conventional expectations, the mean shear generates
eddy motions which transport higher m o m e n t u m toward the surface. In
terms of energetics, this momentum transport corresponds to conversion of
mean kinetic energy to turbulence kinetic energy. T h e d a t a d e c o m p o s i -
t i o n in t h i s s t u d y p r o v i d e s a definition of t h e t r a n s p o r t i n g e d d i e s
w h i c h allows q u a n t i t a t i v e verification of classical c o n c e p t s . Also
as expected, the value of the isotropy coefficient for the transporting eddies
is between t h a t of the large eddies and the fine scale structure. T h e main
transporting eddies are characterized by significant vertical motions which,
however, remain smaller t h a n the horizontal velocity components. This
is because the mean shear directly generates variance in the u-component
which subsequently induces vertical velocity fluctuations through pressure
fluctuations.
T h e remaining small scale deviations make up the fine scale structure.
Very little t r a n s p o r t is associated with the fine scale mode, and t h e energy
An Adaptive Decomposition: Application to Turbulence 125
§5. C o n c l u s i o n
Geophysical time series generally consist of physically distinct modes of
variation, each occurring on a subrange of scales which depend on space and
time. Conventional decomposition or filtering techniques divide t h e time se-
ries according to scales which are constant in space and time. In this study,
distinct modes were isolated using a piece-wise constant (Haar wavelet) de-
composition which allows the scales defining a particular mode to vary with
record position. W i t h this approach sampled covariances are completely
and orthogonally decomposed. Partitioning the flow in this manner allows
assessment of t h e relative contributions of the different modes to traditional
statistics.
T h e turbulence d a t a analyzed in this study has been partitioned into
four modes of variation. Each mode is defined locally in terms of an upper
and lower cutoff scale. T h e cutoff scales for the two larger scale modes are
specified to be constant with respect to record position. T h e scale sepa-
rating t h e two smaller scale modes varies with position according to t h e
local m a x i m a in the spatial distribution of m o m e n t u m flux. A local mo-
m e n t u m flux is quantified in terms of the product of t h e difference t e r m s
Au(am;n) x A w ( a m ; n ) , which is equivalent to a product of wavelet co-
efficients at a fixed scale and position. Using a wavelet decomposition to
examine the spatial or temporal distribution of the scale dependent flux is a
promising approach for distinguishing distinct physical modes of variation.
Adapting the decomposition to the spatial distribution of m o m e n t u m
flux leads to an improvement in the small scale partitioning as interpreted
in terms of globally averaged statistics. T h e spatial dependence of t h e
cutoff scale allows the computed transport mode to capture more of the
m o m e n t u m flux; some of the flux occurring on scales traditionally included
as fine scale structure is now more correctly included in t h e t r a n s p o r t mode.
For t h e t r a n s p o r t mode, t h e gradients in the longitudinal wind component
are negatively skewed in the downstream direction, which verifies t h a t this
m o d e is primarily shear driven. If a constant cutoff scale were used, t h e
skewness of t h e gradients in t h e longitudinal wind for t h e main transporting
motions would b e only —0.230. After the additional small scale variations
are included by varying t h e cutoff scale, the gradient skewness is —0.846.
This change is a result of further resolving the microfronts associated with
126 J. Howell and L. Mahrt
References
1. Alpert, B. K., Wavelets and other bases for fast numerical linear al-
gebra, Wavelets: A Tutorial in Theory and Applications (C. K. Chui,
ed.), Academic Press, 181-216, 1992.
2. Battle, G., Cardinal Spline Interpolation and the Block Spin Construc-
tion of Wavelets, Wavelets: A Tutorial in Theory and Applications (C.
K. Chui, ed.), Academic Press, 73-90, 1992.
3. Bradshaw, G. A., and T. A. Spies, Characterizing canopy gap structure
in forests using wavelet analysis, J. Ecology, 80, 105-215, 1992.
4. Brown, R. A., Longitudinal instabilities and secondary flows in the
planetary boundary layer: A review, Rev. Geoph. Space Phys., 1 8 ,
683-697, 1980.
5. Burt, P. J., The Pyramid as a Structure for Efficient Computation,
Multiresolution Image Processing and Analysis (A. Rosenfeld, ed.),
Springer Verlag, 6-35, 1984.
6. Burt, P. J., T h e Laplacian pyramid as a compact image code, IEEE
Trans. Commun., 3 1 , 532-540, 1983.
7. Chui, C. K., An Introduction to Wavelets, Academic Press, 1992.
8. Daubechies, L, Orthonormal bases of compactly supported wavelets,
Commun. Pure Appl. Math, 4 1 , 909-996, 1988.
9. Farge, M., Wavelet transforms and their applications to turbulence,
Annu. Rev. Fluid Mech. 24, 395-457, 1992.
An Adaptive Decomposition: Application to Turbulence 127
§1. Introduction
Daytime turbulence in the vicinity of vegetation canopies has been ex-
tensively investigated during the past ten years. It is now widely recognized
that, within and just above vegetation canopies, turbulent transport pro-
cesses are to a large extent dominated by intermittent, energetic coherent
structures, with length scales of the order of the canopy depth ([1], [18],
and [20]). A dynamic model for the formation and development of these
'canopy eddies' has been proposed recently ([11] and [19]).
§2. E x p e r i m e n t a l P r o c e d u r e s a n d F l o w C h a r a c t e r i s t i c s
T h e experiment was carried out over a maize crop at Grignon, France
(48°51'N, 1°58Έ), from July to September 1990. T h e d a t a presented here
were obtained on August 13. At this date, the canopy height (evaluated
as the m e a n height of a set of 100 individual plants) was h = 1.55 m, with
a zero-plane displacement height d = 1.15 m (d is the apparent level of
m o m e n t u m absorption by the plants, used as the origin of heights in crop
studies). T h e leaf area index (LAI, or total leaf area per unit soil surface)
was 4.1. A meteorological mast was equipped with slow-response temper-
ature (shielded, aspirated thermocouples) and windspeed (cup anemome-
ters) sensors at 7 heights z ranging from z/h = 0.3 to z/h = 4.6. T h e
distance from the leading edge of the field was larger t h a n 200 m in the
direction of prevailing winds. Two three-dimensional sonic anemometers-
thermometers (model PAC-100, Dobbie Instruments), were also set u p on
the mast at heights z/h = 0.6 and 1.6. All three velocity components (u,
streamwise; t>, lateral; tu, vertical) and air t e m p e r a t u r e T were sampled at
20 Hz, a r a t e high enough for exploring a significant part of t h e inertial
subrange, and stored on a micro-computer.
Only t h e d a t a acquired above the canopy are used here. Linear trends
are first removed from all original time series, split ted into contiguous 30
min runs. In what follows, prime notations ΐί', w' and T' stand for t h e
fluctuations of ii, w and T around their mean values ïï, w and T, calculated
for each run.
Two sets of samples are used in this study, one ('diurnal') acquired from
13:30 to 15:30 U T and t h e other ('nocturnal') from 18:00 to 20:00 U T , just
before and after sunset (which occurred at 19:10 U T ) . An illustration of t h e
132 Y. Brunei and S. Collineau
Table I
Summary of experimental conditions: statistical moments of u, w and T
(means, standard deviations, covariances) and stability parameter
\z - d)/L.
1 Time u T (JT u'w' w'T' — * - * ii
(m/s) (°C) (m/s) (m/s) (K) (m 2 /s 2 ) (K m/s) L
T(eC)
a a (1)
We use only the continuous version of the wavelet transform of a function
/ , noted ( / , ψ ( α ^ ) and defined as:
+oo
/(i)V>(a'6) dt. (2)
/
-oo
Let / be a given experimental time series with energy Ef = ( / , / ) . We
Diurnal and Nocturnal Turbulence Above a Maize Crop 135
where
ΙΦΗΙ 2 ^ (6)
ω
-co
0.2H
. O.H
""Too
D(s)
loo
D(s)
D(s)
We have calculated the wavelet variances of i/, w and T, and the wavelet
covariances of (u,w) and (u;,T), for all the available d a t a samples. T h e
Mexican h a t wavelet was chosen for this, because of its good localization
in frequency. Figure 3a shows the variances and covariances obtained from
14:00 to 15:00 U T , under slightly unstable conditions.
Single, well-defined peaks can be seen on the wavelet variance and
covariance curves. T h e wavelet variance for vertical wind velocity WVW
peaks at Dw « 1-2 s, whereas WVU and WCT peak at Du « DT ~ 3 -
4 s. From a qualitative point of view, these results are very similar to
those obtained by Collineau and Brunet [5] over a pine forest in similar
stability conditions. Only the magnitudes of the Z)-scales differ: in t h e
latter case typical values of 5-7 s and 10-15 s respectively were obtained
for the w-variance peaks, and the peaks for u and T. However, normalizing
the duration scales Du or DT by w* and h as suggested in the introduction
give t h e same mean value Du*/h « 0.85 for b o t h canopies. It therefore
seems t h a t over the maize crop the motions contributing to these peaks are
also typical canopy-scale eddies, of the same type as those depicted in [5],
This will be confirmed further.
T h e covariances WCUW and WCWT also exhibit unambiguous peaks, at
the same D scales as WVU and WVT (Figure 3a). Instead of using WCUW
and WCWT, Collineau and Brunet [5] calculated the wavelet variance of
series of instantaneous cross-products u'w' and w'T'. They were shown to
display a peak at the same scale as WVW. This is very similar to what hap-
pens with Fourier spectra and cospectra in the surface layer: the cospectra
peak at the same frequency as the u and T spectra (see [13] for a review),
whereas t h e spectra of the time series of cross-products such as u'w' and
w'T' peak at t h e same frequency as the w spectra.
Figure 3b shows the wavelet variances and covariances obtained from
19:00 to 20:00 U T , under conditions of moderate stability. T h e wT co-
variance is now negative, following the inversion in sensible heat flux, b u t
all curves are qualitatively similar to those in Figure 3a, with identical D
scales. T h e variances of u and w appear flatter t h a n in Figure 3a b u t this is
just a consequence of a change in the respective magnitude of t h e wavelet
variances and covariances. Full-scale graphs of WVU and WVW indeed show
well-marked peaks at the same D-scales as in Figure 3a. This confirms t h e
visual impression given by Figure l b , t h a t the same kind of structures as
those commonly seen in diurnal time traces are also present in this noctur-
nal sample. T h e normalized peak scales Du*/h are now about 3-4 times
larger t h a n in Figure 3a since ii* has dropped from 0.3-0.4 m s - 1 to a b o u t
0.1 m s - 1 . Using mean wind speed as a velocity scale would lead to a
smaller difference since ΰ has only dropped by a factor of 2.
Considering now t h e period from 18:00 to 18:30 U T (Figure 3c), two
peaks are apparent on all curves b u t WVW. Consequently, there seems t o
138 Y. Brunei and S. Collineau
§4. J u m p D e t e c t i o n a n d C o n d i t i o n a l S a m p l i n g
Having determined a typical event duration, we are now interested in
how frequently these events occur. Several methods for detecting sharp
edges in digital time series, based on the use of localized transforms, have
been proposed recently ([12] and [14]). The combined localization prop-
erties of wavelets in b o t h time and frequency can be exploited together
usefully for designing efficient jump-detection algorithms. Collineau and
Brunet [5] showed t h a t using (Τ,φ^) (or ( w , ^ ( a , 6 ) ) ) as detection func-
tions of ramps (or large excursions from the mean) in time series of temper-
ature (or streamwise wind velocity), after determination of the scale a of
the wavelet variance peak, leads to an accurate detection of the structures
visible in t h e time series. T h e final decision step in the detection algorithm
requires an empirical threshold for first derivative like wavelets (e.g., the
Haar wavelet), but only a slope sign for the Mexican hat wavelet, which
yields zero-crossings in (T, φ(α^) (or (w, φ(α^)) whenever large, sharp rises
or drops occur in the series. W i t h a comparable reliability, the zero-crossing
algorithm is therefore simpler t h a n those based on thresholds, which is the
case of most standard jump-detection algorithms used in turbulence, such
as the Variable Interval Time-Averaging technique (see for example [21]).
However, it has been made clear in [15] t h a t detection through first or sec-
ond derivative like wavelets are equivalent algorithmic problems when the
objective is to classify the 'sharpness' of variations of the signal, which, in
the latter case, requires thresholding the slope of the wavelet transform.
Here, detection was performed on all time series of streamwise wind
velocity, using the Mexican hat wavelet. Wind velocity was chosen because
it has a higher signal-to-noise (as defined in the introduction) ratio t h a n
t e m p e r a t u r e in this particular d a t a set, and also because in this case the
slope sign criterion does not depend on whether the d a t a are diurnal or
nocturnal. Figure 4 shows histograms of the time intervals Δ between the
detected events, for two typical half-hour samples with similar values of
w*, in b o t h diurnal and nocturnal conditions. T h e two distributions are
strikingly similar and provide the same mean value Δ « 9 s. Normalizing
Δ as above, we obtain a mean value Au*/h « 2.1 over the whole d a t a set,
which is again in fairly good agreement with the value of 1.8 obtained over
the forest canopy already mentioned [5].
Finnigan ([9] and [10]) found t h a t the peak frequency of the Fourier
Diurnal and Nocturnal Turbulence Above a Maize Crop 139
U.H--
ri
\l
14:30 1
0.3-
'55
iI
G 1 18:00
1^0.2-
L^
Probabi
1
o
0 5 10 15 20
D(s)
«o1
0.0Û1
001
I I I I I II
0.01
1 1—I I I I I II
0.1
1 1—I I I I HI
1
1 1—I I I I II I
10
f(Hz)
0.15
J . l j i 1 1—i Ii Ii 11
T—I I I ii
11 1 1—I I I I I II 1 1—I I I I III 1 1—I I I I II
10
Wooi o.oi o.i i
f(Hz)
Figure 5. u-, w- and Γ- Fourier spectra (a), and um- and wT-
cospectra (b) for the sample 14:00-15:00 UT. Spectral densities are mul-
tiplied by the frequency and divided by the signal variance.
Diurnal and Nocturnal Turbulence Above a Maize Crop 141
case of the Mexican hat. Instead, we use the ' R a m p ' wavelet defined as:
i=l f*
with Ti — Δ / 2 < t < T{; + Δ / 2 , r,· being the ith detection time.
T h e averaged p a t t e r n s are presented in Figures 6a (diurnal samples)
and 6b (nocturnal samples). Both figures show a characteristic slow ({ti} <
0) upward ({w} > 0) movement of air, rapidly switching to a strong down-
ward ({w} < 0) motion associated with an acceleration of horizontal ve-
locity ({u} > 0). In the diurnal case a sharp t e m p e r a t u r e drop follows a
slow rise, whereas t h e opposite occurs in the nocturnal case. Such p a t t e r n s
have t h e characteristics of typical 'ejection-sweep' processes. It has t o b e
emphasized t h a t detection was performed on time series of t e m p e r a t u r e
only, which suggests a strong coupling between temperature, vertical and
streamwise velocities.
These results invite several comments. Firstly, they show t h a t above
our maize crop t h e motions responsible for t h e peaks in wavelet variances
and covariances, and objectively detected by t h e wavelet transform, do have
t h e 'signatures' of coherent eddies, such as those observed over a variety of
plant canopies or rough surfaces ([18], [19], and [20]). Secondly, t h e nor-
malized p a t t e r n s are very similar to those obtained by t h e same technique
above a forest canopy ([5]), with {T} peaking between ± 0 . 6 and ± 0 . 8 , and
{u} and {w} having equally smaller amplitudes, b o t h comprised between
142 Y. Brunei and S. Collineau
0.8-
<Λ
0.6
a
VH {T}
> 0.4- {u}
cd
i—H
G 0.2- "{w}
O
■<-j
Ό 0-
e
o
X) -0.2- "**---.--- S
<
Nυ
'K! -0.4-
εo
t-<
-0.6-
(a)
-0.8
~A~ÎS Ï 3 F
t(s)
about ±0.2 and ± 0 . 5 . Thirdly, apart from the difference in the tempera-
ture signature due to the opposite sign of the mean vertical t e m p e r a t u r e
gradient, turbulence seems to be dominated by the same processes in our
daytime and nighttime samples. However, this should not b e generalized
since a relatively small stability range has been explored ((z — d)/L between
- 0 . 0 6 4 and 0.404, as shown in Table 1).
We noticed in Figure 3c t h a t just after the inversion in t h e mean tem-
p e r a t u r e gradient, another peak was visible in the wavelet variances and
covariances, corresponding to longer duration scales. As gravity waves have
already been observed in similar conditions [17], one may reasonably won-
der whether those peaks could be due to such phenomena. This is t h e
objective of the next section.
§5. W a v e l e t D e c o m p o s i t i o n
T h e occurrence of linear gravity waves can be predicted theoretically
by a linear stability analysis of the dynamical flow equations, as t h e result
of Kelvin-Helmoltz instabilities (see [7] and [8]). This analysis provides a
characteristic wave frequency depending on the mean vertical t e m p e r a t u r e
gradient. This is the Brunt-Vaisala frequency N , defined as:
0.08
200
§6. S u m m a r y a n d C o n c l u s i o n s
T h e results of this experimental study confirm the importance of co-
herent structures in turbulent transfer processes between plant canopies
and the atmosphere. The features observed in time series of velocity com-
ponents and air temperature recorded above a maize crop are very similar
to those observed above other plant canopies, and quite consistent with
the picture of canopy turbulence acquired over the past few years. Wavelet
analysis has enabled us to detect characteristic signatures of coherent mo-
tions in all series, revealing the occurrence of ejection-sweep processes. T h e
scale of these motions is canopy-dependent and a comparison between the
pine forest and the maize crop suggests t h a t the occurrence frequency of
these motions scales with friction velocity and canopy height. In b o t h cases
a value of Au*/h « 2 was obtained for the mean time interval between
successive structures. Also, a unique value Du*/h « 0.85 was obtained for
the duration scale corresponding to the peak in the wavelet variances of
streamwise velocity and temperature. The wavelet covariances of uw and
wT were also observed to peak at the same scale. These results support
the postulate t h a t 'universal' mechanisms are responsible for the structure
of turbulence in the vicinity of plant canopies, in near-neutral conditions
([1], [11], and [19]).
Analysis of nighttime samples in light to moderate stability also proved
the existence of similar processes. In moderate stability ((z — d)/L œ 0.3)
the time scale Au*/h was found to be somewhat larger, but this needs
further confirmation since only one hour long sample was available for this
Diurnal and Nocturnal Turbulence Above a Maize Crop 147
0.08
0.04
study. In lighter stability, around the inversion time of the mean vertical
temperature gradient, another scale of motions was observed, for which no
satisfactory explanation can be given at the present time.
These results were obtained by performing a wavelet analysis of tur-
bulent time series, consisting in several steps. Firstly, computation of
wavelet variances and covariances provide characteristic duration scales
of the events contributing most to the signal energy. Secondly, the wavelet
transform is used for detecting these events, by 'looking' at the series at
these particular scales. As far as the detection itself is concerned, the use
of a wavelet such as the Mexican hat, which yields zero-crossings whenever
large discontinuities occur in the series, provides a simple detection scheme
which does not require adjustment of an empirical threshold. Conditional
averaging of the detected patterns can then be performed, enabling one to
extract from the time series a clean picture of the signatures of these mo-
tions. Compared to more traditional analysis techniques, this procedure
has the major advantage of being based upon a unique, self-consistent
line of mathematical treatments, which rely on the combined time- and
frequency-localization properties of the wavelet transform.
These properties have also been used for partitioning the original series
into large- and small-scale components, using a cut-off scale defined as the
gap scale between two peaks in the wavelet variances and covariances. This
has promising applications, for instance in the study of wave-turbulence
interactions.
References
1. Brunet, Y., J. J. Finnigan and M. R. Raupach, A wind tunnel study
of air flow in waving wheat: single-point velocity statistics, Boundary-
Layer MeteoroL, 1994, In press.
2. Caughey, S. J., Boundary-layer turbulence spectra in stable conditions,
Boundary-Layer MeteoroL, 11, 3-14, 1977.
3. Caughey, S. J. and C. J. Readings, An observation of waves and tur-
bulence in the earth's boundary layer, Boundary-Layer MeteoroL, 9,
279-296, 1975.
4. Collineau, S. and Y. Brunet, Detection of turbulent coherent motions
in a forest canopy. Part I: Wavelet analysis, Boundary-Layer MeteoroL,
65, 357-379, 1993.
5. Collineau, S. and Y. Brunet, Detection of turbulent coherent mo-
tions in a forest canopy. Part II: Time-scales and conditional averages,
Boundary-Layer MeteoroL, 66, 49-73, 1993.
6. Daubechies, I., The wavelet transform, time-frequency localization and
signal analysis, IEEE Trans. Information Theory, 36, 961-1005, 1990.
7. De Baas, A. F. and A. G. M. Driedonks, Internal gravity waves in a
Diurnal and Nocturnal Turbulence Above a Maize Crop 149
1984.
22. Stull, R. B., An introduction to boundary layer meteorology, Kluwer
Academic Publishers, Dordrecht, 1988.
Yves Brunet
Laboratoire de Bio climatologie
INRA
B P 81
33883 Villenave d'Ornon Cedex
France
email: [email protected]
Serge Collineau
Laboratoire de Bioclimatologie
INRA
BP81
33883 Villenave d'Ornon Cedex
France
Wavelet Spectrum Analysis
and Ocean Wind Waves
Paul C. Liu
§1. Introduction
Ever since Willard J. Pierson [18] adopted the works of John W. Tukey
[22] and introduced the power spectrum analysis to ocean wave studies,
Fourier spectrum analysis has been successfully and persistently used in
data analysis of wind-generated ocean waves. Over the past four decades,
with the increased availability of new instruments for measuring wind and
waves, spectrum analysis has continued to be the fundamental standard
procedure used for analyzing wind and wave data.
Fourier spectrum analysis generally provides frequency information
about the energy content of measured, and presumed stationary, time-
series data. Characteristic properties of waves such as total energy and
dominant or average frequency can be readily derived from the estimated
spectrum. This information, however, pertains only to the time span of
the measured data. Changes and variations within a time series cannot be
easily unraveled. As stationarity in the data simply represents a mathe-
matical idealization, its validity is usually regarded as an approximation of
the real wave field. The effectiveness of applying Fourier spectrum analysis
to a rapidly changing wave field, such as during wave growth or decay, is
§2. W a v e l e t S p e c t r u m
Following a standard formulation [3], we briefly summarize the wavelet
transform. We start with a family of functions, the so-called analyzing
wavelets, ißab(t)i that are generated by dilations a and translations b from
a mother wavelet, ip(t), as
φα,(ή = ^ = φ ( ^ ) (1)
where a > 0, —oo < b < +oo, and f_™ ψ(ί)<ϋ = 0. The continuous wavelet
transform of a time-series, X(t), is then defined as the inner product of ipab
and X as
V.r(t) = 2 - / V ( 2 - t - r ) , (4)
where s and r are integers. Then the continuous wavelet transforms (2)
and (3) for time series data X(t) become
1 /*+°° t
x{s T) = x{t)r{ (5)
' 7Ploo v "T)dt
Wavelet Spectrum Analysis 153
and
(6)
In general, t h e studies of wavelet transforms and wavelet analysis are
centered on two basic questions [4]: (1) Do t h e wavelet coefficients com-
pletely characterize t h e time-series d a t a ? (2) C a n t h e original time series
be reconstructed from t h e wavelet coefficients? T h e answers t o b o t h of
these questions are clearly yes as evidenced by t h e voluminous literature
in recent years. In this paper we rely on t h e affirmative answer t o t h e first
question a n d concentrate on exploring t h e wavelet transform of measured
wind waves. It is an exciting a n d fruitful area for practical application of
the wavelet transform . As d a t a analysis on wind wave studies comprises
mainly of applications of statistics and Fourier transforms, t h e s u m m a r y
shown in Table 1 indicates t h a t wavelet transform analysis is a logical ex-
tension t o t h e currently available analyses.
In analogy with Fourier energy density spectrum, we can readily define
a wavelet spectrum for a data series X(t) as
and accordingly,
T ( s τ ) = WXY(S,T)
VWXk(s,T)Wyk(s,T)
and
WXk(s,T)WYk(s,T)
as t h e complex-valued wavelet coherency and its square, t h e real valued
wavelet coherence, respectively, between t h e two d a t a sets. T h e functions
S W x y ( s , r ) a n d 3 W x y ( s , r ) in (10) are respectively t h e real a n d imagi-
nary parts of WXY (s, r ) , and hence t h e co- and quadrature- wavelet spectra
of X(t) and Y(t).
154 P. Liu
Here we should point out that this wavelet is not an admissible wavelet
since a correction term is needed because φ(0) Φ 0. However, in prac-
tice choice of a large enough value for the parameter m, (e.g. m > 5)
generally renders the correction term negligible. In this study, we follow
Daubechies [5] and use m = π v/27^2. While there are admissible wavelets
available, the Morlet wavelet has been widely used in signal analysis and
sound pattern studies. Aside from its convenient formulation and histori-
cal significance, its localized frequency is independent of time, a feature of
particular advantage for wind wave studies.
Wavelet Spectrum
0.3 '
N0.25 - -
fr 0.2 r*-~S\ Λ Γ\ '
; Ç^àVl â ^ Aß^k
Λ Λ Λ^
Φ
|0·15
u- 0.1
0.05 - 1 1
-
1650 1750 1800
time s
Figure 1. A sample plot of a time series of wind waves and its respec-
tive wavelet spectrum.
§3. Applications
In the following three subsections we present three wavelet transform
analyses of wind wave data leading to distinct results that would be dif-
ficult, if not impossible, to obtain from the usual Fourier transform. The
data used in the applications were measured during the recent SWADE
(Surface Wave Dynamics Experiment) program [25]. The wind and wave
data were recorded from a 3 m discus buoy during the severe storm of Oc-
tober 26, 1990. The buoy was located at latitude 38°22.1' N and longitude
73°38.9' W, with a water depth of 115 m near the edge of the continental
156 P. Liu
shelf offshore of Virginia in the Atlantic Ocean. Time series of wind and
waves were both recorded at 1 Hz from a combined design of a three-axis
accelerometer and magnetometer along with the Datawell Hippy system.
A total of 100 sets of data, each 1024 s in length, were used in the anal-
yses. The data, predominantly wind-generated waves, covered the entire
duration of the storm with wind speeds ranging from calm to 18 m/s and
significant wave heights approaching 7 m.
(iv) The dominant group wave height, hp, which can be obtained from
the time series as the maximum trough-to-crest wave height over
the time length tg.
The variability of these parameters indicates that wave groups are ap-
parently diverse, irregular, nonperiodic, and independent from each other.
The formidable task is to determine the significance and usefulness of these
parameters. Here we consider a simplified approach of forming two nor-
malized parameters:
wave group tends to generate more waves in the group. This interesting
result, while intuitively understandable, is new.
A scatter plot of averages of dominant group wave heights versus sig-
nificant wave heights is shown in Figure 3. The significant wave height,
defined as the average of the highest one-third wave heights in the wave
record, is a familiar and widely-used parameter. For practical applications,
such as in engineering design, mean dominant group wave height would be
more pertinent than the significant wave height. Figure 3 shows that sig-
nificant wave heights are slightly less than the averages of dominant group
wave heights.
growth.
How do wind waves grow? It is a question that several generations of
scientists have addressed. In addition to the early work of Jeffreys [11] and
Ursell's [23] famous "nothing very satisfying" summary, modern conceptual
perceptions of wind waves primarily stem from the theoretical conjectures
of Phillips [17], Miles [15], and Hasselmann [8]. The current proliferation
of numerical wave models is basically developed from these early theories.
Numerous measurements of wave energy spectra with average wind speeds
have been conducted for the validation and possible enhancement of the
available models. Now with the latest SWADE measurements and the
advancement of wavelet transforms, we are able to examine wind wave
processes from new perspectives.
Wavelet Spectrum
plots can tell us. T h e five separate graphs in Figure 5 display, respectively
from top down, t h e wavelet spectrum for wind, t h e wavelet spectrum for
waves, t h e real p a r t , t h e imaginary p a r t , and t h e phase of wavelet coher-
ence. All of t h e plots contain the three frequency components of 0.1131,
0.1199, and 0.127 Hz for which the energy density is highest.
Note t h a t in Figure 1 there are five groups of waves t h a t can b e iden-
tified from t h e wavelet spectrum. In t h e second graph of Figure 5 in which
energy densities increase and decrease with respect t o time, only three
stronger groups (i.e., at time marks 1570, 1630, and 1695) are reflected
from the fluctations of these frequency components. T h e top graph of Fig-
ure 5 shows t h a t the wavelet spectrum components for wind speeds exhibit
similar, b u t more, energy fluctations with time. Some of the fluctuations
correspond closely to those of the waves. By examining the b o t t o m three
graphs of Figure 5, it shows quite clearly t h a t for t h e three wave groups
identified with appreciable energy contents, the real p a r t of their coherence
is close t o 1 1, their imaginary part close to 0, and their phase is also close
to 0. Therefore, during wave growth, t h e frequency components for peak
wave energy between wind and waves are inherently in phase. Wave groups
constitute t h e basic elements of wind wave processes, and the wave growth
are primarily taking place within the wave group.
As t h e growth of wind waves is an extremely complicated process,
the above results contribute still qualitatively toward an understanding of
t h e n a t u r e of how do waves grow. While we are accustomed to correlate
wave growth with "average" wind speeds, the results presented here clearly
show t h a t waves are in fact responding to wind speeds instantly. Further
detailed studies may challenge or counter more familiar notions of wind
waves. Using cross wavelet spectrum analysis not only introduces new
d a t a analysis techniques, it may also leads to new courses of exploration.
in the laboratory and in the field, have been done with specialized methods
based on radar reflectivity, optical contrast, or acoustic output of the ocean
surface. Here we show that with the help of wavelet spectra [12], instead
of using specialized measurement devices, a basic wave-breaking criterion
can be easily implemented to wind wave time series to distinguish breaking
from non-breaking waves. This simple and fairly efficient approach can
be readily applied to indirectly estimate wave breaking statistics from any
available time series of wind-generated waves.
m
X
|x X
x
o x 1 ° * ° x II1 ï 1 *
_6I , ,i , , . ..,_. . ,1 1
1500 1550 1600 1650 1700 1750 1800
Time s
One of the most frequently used approaches for the study of wave break-
ing is the use of a limiting value of the wave steepness beyond which the
surface cannot be sustained [13]. Alternatively, assuming a linear dispersion
relationship, the wave surface will break when its downward acceleration
exceeds a limiting fraction, 7, of the gravitational acceleration, g, that is
ασ2 ^ 7#. The quantity ασ 2 can be calculated for a time series of wave
data since the local wave amplitude, a, is available from the measured time
series while the local wave frequency, σ, can be obtained from the wavelet
Wavelet Spectrum Analysis 163
18|-
16
ι
I12 o o
o°o
ce
2> 9p°
ω 10
O X
9> xä
X
'χΧ o x
**
I Ή H nil
4 6 8 10 12
Wind Speed m/s
frequency, and λ is a number greater than 1 that denotes the start of the
high frequency range beyond ωρ. The exact location of this high frequency
range has not been clearly defined. Considering this range as corresponding
to the familiar equilibrium range, one frequently used value of λ has been
1.35 [7].
To test this approach, Figure 6 presents an illustration of the analysis
where estimated breaking waves are marked on the same time series seg-
ment given in figure 1. The x's and o's represent the results with a high
frequency range between 1.15 and 1.35 times, respectively, of the local peak
energy frequency, CJP, and cut-off frequency, ωη. While the λ values of 1.15
or 1.35 has been chosen rather arbitrarily for comparisons, they are clearly
not always recognizing the same breaking waves. In general with the same
cut-off frequency, the lower end of the frequency range farther away from
the local peak frequency, i.e. large λ value, would yield higher local average
frequency σ and more breaking waves. Therefore an exploration of break-
ing waves could potentially serve to resolve the definition of the well-known
but still not yet well-defined equilibrium range. Figures 7 present plots of
overall percentages of breaking waves from all the data analyzed in this
study as a function of wind speed. While the data points are scattered
considerably, there is an approximate linear trend indicating an increase
in the percentage of breaking waves with an increase in wind speed. The
results shown in Figure 7 are in general accord with various available ob-
servations [9]. According to these results, breaking waves become prevalent
when wind speeds exceed 10 m/s.
At the present, the limiting fraction of downward wave acceleration
from the gravitational acceleration, 7, and the parameter locating the lo-
cal equilibrium range beyond local peak frequency, λ, are both tentative.
Therefore, the wavelet transform approach that leads to these results is
useful, convenient, and also exploratory. Perhaps a better simultaneous
measurement of wind-wave time series and wave breaking would suffice to
substantiate the approach. Unfortunately operational and sufficient instru-
ment for this simple purpose is still lacking.
References
1. Agrawal, Y. C., E. A. Terray, M. A. Donelan, P. A. Hwang, A. J.
Williams III, W. M. Drennan, K. K. K a h m a , and S. A. Kitaigorodiski,
Enhanced dissipation of kinetic energy b e n e a t h surface waves, Nature,
3 5 9 , 219-220, 1992.
2. Banner, M. L., and D. H. Peregrine, Wave breaking in deep water,
Annu. Rev. Fluid Mech., 25, 373-397, 1993.
3. Combes, J. A., A. Grossmann, and P h . Tchamitchian (Eds.), Wavelets,
Time-Frequency Methods and Phase Space, 2nd ed. Springer-Verlag,
1989.
4. Daubechies, I., Ten Lectures on Wavelets, Society of Industrial and
Applied Mathematics, 1992.
5. Daubechies, I., T h e wavelet transform, time-frequency localization and
signal analysis, IEEE Trans. Inform. Theory, 3 6 , 961-1005, 1990.
6. Farge, M., Wavelet transforms and their applications t o turbulence,
Annu. Rev. Fluid Mech., 24, 395-457, 1992.
7. Günther, H., W. Rosenthal, T. J. Weare, B. A. Worthington, K. Hassel-
m a n n , and J. A. Ewing, A hybrid parametrical wave prediction model,
J. Geophys. Res., 8 4 , 5727-5738, 1979.
8. Hasselmann, K., On the non-linear energy transfer in a gravity-wave
spectrum, P a r t 1, General theory, J. Fluid Mech., 12, 481-500, 1962.
9. Holthuijsen, L. H., and T. H. C. Herbers, Statistics of breaking waves
observed as whitecaps in the open sea, J. Phys. Oceanogr., 16, 290-297,
1986.
10. Hwang, P. A., D. Xu, and J. Wu, Breaking of wind-generated waves:
measurements and characteristics, J. Fluid Mech., 2 0 2 , 177-200, 1989.
11. Jeffreys, H., On the formation of water waves by wind, Proc. Roy. Soc,
bf A 107, 189-206, 1925.
12. Liu, P. C , Estimating breaking wave statistics from wind-wave time
series data, Annales Geophysicae, 4, 970-972, 1993.
13. Longuet-Higgins, M. S., On wave breaking and the equilibrium spec-
t r u m of wind -generating waves, Proc. Roy. Soc, A 3 1 0 , 151-159,
1969.
14. Masson, D. and P. Chandler, Wave groups, a closer look at spectral
methods, Coastal Engineering, 20, 249-275, 1993.
15. Miles, J. W., On t h e generation of surface waves by shear flows, J.
Fluid Mech., 3 , 185-204, 1957.
16. Morlet, J., G. Arens, I. Fourgeau, and D. Giard, Wave propagation
and sampling theory, Geophysics, 4 7 , 203-236, 1982.
166 P. Liu
Sarah A. Little
Abstract. 1-D wavelet analysis has been shown to be useful in studying bathy-
métrie profiles [7]. 2-D bathymétrie maps are less common than 1-D profiles,
but offer immensely more information about seafloor generation processes. 2-
D wavelet analysis is applied to swath-mapped bathymétrie data from the Mid-
Atlantic Ridge. Both image enhancement and feature identification are performed
with excellent results in the identification of the location and scarp facing direction
of ridge-parallel faulting. Wavelet image processing techniques enable computer
analysis of distribution and spatial patterns in faults to be performed without the
tedious job of transcribing hand picked and ruler-measured fault parameters from
printed images to a digital data base.
§1. I n t r o d u c t i o n
T h e wavelet decomposition of bathymétrie d a t a reveals structures and
p a t t e r n s which are easily overlooked in the raw data. In addition, it can
be used to isolate features of interest, such as fault scarps, for use in subse-
quent quantitative analysis. This paper describes the application of wavelet
analysis to seafloor topography. Much of seafloor bathymetry has been col-
lected by ships of opportunity traversing various sections of ocean. These
depth measurements are closely spaced in the along-track direction, b u t
t h e tracks are often many kilometers apart. T h e resultant d a t a sets are es-
sentially 1-dimensional spatial series. In a few areas of the seafloor, swath
bathymétrie surveys have been conducted which return detailed, 2-D, maps
of limited sections of the ocean. 1-D spatial series are more common, and
1-D wavelet analysis of bathymétrie profiles can be used to improve our un-
derstanding of the shape of the seafloor. Swath m a p p e d areas have much
more information t h a n single topographic profiles, of course, and I present
an example of image enhancement and fault scarp identification using 2-D
wavelet analysis.
w{t) = e^ikth-£
where t is space, k is wavenumber, and σ is the Gaussian variance.
A single scale wavelet transform is the convolution of t h e wavelet at
t h a t scale with the data, b o t h of which are functions of the spatial domain.
For increased computational speed, this convolution may be computed in
the Fourier domain by multiplying the Fourier transform of the d a t a by
the Fourier transform of the wavelet, then taking the inverse transform of
the product. On a single scale, the wavelet transform can be thought of
as a linear filter. For a multiscale analysis, the wavelet shape is rescaled,
in powers of 2, to longer and shorter lengths, creating a bank of filters
of different sizes. T h e power in each size wavelet is normalized to one,
and then the wavelet is convolved with the data. In this way the series of
wavelet filters is used to scan for small- and large-scale events in the data.
W h e n a wavelet of a given size and shape is convolved with a spa-
tial series, the magnitude of the resultant series is a measure of the m a t c h
between the wavelet and d a t a series—small-scale events will m a t c h small
wavelets b u t not large wavelets. Therefore, when a bank of wavelets is
applied to a spatial series, one can b o t h identify transient events and dis-
tinguish events on different scales. In this way one can scan a d a t a series
for interesting features on a broad range of scales, and also identify re-
gions of t h e d a t a where events of a certain scale are entirely missing. Little
et. al. [7] used the Morlet wavelet transform, in this way, to discover t h e
anomalous low-frequency zone in a 1-D bathymétrie profile.
§3. 2 - D B a t h y m e t r y
T h e following section describes the application of wavelets generated
from spline functions to image enhancement and fault scarp identification
in 2-D swath-mapped bathymétrie data. T h e bathymétrie d a t a cover an
approximately 100 km x 70 km section of Mid-Atlantic Ridge near 29°N,
43°W, which is about 1/10 of the d a t a available in this region [13].
Topography near the Mid-Atlantic Ridge (MAR) is characterized by a
deep central median valley surrounded by faulted blocks leading out over
crestal mountains and down the outer flanks. These faulted blocks tend
to b e long linear ridges running parallel to t h e axis of spreading, ranging
in width from order of 10 m to 10 km. Swath m a p p e d areas of the M A R
reveal scales ranging from order of 100 m up to 100 km, and hence offer a
reasonable d a t a set with which to examine this major faulting process at
the MAR. Two major questions to ask about the M A R are 1) what does
170 S. Little
3.1. Data
T h e d a t a used in this analysis was collected by a 16-beam Sea Beam
swath echo sounder. This instrument is mounted on the hull of a ship
and sends out sound in beams whose individual footprints are 2 | x 2 |
degrees, which translates to about 150 m on a side in a water depth of
3000 m. As the ship steams forward, it continuously collects an approxi-
mately 2 km wide swath of depth readings. These d a t a are tied to latitude
and longitude via satellite navigation and put together in a regional m a p
of bathymetry. There are places where the swaths do not overlap, and a
linear interpolation has been performed in these gaps so t h a t the d a t a can
be smoothly represented in a 2-D matrix of depth values. T h e d a t a are
shown in Figure 1 in a contour/gray scale image. There are 755 points
of longitude between 43.567°W and 42.552°W, and 495 points of latitude
between 28.581°N and 29.917°N. Depths range from 4000 m in the deep-
est areas (black) to 1800 m in the shallowest areas (white). There are two
white areas inside the image which represent no data, a large one centered
on longitude point number 200 and latitude point number 50, and a small
gap located at longitude point number 240, latitude point number 100. T h e
white areas to the left and right of the image also contain no bathymétrie
information. T h e central valley of the M A R runs up the middle of the
image at about 20° off vertical. On each side of the central valley are t h e
crestal mountains, and beyond them a small portion of the flanks. This
area of the Mid-Atlantic Ridge contains a ridge offset, visible in t h e upper
central part of the image, where the two major valleys are offset from one
another. T h e d a t a used in this paper have previously been analyzed for
fault locations using a curvature method [14], [15].
-1500
g -2000
f, -2500
Û
-3000
-4000
100 200 300 400 500
Sample point number, Longitude
3 . 3 . Linear B - s p l i n e w a v e l e t
Spline wavelets are constructed by a shifting, weighting and summing
of the B-spline functions, Ni(x), where x is the spatial variable, and i is
the order of the spline. Details of the construction of the linear and other
B-spline wavelets can be found in [2],
T h e linear B-spline function, N2(x), is defined as:
r o x <0
0<x < 1
1 <x < 2
ι o x>2.
Figure 3a shows this wavelet (at 16 points) and the 1-D convolution
of this wavelet with three, noise free, 1-D ridges of different sizes. Notice
t h a t , in the transform, each step edge is converted to a low and a high.
T h e ridge whose width is most similar to the central portion of t h e wavelet
is transformed into a single high with two lower amplitude lows on either
side. This creates contrast which will visually "bring out" ridges.
T h e 2-dimensional filter is formulated to take advantage of the strongly
linear and oriented n a t u r e of the faults. Each linear ridge is qualitatively
like a long series of 1-D ridge-edges stacked together. Therefore, a 2-D filter
can be built by stacking up a number of 1-D ridge-edge detecting wavelets.
This is a departure from strict wavelet construction because the ridge-edge
parallel direction of this 2-D filter is not a wavelet. T h e 2-dimensional
version of t h e filter is specifically created in the following way: generate a
1-dimensional linear B-spline wavelet of a desired length; replicate this filter
to create a 2-dimensional matrix of identical filters; taper the edge-parallel
direction with a Hanning window [12]; and rotate the entire m a t r i x to t h e
desired orientation. A 3-D plot, contour/gray scale plot, and horizontal and
vertical profiles are presented in Figure 4 for this filter. Figure 4e shows
the filter rotated by —110 degrees from horizontal to match the dominant
orientation of the faults in the bathymetry. T h e length of the 1-dimensional
B-spline (BS) determines the width of the faults to be isolated. T h e number
of identical filters (NF) in the 2-dimensional matrix determines t h e fault
length over which to average (or smooth). If N F is large, then very long
174 S. Little
and straight faults will be identified; if N F is small, short and curved faults
will be identified as well as long and straight. A small N F , although it picks
up b o t h short and long faults, also picks up short curved features which are
volcanic rather t h a n faults, and hence there is a trade-off between getting
all the curved faults and getting too many short, non-fault features.
T h e 2-D linear B-spline filter, with BS=16 and N F = 7 is convolved, via
the 2-D Fourier transform, with the Sea Beam bathymetry and the results
shown in gray scale in Figure 5. Light areas correspond t o topographic
ridges, and dark areas correspond loosely to valleys. T h e fabric of t h e
Wavelet Analysis of Seafloor Bathymetry 175
3 0.5
1 o (c)
< -0.5
5 10
Sample point
(a) NF direction, NF=7
(d)
BS direction 2 4 6
Sample point
Rotated-110
(b) (e)
Figure 4. a) A 3-D mesh plot of the 2-D linear B-spline filter with
BS=16 and NF=7. b) Contour/gray scale plot of this filter, dark is
high, c) Profile of this filter in the B-spline direction, d) Profile of
this filter in the NF direction, e) Contour/gray scale plot of this filter
rotated by —110° to match the dominant orientation of MAR-parallel
fault ridges.
4 6
(a) Sample point
NF direction, NF=7
(d)
BS direction 10 0 NF direction 2 4 6
Sample point
Contour Rotated-110
(b) (e)
0 £<0
0<x < 1
1 <x < 2
N4(x) = { § - ( x - 2 ) 2 + I(*-2)3 2< x < 3
|_I X + I ( X _3)2_I( X _3)3 3 < x <4
0 x >4.
A sixteen point long, 1-D filter created from the derivative of this
function is shown in Figure 3b. This filter is convolved with ridge edges of
three different sizes to show how it transforms a rising slope to a numerical
high, and a falling slope to a numerical low. In this way it locates fault
scarps and identifies t h e direction they face.
T h e 2-D filter is created from the derivative of the cubic B-spline t h e
same way as for the linear B-spline wavelet: Generate a 1-D filter of a de-
178 S. Little
■' /
§4. C o n c l u s i o n s
Bathymétrie d a t a from the seafloor contain the superposition of ridges,
valleys, and volcanos at many different scales. Wavelet analysis offers a use-
ful m e t h o d for decomposing the texture of the seafloor to help understand
processes which occur at many different scales. Much of the bathymétrie
d a t a available is in the form of 1-D spatial series. Scale decomposition
of these series has been shown to yield an understanding of geophysical
processes which originally were overlooked in the raw data. Smaller por-
tions of the seafloor have been swath-mapped, and these 2-D d a t a sets
offer an unusually good opportunity to decipher the complexity of seafloor
topography. This paper has presented two wavelet techniques for working
with swath-mapped bathymetry, one a qualitative image enhancement, and
the other a quantitative fault scarp identifier. A linear B-spline wavelet is
used to design a 2-D filter which produces an image in which the linear,
MAR-parallel fault ridges are easily seen in the context of the broader to-
Wavelet Analysis of Seafloor Bathymetry 181
50
100
150
I
S 200
350
400
450
ι - , t1 .
100 200 300 400 500 600 700
Sample point number
References
1. Canny, J., A computational approach to edge detection, Trans, on
Pattern Anal, and Machine Intelligence, 8(6), 679-698, 1986.
2. Chui, C. K., An Introduction to Wavelets, Academic Press, New York,
1992.
3. Daubechies, L, Orthonormal bases of compactly supported wavelets,
Comm. Pure Appl. Math. , 4 1 , 909-996, 1988.
4. Daubechies, I., Ten lectures on wavelets, CBMS Regional Conference
182 S. Little
Sarah A. Little
Department of Geology and Geophysics
Woods Hole Océanographie Institution
Woods Hole
MA 02543, USA
e-mail: [email protected]
A n a l y s i s of H i g h R e s o l u t i o n M a r i n e S e i s m i c D a t a U s i n g t h e
W a v e l e t Transform
Chris J. Pike
§1. I n t r o d u c t i o n
T h e Centre for Cold Ocean Resources Engineering (C-CORE) is an
independently-funded research institute of Memorial University of New-
foundland in St. John's, Newfoundland, Canada. C-CORE is involved in
solving engineering problems related to resource development in the ocean
environment. T h e Centre has traditionally focused on a limited number of
research areas related to ocean resources and has built up expertise reflected
by t h e four research groups operating at C-CORE. They are t h e R e m o t e
Sensing group, t h e Geotechnical Engineering group, the Ice Engineering
group and t h e Seabed Geophysics group.
One of the areas of research engaged by the Seabed Geophysics group
has been t h e high resolution acoustic investigation of the sub-seabed. A
central goal of the acoustic program is the development of methods for
extracting marine soil properties by correlating acoustic properties derived
from acoustic signals with the geotechnical properties of the soil. During
t h e course of this research it was evident t h a t the high frequency acoustic
signals (500 Hz to 10,000 Hz) used were rapidly attenuated, especially the
higher frequencies, in those soils that contained many scatterers - boulders
and cobbles with dimensions comparable to the wavelengths of the acoustic
waves generated by the source. Spectra for the acoustic signals are usually
calculated by using the Fourier transform (FT) or the fast Fourier transform
(FFT). From the equation for the Fourier transform it is evident that this
equation is not well suited to the study of localized (in time) disturbances
in a signal nor is it appropriate to apply it to non-stationary time series.
The power spectrum of an acoustic signal is often analyzed in an attempt
to investigate attenuation of an acoustic signal. Techniques that compare
the spectra of the emitted and transmitted or reflected pulse are used to
assess attenuation as a function of frequency. The typical frequency range
for explortion seismic data is between 10 and 100 Hz. Over this range,
and for the length of duration of the seismic record, the basic assumptions
of stationarity are not grossly violated. High resolution seismic signals
can often exhibit severe loss of high frequencies as they travel down into
the earth and return again. A method that reflected the time-frequency
structure in high resolution signals would be more diagnostic and more
appropriate for the analysis of very high resolution signals.
Experiments designed to measure attenuation within the earth some-
times use down hole geophones to record the direct pulse from a source
near the surface or a core sample may be taken to the laboratory where
a high frequency acoustic wave is transmitted through the sample. If the
outgoing pulse and the transmitted pulse can be isolated then the spectra
can be compared. The isolation of the first arrival is usually accomplished
by using a short window Fourier transform (SWFT) and often the spectra
exhibit properties associated with the window function used as well as the
signal. The wavelet transform offers a means of avoiding this problem.
High resolving power of the acoustic wave is desirable for a range of
depths but this ability is compromised when multiple scatterers are present.
Morlet et al. ([21] and [22]) investigated the propagation of plane waves for
normal incidence, through periodic multi-layered media for wavelengths
ranging from much greater than the spatial period to periods on the or-
der of the spatial period. They found that for large wavelengths (16 times
the spatial period of the medium) the composite medium was transpar-
ent but phase-delaying; for short wavelengths (2 times the spatial period
of the medium) the signal was strongly attenuated and super-reflectivity
could occur; for intermediate wavelengths (8 times the spatial period of the
medium) velocity dispersion versus frequency appeared.
These observations are quite diagnostic of the medium, offering the geo-
physical investigators of shallow marine soils a new and potentially pow-
erful means of quantifying subseabed soil conditions through the use of
time-frequency methods. Morlet et al. ([21] and [22]) conclude that when
High Resolution Marine Seismic Data Analysis 185
§2. A c o u s t i c - G e o t e c h n i c a l C o r r e l a t i o n s : P h y s i c a l a n d H i s t o r i c a l
Context
T h e motivation for research into correlations between acoustic/seismic
responses of a marine soil and the soil's geotechnical properties is to provide
the geotechnical engineer with cost-effective and reliable estimates of t h e
geotechnical properties of submerged soils. Reliable estimates of geotech-
nical properties are needed, for example, in the selection of a site and t h e
design of a structure t h a t is intended to be supported by the seabed. Such
structures can include gravity-based oil production platforms or support
columns for causeways or bridges. Porosity, density and grain size distri-
bution are some of the soil properties t h a t affect the acoustic response of
a signal and are also of interest to the geotechnical engineer [31]. Acoustic
wave properties related to these parameters through empirical relationships
include acoustic impedance, velocity and attenuation ([1], [9], [20] and [13]).
These properties can be determined from digital acoustic d a t a with greater
accuracy t h a n was possible in the past using analogue records.
T h e recent and rapid development of smaller, faster, and more power-
ful computers has had an impact upon many branches of science b u t none
so dramatic, perhaps, as in the area of high resolution sub-bottom marine
seismic reflection profiling. At about the same time as the first papers on
wavelet transforms were being written, analogue paper recordings of acous-
tic returns from t h e seabed and sub-seabed represented t h e current practice
in d a t a collection and presentation. Practitioners of land-based engineer-
ing seismic d a t a acquisition, and the oil and gas exploration industry, have
186 a Pike
been acquiring and processing digital data for almost as many years as there
have been computers. During the past ten years the marine high-resolution
(> 500 Hz) or site surveying industry has made the leap from analogue data
acquisition systems to digital data acquisition systems. This technological
leap can be attributed to the development of high-speed analogue-to-digital
computer boards and digital signal-processing chips, which together effec-
tively allow real-time data acquisition and processing of high-resolution
acoustic data.
The availability of digital acoustic data has led to the desire for a more
quantitative assessment of seismic signals. Empirical relationships between
acoustic properties and soil properties have been reported ([1], [32] and
[10]). Studies that relate acoustic properties to rock or sediment properties
involve the use of frequencies below 100 Hz ([26] and [12]) or above 10
kHz ([10] and [11]) but very few studies provide information in the 100
to 10,000 Hz range which is the typical range for many high-resolution
sub-bottom marine surveys. In soils characterized by large scatterers, such
as boulders in glacial tills, the 100-1,000 Hz range of frequencies is more
efficient in achieving greater depth penetration. Testing carried out in
laboratories, usually at very high (> 100 kHz) frequencies, on samples
collected in the field or on cone-penetrometer data acquired in situ, provide
data for correlations that are sometimes biased toward high frequencies.
But the means to correlate geotechnical properties of a soil over a more
complete range of frequencies is also needed. Some commercial acoustic
systems already offer classification schemes based upon how the transmitted
pulse is modified by the water-sediment interface upon reflection [18] or
internally upon transmission to deeper layers and subsequent return to the
receiver [29]. It is the seafloor sediment type that is most easily determined
by evaluating the seabed return as in Leblanc et al. [18] but characterization
derived from deeper returns becomes quite complex due to constructive and
destructive interference by scattered waves. Time-frequency analyses would
aid in assessing the degree of interference as well as capitalizing upon the
tuning or detuning of acoustic waves that can occur within the sub-bottom
layers.
There are several factors that contribute to the attenuation of an acous-
tic wave. Acoustic energy attenuates with increasing distance from the
source, a function described by the term geometrical spreading loss. Ge-
ometrical spreading losses are not related directly to the properties of the
medium in which the energy is propagating. Usually a means of correcting
or nullifying this effect is sought before attenuation due to the medium is
determined. Attenuation of acoustic energy can be divided into two basic
categories, intrinsic attenuation and apparent attenuation. Intrinsic atten-
uation is the process that degrades the amplitude of the signal such as
absorption due to frictional heating or viscous losses [14]. Apparent atten-
High Resolution Marine Seismic Data Analysis 187
1
FT(9b(t)) = Φ 6 (ω) = e ^ ^ - e~* e ^ ^ . (3)
This wavelet is shown in Figure 1. The envelope of the wavelet and its
amplitude spectrum are Gaussian-modulated functions. The variable b in
Equations (2) and (3) has the same value of 5.336 and is the same as t h a t
given by Goupillaud et al. [6].
High Resolution Marine Seismic Data Analysis 189
0.001 0.002 0.003 0.004 0.005 0.006 0.007 0.008 0.009 0.01
Time (sec)
relevant part of the derivation of the spectral ratio equation. After John-
ston and Toksöz [16], a plane wave propagating in a homogeneous medium
can be described by:
A(x,t) = A0e^kx-^ (5)
where attenuation may be represented by allowing either A:, the wavenum-
ber, or ω, the angular frequency, to be complex. Following [16] and in-
cluding a geometrical spreading term as in [34], the ratio of the reference
amplitude spectrum versus the spectrum at some time, ίχ, after taking the
natural logarithm of the ratio of the two amplitude spectra can be written
as:
A G
In —^ = (7 m - 7i)vtf + In -^ (6)
A\ Cxi
where v is velocity, t is one-way travel time, / is frequency, a(f) = 7 / is
the attenuation at / , and the left hand side of Equation (6), the spectral
ratio, is represented by the data displayed in Figure 2c. Equation (6) can
be interpreted in two ways; first the free variable can be the frequency, / ,
and then the slope, of a line fitted to the data when the spectral ratio is
plotted against frequency, would be:
slope = (7 m - 7i)ttf. (7)
The geometrical term, ln(G? i/G2), can be ignored if one assumes that the
geometrical spreading terms are independent of frequency. Alternatively,
travel time can be taken as the free variable but in this case the geometrical
factor does depend upon t and it will influence the value of the attenuation
derived from the slope of a line fitted to the curve of the spectral ratio
versus time calculation.
For this study the spectral ratio method was applied by taking the
frequency distribution for the time at which the maximum magnitude for
the direct pulse is greatest and using this spectrum as the reference spec-
trum. This occurs at approximately 0.0007 sec. in Figure 2b. Therefore,
any determination of the slopes of lines fitted to the data in the direction
parallel to the vertical axis satisfies Equation (7) and those that are fitted
to the data in the direction parallel to the horizontal axis are subject to
the time-dependency condition.
The procedure for estimating the attenuation from normally-incident
acoustic waves is demonstrated using the simple synthetic seismic example
described in the preceding text. The results of a comparison of the value
of the attenuation obtained using the synthetic trace corrected for geomet-
rical spreading losses and an uncorrected synthetic trace are displayed in
Figures 3a and 3b respectively. The synthetic trace was corrected for the
known geometrical spreading factor and the wavelet transform was then
calculated and the modulus displayed (Figure 2b). The mean power at
192 C. Pike
each time and for all frequencies in the transform was calculated and plot-
ted in Figure 3a. There are five broad peaks shown in Figure 3a associated
with each of the attenuated pulses in Figure 2a. The largest local m a x i m a
for a particular broad peak was chosen interactively and a least squares
fit was made to these selected values. The criterion for selecting the local
maxima is based upon providing a guideline for consistent picking, and in
selecting the large broad peaks it was assumed t h a t boundaries or acoustic
reflectors will reflect energy independent of frequency content which would
not be the case for random scatterers. The line and the points chosen
for the least squares fit are indicated (straight black line and black x's re-
spectively). T h e value of the slope was found to be 360 nepers/sec which,
when converted to meters by dividing out the velocity term, yields a value
of 0.24 nepers/meter. The conversion between nepers/unit length and dB
(decibels) per unit length is a (dB/unit length) = 8.686π (nepers/unit
length). The actual value used was 0.25 nepers/meter, a 4% difference be-
tween the actual and estimated value. If the geometrical spreading losses
are not removed then there should be some effect on the result for the
attenuation calculation, and this is observed in Figure 3b. Using the spec-
tral ratio d a t a from Figure 2c, the variance in the direction parallel to
the frequency axis was calculated for every time sample. The slope of a
least squares fit to the user-selected points from this plot yielded a value of
277 nepers/sec or 0.185 nepers/meter, a 27% difference. When all digital
sample values are considered, the least squares estimate for the corrected
d a t a is 0.29 nepers/meter, a 16% difference. As a final comparison t h e
primary event was windowed (150-point window) as well as t h e event at
0.008 sec. T h e amplitude spectrum was calculated for each event using t h e
fast Fourier transform and then the difference was divided by the separa-
tion between the two events yielding a value of 0.22 nepers/meter or a 13%
difference between the actual and estimated value.
T h e results given in Figures 3a and 3b are very encouraging, and in
spite of the large difference (16%) for the case where all values for a cor-
rected transform were used an order of magnitude for the attenuation is
nevertheless achievable. For real acoustic signals the zones or valleys in
Figure 3b, where the spectral ratio falls below 5 nepers with respect to
the primary signal, the interference due to scattering of acoustic energy
would tend to fill in these valleys possibly resulting in a lower value for the
attenuation estimate when including all values in the estimation.
Application of the wavelet transform to the problem of determining
attenuation offers the potential for automated order-of-magnitude deter-
minations of attenuation from digitally-recorded acoustic signals. As well,
the potential to surgically isolate energy in time and frequency can enhance
its application for sub-bottom attenuation measurements using reflected
seismic signals.
High Resolution Marine Seismic Data Analysis 193
Time (sec)
Figure 4. (a) Site plan for experimental field program and (b) cross-
section of seabed recording geometry.
196 a Pike
tic source was then moved to stations 6, 7 and 8 with the above acquisition
procedure repeated for each source position. After the shot at station 8 the
entire b e a m was rotated about its central point in 45 degree increments and
the recording process described above was repeated. The complete source
receiver coverage is shown in Figure 4a. The four recorded profiles are la-
belled Beams 1 to 4 and station number 1 is identified for each b e a m with
a total of twelve stations per beam. The outer ring of sparker source point
locations, at a radius of 2.13 m, defines the extent of the subsurface normal
incidence coverage. T h e common depth point (CDP) coverage is defined
by a circle of radius 4.87 m, passing through stations 3 and 10 on beams 1
through 4.
T h e entire procedure described above was carried out for two separate
power settings, 480 J and 1080 J. Thus two complete surveys were acquired
for this site, and are referred to as the low and high power surveys.
T h e physical-sampling program provided marine soils and bedrock
cores for analysis and a geotechnical program provided some in situ mea-
surements using a limited series of cone penetrometer tests (CPT's) as well
as a series of seismic cone penetrometer tests (SCPT's) to acquire velocity
information. Figure 4a shows the locations of the C P T ' s and boreholes.
T h e cone penetrometer was capable of recording shear (S) and com-
pressional (P) wave data. This information was acquired when the cone
penetrometer was stopped at 1 meter intervals as the cone was pushed into
the seabed. T h e compressional wave arrival times were used to calculate
average and pseudo interval velocity values for the soils. T h e horizontal
offset between the three acoustic sources and the cone rods and the off-
set of the geophones from the cone tip were accounted for in all velocity
calculations.
Three continuously-sampled boreholes were drilled by rotary techniques,
using N-size casing. Standard penetration tests were conducted and repre-
sentative b u t disturbed soil samples were obtained in the overburden, using
a conventional 51 m m OD split-spoon (SS) sampler. Bedrock was cored in
NQ size. All borehole depths reported are with respect to the sea floor.
The exact hole locations were measured with respect to the underwater
acoustic survey by divers.
§5. D a t a A n a l y s i s
Acoustic d a t a analysis was divided into two phases with the first fo-
cused on determining subsurface stratigraphy by using the multichannel
acoustic data, the subsurface samples and results from the penetrometer
tests. T h e second phase involved analysis of the normal-incidence d a t a and
was concerned with measuring and quantifying changes in the waveform of
the reflected energy as well as finding some means to visually compare t h e
High Resolution Marine Seismic Data Analysis 197
5 . 1 . C o n v e n t i o n a l d a t a analysis a n d p r o c e s s i n g
Core samples extracted from the site confirmed the composition of
the marine sediments for this area. T h e sediments consisted of a fine to
medium-grained gravelly sand with some organic content overlying a sand-
gravel deposit of between 4.2 and 4.6 m thick. Sandwiched between t h e
fractured, layered bedrock and the sand-gravel deposit was a thin ( < 1.0
meter) till. T h e depth to bedrock ranged between 4.7 and 5.1 m below t h e
sea-bot torn.
T h e cone penetrometer d a t a were of limited value as this tool is de-
signed for use with fine-grained soils. However, some useful d a t a included
the cone tip resistance which provided information on soil stiffness. T h e
seismic d a t a were used to estimate independently the compressional (P)
wave velocity of the soil.
D a t a from the acoustic program were processed using conventional
common depth point procedures for multichannel seismic analysis [38]. An
example of a brute stack from the Beam 3 profile (Figure 4a) is given in
Figure 5. Two principal acoustic horizons are easily observed; the seabed
event, a positive (black) peak (trace excursions to the right), at about 2.0
ms and a second event, a trough at about 5.0 ms, b o t h indicated by t h e
arrow heads on either side of the plot. The seismic section is displayed with
two-way vertical travel times versus horizontal position in the profile. There
are many other weaker events evident but below about 6 ms the events
become more disjointed and incoherent. Figure 6 shows (a) a b r u t e stack for
the Beam 3 profile, (b) filtered and gained normal incidence trace (repeated
four times as a visual aid), (c) stratigraphie indicators from Borehole 1 and
(d) the geotechnical field report of the samples showing stratigraphy. T h e
reflection even at 5.0 ms in Figure 6a correlates with a boundary t h a t was
not inferred from the sampling program possibly due to low sample recovery
rates (see Figure 6d). A boundary was evident in the geotechnical analyses
and its existence corroborated from the force needed to push t h e cone
penetrometer past this boundary. It is indicated by the triangle symbol
(Figure 6c).
5 . 2 . W a v e l e t t r a n s f o r m analysis of a n a c o u s t i c signal
Using a Morlet analyzing wavelet, the wavelet transform was applied to
the normal incidence d a t a from the field program. These d a t a are recorded
at t h e locations in Figure 4 where t h e sparker source and hydrophones are
coincident. Figure 7a shows the normal incidence trace from profile Beam 3,
station 5 with a geometrical spreading correction applied to the d a t a . T h e
geometrical spreading correction was estimated using the average velocity
198 C. Pike
F i g u r e 6. (a) Brute stack for profile Beam 3; (b) filtered normal in-
cidence trace for station 5, Beam 3 (trace is replicated four times for
visual aid); (c) borehole 1 showing identified stratigraphie boundaries;
(d) soil profile description based on physical samples and cores.
the 0.005 sec event are several more pronounced excursions from t h e signal
and these could be due to scattering from the fractured bedrock or internal
multiple reflections.
T h e modulus of t h e wavelet transform for the acoustic signal (Fig-
ure 7b) exhibits bright areas (shown as white) where the signal strength is
strong (note the direct arrival and the seabed event in Figure 7a). There
is more detail represented in this display t h a n would be found in a con-
ventional power spectra display. T h e modulus of the wavelet transform
associated with t h e direct pulse exhibits three lighter areas (see Figure 7b,
between 7.5 and 10 ln(Hz) and 0.00025 and 0.001 sec). Comparison be-
tween the waveform of the direct pulse (Figure 7a) and t h e modulus of t h e
200 C. Pike
0 0.001 0.002 0.003 0.004 0.005 0.006 0.007 0.008 0.009 0.01
Time (sec)
5 . 3 . E s t i m a t i o n of a t t e n u a t i o n
For the attenuation analysis a similar procedure was followed as de-
scribed in section 3.2. Figure 10 shows a comparison between the variance
of t h e spectral ratio (black dots) and the value for the variance of the cor-
rected d a t a using the estimated spherical spreading correction (solid line).
T h e points cluster fairly closely about the solid line indicating t h a t either
the spectral ratio or the corrected d a t a could be taken for analysis. Atten-
uation estimates were derived from the modulus of the wavelet transform
of the d a t a corrected for geometrical spreading losses. This approach was
High Resolution Marine Seismic Data Analysis 203
Table I
Wavelet Transform Attenuation Values
Beam 1 (nepers/m) Beam 3 (nepers/m) |
1 Station 5 4.4 ÉÛ f
| Station 6 3.5 2.9
| Station 7 3.5 2.8
| Station 8 5.6 6^ [
§6. C o n c l u s i o n
Two analyses based upon the wavelet transform have been presented.
The qualitative analysis discusses the features observed in the modulus
of the wavelet transform in relation to the acoustic signal and the sub-
seabed. It represents a starting point from which feature extraction and a
quantitative analysis can begin with the goal of developing classifiers for
the scatterers. It should be noted t h a t only the Morlet analyzing wavelet
has been utilized in this study. Examination of alternate wavelets t h a t may
be more diagnostic for classification of scattering should be pursued. T h e
next stage of the study of wavelets for this application will employ other
High Resolution Marine Seismic Data Analysis 207
0 0|
-1 -1
1 - 4 ■■!■■<■■: H
Φ Q
Φ . p Φ Φ
Q :^» : : o
Q
-δ[· -·:·>Ν·- -| ;· j -5
-5 -5
-6l· Λ i- | -6
-6 -6
-7
-7 -7
7 8 9 7 8 9 7 8 9
(nepers) (nepers) (nepers) (nepers)
References
1. Buchan, S., D. M. McCann, and D. Taylor Smith, Relations between
the acoustic and geotechnical properties of marine sediments, Q. Jl.
Eng. GeoL, 5, 265-284, 1972.
2. Courtney, R. C. and L. A. Mayer, Calculation of acoustic parameters
by a filter- correlation method, J. Acoust. Soc. Am., 93, 1145-1154,
1993.
3. Dodds, D. J., Attenuation estimates from high resolution s u b b o t t o m
profiler echoes, Presented at the Saclant Asw Research Conference on
Ocean Acoustics Influenced by the Sea Floor, La Spezia, Italy, J u n e 9 -
12, NATO conference on b o t t o m interacting ocean acoustics, 173-191,
1980.
4. English, J. M., S. T. Inkpen and J. Y. Guigné, A new, high-frequency,
broadband seismic source. Presented at the 23rd Annual Offshore Tech-
nology Conference in Houston, Texas, O T C 6558, 1991.
5. Goupillaud, Pierre L., Three New Mathematical Developments on the
Geophysical Horizon, Geophysics: The Leading Edge of Exploration,
40-42, J u n e 1992.
6. Goupillaud, P., A. Grossmann and J. Morlet, Cycle-Octaves and Re-
lated Transforms in Seismic Signal Analysis, Geoexploration, 23, 8 5 -
102, 1984.
7. Goupillaud, P., A. Grossmann, J. Morlet, Cycle-Octave Representation
for Instantaneous Frequency Spectra, SEG Expanded Abstracts with
Biographies, 53rd Annual International SEG Meeting, 613-615, 1983.
8. Grant, F. S. and G. F. West, Interpretation Theory in Applied Geo-
physics, McGraw-Hill Book Company, New York, 1965.
9. Hamilton, E. L., Geoacoustic modelling of the sea floor, Journal of the
Acoustical Society of America, 68, 1313-1340, 1980.
High Resolution Marine Seismic Data Analysis 209
Chris J. Pike
Centre for Cold Ocean Resources Engineering (C-CORE)
Memorial University of Newfoundland
St. John's, Newfoundland
Canada, A1B 3X5
e-mail: [email protected]
Including Multi-Scale Information in the Characterization of
Hydraulic Conductivity Distributions
§1. Introduction
Various authors ([14] and [19]) have shown that the spatial variability
of hydraulic conductivity, K, produces most of the transport dispersion ob-
served at the field scale. If we could accurately reproduce the K-distribution
at a fine scale in a numerical model, there would be no need for stochas-
tic theories. However, it is impractical to collect fine scale K values for
each grid block in a high resolution model. Moreover, even if we had the
information, the model would likely contain enough grid blocks (and there-
fore enough unknowns) that it would take excessive computation time to
struct a fine scale grid. For this initial work, we have assumed t h a t there
is no noise in our sample data; t h a t is, our sample values perfectly rep-
resent the unknown information for the given sample location and scale.
Inclusion of measurement uncertainty is delegated to future research, and
might possibly follow along the lines of [2]. The wavelet reconstruction
technique is compared with other traditional interpolation techniques by
using a 256 x 256 node synthetically generated two-dimensional porous
media and solving the flow and transport equations.
Additionally, we have used a modification of the wavelet reconstruc-
tion technique, dubbed wavelet extrapolation, to investigate what, if any,
important scales of information exist in various synthetic representations
of porous media K distributions. This extrapolation m e t h o d creates es-
timates at each location for any scale finer t h a n the given, completely
sampled, scale. Here, we a t t e m p t to understand the relative importance of
each scale to identify the inherently significant coarse scale representations
of the fine scale variability.
§2. W a v e l e t B a s e d I n t e r p o l a t i o n
2.1. Discrete transform introduction
To understand our wavelet based reconstruction and extrapolation
techniques, we must first familiarize ourselves with some mathematical
details of wavelet transforms. The wavelet transformation is similar to
the Fourier transform in t h a t a d a t a set is transformed into the frequency
(or scale) domain. T h e Fourier transform (via spectral analysis) has been
used extensively to study the characterization of porous media [7], b u t a
limitation of the Fourier analysis technique is the assumption t h a t the in-
formation at each scale is homogeneous and can be essentially represented
by a single "power" value. W h a t makes the wavelet transform useful as
a tool for multi-scale reconstruction of porous media is t h a t this assump-
tion is not necessary, resulting in the preservation of positional information
(i.e. t h e heterogeneity) at each scale. Daubechies in [6] best expresses t h e
wavelet transform as "a tool t h a t cuts up d a t a or functions or operators
into different frequency components, and then studies each component with
a resolution matched to its scale." Thus, with the wavelet transform, one
can analyze d a t a sets with heterogeneous information at each scale, which
is commonly the case for porous media distributions.
T h e forward discrete wavelet transform consists of a convolution of a
basis function (sometimes confusingly called the wavelet function) with a
d a t a set. Unlike the Fourier transform, which uses only sines and cosines
as basis functions, the wavelet transform can use a variety of basis func-
tions. Basis functions are discussed in detail in [6], among others. Here we
discuss the important basis function characteristics t h a t will concern our
Characterization of Hydraulic Conductivity Distributions 217
1 (a)
π
U
F i g u r e 2. Graphs of the (a) (left) Haar basis function, (b) (middle)
Daubechies4, and (c) (right) Daubechiesl2 basis functions.
Table I
Basis function coefficients for the Haar and Daubechies4 functions.
Haar
ci 0.70710678
c2 0.70710678
Daubechies4
d 0.48296291
c2 0.83651630
c3 0.22414387
c4 -0.12940952
fine scale variability (the detail coefficients) and the coarser scale smooth-
ness (the smooth coefficients) according to:
{sm} = [H]{sm+1}; {dm} = [G]{sm+l} (1)
where s represent smooth coefficients, d represent detail coefficients, m is
the level, and H and G are the convolution matrices based on the wavelet
basis function. Higher values of m signify finer scales of information. The
complete wavelet transform is a process that recursively applies Equa-
tion (1) from the finest to the coarsest level (scale). This describes a scale
by scale extraction of the variability information at each scale. The smooth
coefficients generated at each scale are used for the extraction in the next
coarser scale.
The matrices H and G are created from the coefficients of the basis
functions, and represent the convolution of the basis function with the data.
The following is the general makeup for the H and G matrices for a four
coefficient basis function that has a scale-to-scale relationship of 2 (basis
functions that have more or less coefficients would be similar):
C\ C2 C3 C4 0 0 0 0 0 0
0 0 Ci C2 C3 C4 0 0 0 0
H = 0 0 0 0 ci c2 c3 c4 0 0 (2)
c3 c4 0 0 0 0 0 0 c\ c2
c4 -c3 c2 -ci 0 0 0 0
0 0 c4 —c3 c2 — c\ 0 0
G = 0 0 0 0 c4 —c3 c2 —ci (3)
C2 -d 0 0 0 0 0 0
c4 - c 3
As a forward wavelet transform example, Figure 3a shows a simple 4
element one-dimensional data set {6,4,5,3} and the first convolution appli-
Characterization of Hydraulic Conductivity Distributions 219
[ 6
2^- 1 1
V2 "V2 4
2 — 0 0 1 1
5 (5)
3
for the detail coefficient calculation. Continuing, we can represent the
final calculation in Figure 3b in matrix notation for the smooth coefficient
upscaling, and the detail coefficient calculation as, respectively:
H* ^} (6)
Traditionally, only the detail coefficients at every scale, and the coarsest
level smooth coefficients, are considered a complete set of wavelet coeffi-
cients. In our example of Figure 3, the complete set is {9, 1, 1.414, 1.414
}. In our work however, we are not so much interested with just t h e tra-
ditional wavelet coefficients, but in the entire suite of smooth and detail
coefficients at each scale (or level) since we will show in t h e next section
t h a t our multi-scale samples can be related to smooth coefficients at various
scales and locations.
T h e inverse discrete wavelet transform is similarly implemented via
a recursive recombination of the smooth and detail information from t h e
coarsest to finest level (scale):
λ
■1/V5
m = 3{
.-{
.-3{
m-itH-%} (9)
Characterization of Hydraulic Conductivity Distributions 221
9.00 1.00
s, d,
/ 1/V2 \ 1
n-2{
7.07 5.66
si S2
5.66 1.41
„ = 2{ s2 d2
1/V2 / / 1/V2 \ i
„.,{
41r
r 61
4 ±
4r 0
0
0
0
5 0 1 0 1 2 — • (io)
3 0 4r J 0 —fe
V2 s/2
Level
(m)
§3. Traditional I n t e r p o l a t i o n M e t h o d s
T h e wavelet reconstruction technique was compared to three tradi-
tional interpolation methods: simple assignment, kriging, and conditional
simulation. These methods are briefly reviewed below. Kriging and con-
ditional simulation are classical geostatistical methods t h a t have been in
use in mining engineering and geological fields for over a decade. Journel
226 K. Brewer and S. Wheatcraft
Table II
Pseudo-code of the wavelet multi-scale reconstruction algorithm.
C SETUP
READ D — FRACTAL DIMENSION
BETA = 2*(3-D)+l
C READ IN SAMPLE INFORMATION
FOR ALL SAMPLES
READ SAMPLE => LOCATION, VALUE, LEVEL
SMOOTH_COEF(LOCATION, LEVEL) = ADJUST_VALUE(SAMPLE VALUE)
KNOWN_FLAG(LOCATION, LEVEL) = 1
END FOR
C ASSIGN DETAIL COEFFICIENTS
FOR EVERY LEVEL
FOR EVERY LOCATION
DETAIL_COEF(LOCATION, LEVEL) = GAUSS*2**(-0.5*(LEVEL-1)*BETA)
END FOR
END FOR
C PERFORM INITIAL DOWN SWEEP
FOR LEVEL = MINLEVEL+1 TO MAXLEVEL-1
FOR EVERY LOCATION
IF KNOWN_FLAG(LOCATION, LEVEL) IS NOT 1 THEN
XWT = WAVELET_TRANSFORM(LEVEL, LOCATION, FORWARD_FLAG)
XIWT = WAVELET_TRANSFORM(LEVEL, LOCATION, INVERSE_FLAG)
SMOOTH_COEF(LOCATION, LEVEL) = AVERAGE(XWT, XIWT)
END IF
END FOR
END FOR
C CALCULATE THE FINEST SCALE (NO FORWARD TRANSFORM POSSIBLE)
LEVEL = MAXLEVEL
FOR EVERY LOCATION
IF KNOWN_FLAG(LOCATION, LEVEL) IS NOT 1 THEN
XIWT = WAVELET_TRANSFORM(LEVEL, LOCATION, INVERSE_FLAG)
SMOOTH_COEF(LOCATION, LEVEL) = XIWT
END IF
END FOR
C GO UP AND DOWN UNTIL CONVERGENCE IS ACHIEVED
CONVERGENCE_FLAG IS FALSE
DO UNTIL CONVERGENCE_FLAG IS TRUE
C GO UP
FOR LEVEL = MAXLEVEL-1 TO MINLEVEL+1 BY -1
FOR EVERY LOCATION
IF KNOWN_FLAG(LOCATION, LEVEL) IS NOT 1 THEN
XWT = WAVELET_TRANSFORM(LEVEL, LOCATION, FORWARD_FLAG)
XIWT = WAVELET_TRANSFORM(LEVEL, LOCATION, INVERSE_FLAG)
SMOOTH_COEF(LOCATION, LEVEL) = AVERAGE(XWT, XIWT)
END IF
END FOR
END FOR
C GO DOWN
FOR LEVEL = MINLEVEL+2 TO MAXLEVEL-1
FOR EVERY LOCATION
IF KNOWN_FLAG(LOCATION, LEVEL) IS NOT 1 THEN
XWT = WAVELET_TRANSFORM(LEVEL, LOCATION, FORWARD_FLAG)
XIWT = WAVELET_TRANSFORM(LEVEL, LOCATION, INVERSE_FLAG)
SMOOTH_COEF(LOCATION, LEVEL) = AVERAGE(XWT, XIWT)
END IF
END FOR
END FOR
C DO FINEST LEVEL
LEVEL = MAXLEVEL
FOR EVERY LOCATION
IF KNOWN_FLAG(LOCATION, LEVEL) IS NOT 1 THEN
XIWT = WAVELET_TRANSFORM(LEVEL, LOCATION, INVERSE_FLAG)
SMOOTH_COEF(LOCATION, LEVEL) = XIWT
UPDATE MEAN_CHANGE AND MAX_CHANGE
END IF
END FOR
IF MEAN_CHANGE < DESIRED_MEAN_CONVERGENCE _AND_
MAX_CHANGE < DESIRED_MAX_CONVERGENCE THEN
CONVERGENCE_FLAG IS TRUE
END IF
END DO
C WRITE THE FINEST LEVEL VALUES
FOR EVERY LOCATION
WRITE SMOOTH_COEF(LOCATION, MAXLEVEL)
END FOR
STOP END
Characterization of Hydraulic Conductivity Distributions 227
Table III
Pseudo-code of the wavelet multi-scale extrapolation algorithm.
C SETUP
READ D — FRACTAL DIMENSION
BETA = 2*(3-D)+l
READ MAXLEVEL
READ OTHER VARIABLES
C READ IN SAMPLE INFORMATION ~ COMPLETE ON ONE SCALE
FOR ALL SAMPLES
READ SAMPLE => LOCATION, VALUE, LEVEL
SMOOTH_COEF(LOCATION, LEVEL) = SAMPLE VALUE
MINLEVEL = LEVEL
END FOR
C ASSIGN DETAIL COEFFICIENTS WHEN SWEEPING DOWN
FOR LEVEL = MINLEVEL+1 TO MAXLEVEL
FOR EVERY LOCATION
DETAIL_COEF(LOCATION, LEVEL) = GAUSS*2**(-0.5*(LEVEL-1)*BETA)
XIWT = WAVELET_TRANSFORM(LEVEL, LOCATION, INVERSE_FLAG)
SMOOTH_COEF(LOCATION, LEVEL) = AVERAGE(XWT, XIWT)
END FOR
END FOR
C WRITE THE FINEST LEVEL VALUES
FOR EVERY LOCATION
WRITE SMOOTH_COEF(LOCATION, MAXLEVEL)
END FOR
STOP END
and Huijbregts [9] and Hohn [8] detail the development and mathematical
justification behind these and other geostatistical methods.
3.2. Kriging
Kriging is a best linear unbiased estimator (BLUE) and uses the spatial
covariance of the samples to determine the best estimate at a particular
point. All samples are assumed to have point support and thus to be from
only one scale. As developed in [9], the kriging system consists of solving
for sample weights, λ:
[W}{\} = {B} (14)
where W is the sample-point to sample-point covariance matrix and B is
the sample-point to target-point covariance vector. The weights are then
used on the sample values to calculate an estimate at the target point. This
procedure is applied at each unknown grid point. The kriging algorithm in
[1] was used with a slight modification to incorporate the simple assignment
technique, described above, for those fine grid locations which were beyond
the algorithm range.
F i g u r e 6. Sampling schematic.
•
2-
T Sample •
• Known •
•
•
1.5- •
•
•
1-
• ▼ ▼
••
0.5- •
T
• ▼
▼ ▼ ▼ ▼ ' ▼
0- Ml 1 1
▼ ▼
1 1
0 —i5 1
10 15r 20 25 30 35 40
Lag
T h e medium and large scale sample values were calculated using a sim-
ple geometric mean of all known K values in each sample area [11]. This
sample set (sample value, location, and scale) was used as input for t h e
simple assignment and wavelet interpolation techniques. For the kriging
and conditional simulation interpolation techniques, the sample set was
modified by locating the medium and large scale samples at each sample
centroid, and discarding the sample scale information. T h e sample semi-
variogram, Figure 7, was used for determining the covariance structure for
the kriging and conditional simulation reconstruction techniques. A range
of 12, sill of 0.5, and a nugget of 0.0 was determined from the sample
semivariogram. The sample semivariogram is not similar to the known
semivariogram, and this difference is a likely cause for part of any resulting
error for the kriging and conditional simulation techniques.
Only one fine scale reconstruction realization from the sample set was
required for the deterministic simple assignment and kriging techniques.
The wavelet interpolation and the conditional simulation techniques, how-
ever, used a Monte Carlo ensemble average. The Daubechies4 wavelet basis
function [6] was used for the wavelet reconstruction technique.
For each of t h e fine scale reconstructions, as well as for t h e known grid,
the flow equation was solved with a multigrid approach [4] with no-flow side
Characterization of Hydraulic Conductivity Distributions 231
r- 38000
I, A
\\ -37000
J| f\ -36000
-35000
N,\,v- - +- v ~ \^ - - - - - - -34000
~v^/*- -33000
j -32000
-31000
-30000
where r; is the time for the i t h particle to travel through the reconstructed
domain, k{ is the time for the ith particle to travel through the known
domain, and for each of the n particles. These resulting mean error values
are shown in Table IV. As the percentages show, the wavelet and condi-
tional simulation techniques recreate breakthrough curves b e t t e r t h a n t h e
simple and kriging techniques. For this particular sample set, it appears
t h a t conditional simulation technique is slightly b e t t e r t h a n wavelet recon-
struction. This result may be an artifact of the sample suite or the basis
function used in the wavelet method.
4 . 2 . O p t i m u m scale
Addressing the question of whether there are optimum sample scales
for various synthetic two-dimensional hydraulic conductivity distributions,
the wavelet extrapolation technique was used with complete sample cover-
age at various scales. An optimum sample scale is defined as one where t h e
greatest reduction in breakthrough curve error occurs, which is indicative
of important scales representing the coarsening process of the fine scale
variability. Identification of optimum sample scales are also important in
developing an effective and efficient site characterization procedure. T h e
experiments consisted of using the wavelet extrapolation technique t o ex-
trapolate complete sample information at each scale down to t h e original
finest scale. T h e Daubechies4 wavelet basis function was again used for all
Characterization of Hydraulic Conductivity Distributions 233
Figure 9. Fine scale ln(K) shaded contour maps for: (a) The known
grid.
234 K. Brewer and S. Wheatcraft
/
0.9-
0.8- /J
/ ,· r
/ / / ' /
0.7-
0.6-
S i' 1 / /
1
0.5- I ' /, /
! ' iM
0.4-
Known
0.3-
i ii 1
i 1/ ) ——— Simple
0.2- Kriged
■ · s- y
Wavelet
0.1- 1 ' lx
1 · l: ..... Cond. Sim.
n 1 / 1J: 1 1 1 1 1 1 1 1
0 100 200 300 400 500 600 700 800 900
Time
Figure 10. Particle breakthrough curves.
Table I V
Mean breakthrough curve error for reconstruction
techniques.
Reconstruction Technique Mean BTC error
Kriging 48%
extrapolations.
W i t h a 128 x 128 regular grid, 6 separate levels can be completely sam-
pled. Samples at each level were generated using the geometric mean of the
known fine scale values. Each complete level sample set is then used, sepa-
rately, to reconstruct the fine grid via the wavelet extrapolation technique.
T h e flow equations are solved and particle breakthrough curves are gener-
ated. Since the extrapolation technique is stochastic, a Monte Carlo suite
of thirty reconstruction realizations was necessary to ensure stability of t h e
first and second order moments of the ensemble breakthrough curve. These
techniques are similar to those described in Subsection 4.1. An example
of a breakthrough curve suite for all levels is given in Figure 11. This fig-
Characterization of Hydraulic Conductivity Distributions 239
i 1 1 1 1 1 1 1 1
0 100 200 300 400 500 600 700 800
Time
50
45
p 40
LU
Φ 35-
30
5 -
O) 2 5 -
I i1
2 20- 11
CO
15-
Φ 1
ω 10-
c
co
Φ
-i 1 r-ü
10 100 1000 10000 100000
Number of Samples
10000 100000
Number of Samples
Figure 13. Mean error change for increasing finer samples levels, 128 X
128 correlated stochastic field with a correlation length of 12 units.
Characterization of Hydraulic Conductivity Distributions 241
1-
0- 1 1 1 — 1
10 100 1000 10000 100000
Number of Samples
F i g u r e 14. Mean error change for (a) 128 X 128:12, (b) 512 X 512:12,
and (c) 512 x 512:48 correlated stochastic fields.
grid of 512 x 512 and the same fractal dimension, no change in the sample
o p t i m u m (64 samples) was evident, as shown in Figure 16. In this case,
however, the sample size is 64 x 64 grid units. T h e cause for the difference
in o p t i m u m sample size, yet consistency in sample quantity is not clear.
4 . 3 . W a v e l e t basis effects
As noted in Subsection 2.1, many different wavelet basis functions can
be used in the wavelet transform, and consequently, also in the wavelet
reconstruction and extrapolation techniques. Different wavelet basis func-
tions will preferentially move, between scales, different characteristics of
the target dataset. For example, use of the Haar basis function will em-
phasize discontinuities in the target dataset, and similarly, would enhance
reconstruction of a "discontinuous" fine grid field from a sample set. Other
wavelet basis functions, such as the Daubechies family used throughout
this research, emphasize the smoothness of the examined data.
T h e wavelet reconstruction experiments for fine scale ln(JiT) grids proved
to be highly dependent on the wavelet basis chosen. Figure 17 shows a
Haar basis wavelet reconstruction realization ln(Ä') shaded contour m a p .
T h e resulting discontinuous field is evident, especially compared with t h e
Daubechies4 reconstruction in Figure 9e. These discontinuities in the Haar
reconstruction, which are not inherent in the known grid, m a d e it unsuit-
able as a basis function for reconstructing our known fractal distributions.
An example of the effect on breakthrough curve error by utilizing dif-
ferent wavelet basis functions is shown in Figure 18. T h e Haar wavelet
basis function has greater breakthrough curve error t h a n the Daubechies4
wavelet function at all sample scales. The error decrease between scales
is also generally slower for the Haar basis. This indicates t h a t t h e H a a r
wavelet basis function does not interpolate the finer scale variability as
well as the Daubechies family. This effect is again primarily due t o the
difference in t h e smoothing properties between the Haar and Daubechies
wavelet basis functions. Since our known field (a pseudo-fractal) is rela-
tively smooth, the Daubechies4 wavelet function performs better. Similar
results should be expected for correlated stochastic grids.
§5. C o n c l u s i o n s a n d R e c o m m e n d a t i o n s
A primary purpose of this initial work was to investigate the usefulness
of t h e wavelet technique for reconstructing fine scale variability and mov-
ing multi-scale information. A secondary purpose was to use the wavelet
extrapolation technique to begin to understand what, if any, optimal scales
may exist. Based on the results discussed in Section 4, a number of con-
clusions can be stated:
244 K. Brewer and S. Wheatcraft
Figure 17. Shaded contour map of fine scale ln(JsT) field reconstructed
using Haar basis function wavelet.
Characterization of Hydraulic Conductivity Distributions 245
bu-.
^
o^ 45-
l_
2 40-
LU
CD 35-
£
3
ü 30-
25- b
13
2 20- H
sz
V
CO 15-
ω
m 10-
c
CO
φ h-
:> 0 1 1 1 r-ü 1
10 100 1000 10000 100000
Number of Samples
§6, C o n c l u d i n g R e m a r k s
Since fractals have correlation at all scales, the consistency of the opti-
m u m sample quantity (not scale) for the fractal field experiments is some-
what puzzling. One possible explanation might be t h a t there is a similarity
in the way wavelet techniques function compared to statistical relationships
of the number of required samples necessary for adequate representation
of a normal distribution. This may be an indication t h a t fractal distribu-
tions are best modeled by the wavelet method. Additional experiments are
needed.
Besides the areas of future research described above, part of a future
research plan should include increasing the robustness of the wavelet tech-
nique to handle irregular sample locations and overlapping sample volumes.
In addition, applying the reconstruction technique to a tightly controlled
field-generated multi-scale d a t a set is needed. Comparison of the recon-
struction of a fine scale grid, and various contaminant plumes should fur-
ther clarify t h e limitations and expectations of the wavelet reconstruction
application.
References
1. Carr, J. R., UVKRIG: A FORTRAN-77 program for universal Kriging,
Comp. and Geosci., 16, 211-236, 1990.
2. Chou, Kenneth C , A stochastic modeling approach to multiscale signal
processing, P h . D . Thesis, Massachusetts Institute of Technology, May
1991.
3. Chou, K. C , A. S. Willsky, A. Benveniste, and M. Basseville, Recur-
sive and iterative estimation algorithms for multi-resolution stochastic
processes, Proceedings of the 28th Conference of Decision and Control,
Tampa, Florida, IEEE, December 1989.
Characterization of Hydraulic Conductivity Distributions 247
Kevin E. Brewer
Department of Geological Sciences/ 172
University of Nevada, Reno
Reno, Nevada 89557
USA
e-mail: kevinbUhydro.unr.edu
Stephen W. Wheatcraft
Department of Geological Sciences/ 172
University of Nevada, Reno
Reno, Nevada 89557
USA
e-mail: steveuhydro.unr.edu
Wavelet-Based Multifractal Analysis of Non-Stationary and/or
Intermittent Geophysical Signals
§2. P r e l i m i n a r y C o n s i d e r a t i o n s o n S t a t i o n a r i t y , E r g o d i c i t y a n d
Scale-Invariance
2 . 1 . D a t a specification a n d s p e c t r a l analysis
We consider given a discrete set of N + 1 d a t a points:
fi = f(xi), Xi = U (i = 0 , l , . . . , J V ) , (1)
sampled from / ( # ) , x G [0, L], some scalar geophysical field (x is position) or
time-series (x is time). T h e overall length of the one-dimensional dataset is
L and / is t h e grid constant (or 1//, t h e sampling r a t e ) . We will furthermore
assume t h a t
i V = y » l . (2)
252 A. Davis et al.
dropping subscripts for simplicity. This is probably the most popular statis-
tic in time-series analysis, beyond 1-point p.d.f.'s (i.e., simple histograms
of /-values as in Figure l a ) . T h e energy spectrum E£(k) of the ε-field in
Figure l b is computed in the same way and b o t h spectra for the d a t a in
Figure 1 are plotted in Figure 2.
In Equation (3) we use (·) to denote an "ensemble" average which in-
volves in principle every possible realization of the random process f(x).
In geophysical practice however, we are generally provided with the out-
come of a small number of experiments - possibly even a single one - and
we are forced to sacrifice spatial information to obtain spatial averages,
as estimates of their ensemble counterparts. Processes for which spatial
averages converge to the corresponding ensemble statistics in the limit of
large averaging sets are called "ergodic." Estimating an energy spectrum
with a single realization, dropping the triangular brackets in Equation (3),
amounts to making an ergodicity assumption. Throughout this paper, we
Multifractal Analysis of Geophysical Signals 253
use the same notation for ensemble averages and their spatially obtained
estimates. The interested reader is referred to [32] and [29] for a detailed
analytical study of multiplicative cascade models where cases of ergodic-
ity violations can be found; this important class of models will be briefly
discussed in section 4 and in the Appendix.
410
time (sec)
I i I i I i I i I i I i I i I i I i I i I i I i I i I . I i
lo
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 92(kL)
§3. Q u a n t i f y i n g a n d Qualifying N o n - S t a t i o n a r i t y w i t h S t r u c t u r e
Functions
In this section we interest ourselves primarily in non-stationary pro-
cesses with stationary increments (1 < ß < 3) which are ubiquitous in na-
ture. So, following t h e ideas surveyed in t h e Appendix, we will focus on t h e
field of increments of a random process f(x) over a distance r: Δ / ( Γ ; X) =
f(x + r / 2 ) - f(x - r / 2 ) , r / 2 < x < L - r / 2 (0 < r < L). T h e statistical
moments of | Δ / ( Γ ; Χ ) | are a two-point statistic known as "structure func-
tions of order p " and they are independent of x: ( | Δ / ( Γ ; X)\P) = ( | Δ / ( Γ ) | Ρ ) ;
this incremental stationarity is a necessary condition to feel justified in es-
timating ( | Δ / ( Γ ) | Ρ ) by averaging | Δ / ( Γ ; Χ ) | Ρ over x. In most cases small
increments are a frequent occurrence so one cannot take arbitrarily large
negative values for p.
Arnéodo, Bacry and Muzy have shown in a series of papers ([5], [7], [9]
and [45]) how wavelets can be explicitly used to perform structure function
analysis in its standard and more general forms. We summarize their ideas
in the upcoming subsection then review the implications of scale-invariance;
finally, we discuss turbulence studies and apply the technique to the LWC
d a t a presented above, underscoring the dynamical meaning of structure
function exponents in connection with non-stationarity.
τ ( λ / 1 \x\ < r / 2 fa Λ
= (6a)
W ( 0 |*| > r/2 '
Multifractal Analysis of Geophysical Signals 257
where boundary complications are dealt with by making f(x) periodic with
period L. Since the kernel dlr of the integral operator in Equation (7) is
parameterized by a variable scale r, we are reminded of a wavelet transform.
Indeed, it is easy to show t h a t
Φσ(α) = - V i / 2 0 * 0 = 4 ^ i / 2 ( z ) , (9a)
on the one hand, and
W
*«M = { r W>; <9b)
dl,(x)
ψ Η (χ) >
ii
λ
\
Γ
ψ β (χ)
in Equation (11). These authors also show how wavelets with more os-
cillations (e.g., successive derivatives of φσ(χ), orthogonal to higher order
polynomials) can be used to find the scale-invariant properties of a struc-
t u r e function-like statistic in Equation (11) for p < — 1, which is dominated
by regular points where t h e signal f(x) is differentiate.
3.2. Scale-invariant s t r u c t u r e f u n c t i o n s , g e n e r a l i t i e s
R a n d o m scale invariant fields/signals f(x) are characterized by scale-
conditioned statistics t h a t follow power-laws with respect to t h e scale pa-
rameter. Equation (11) describes a way of estimating such a statistic (r is
held constant during the spatial averaging), we thus anticipate
<|Δ/(Γ)|*>«Γ«*>, (12)
over some large range of scales [77, R] which may partially or entirely overlap
with the instrumentally accessible range [/,L].
Some general statements can be made about t h e family of exponents
ζ(ρ). Firstly, one exponent is known a priori, from definitions: ζ(0) = 0.
Secondly, ζ(ρ) will be a smooth differentiate function of p no m a t t e r how
rough the d a t a / ( # ) , being essentially a sum of many exponential functions
of p . Thirdly, it can be shown t h a t , if the signal f{x) has absolute bounds,
then ζ(ρ) is monotonically non-decreasing ([23] and [39]). Finally, it can
be shown on more general grounds t h a t ζ(ρ) is concave (ζ"(ρ) < 0).
To see this, choose units of length where L = 1 and / - u n i t s where
(\Af(l;x)\p) « 0 ( 1 ) ; we then have (\Af(r;x)\P) « r<&\ making explicit
the weak dependence of the prefactors on p. It follows t h a t — £ ( p ) l n r is
the second characteristic (or cumulant generating) function ln(exp(— ρξΓ))
of the random variables £ r = — In | Δ / ( Γ ; Χ)\ and is therefore convex [22].
Stationary processes of course have stationary increments b u t these
will scale trivially due to the invariance under translation: ζ(ρ) = 0 (scale-
independent increments). However, due to the effects of finite spatial res-
olution, even theoretical models lead to numerically small C(p)' s ( s e e [39]
for an example). At t h e other end of the non-stationarity scale, we have
continuous functions with bounded non-vanishing first derivatives, yielding
| Δ / ( Γ ; X)\ OC r (for almost every x) hence ζ(ρ) = p.
Any concave ζ(ρ) with ζ(0) = 0 allows the definition of a hierarchy of
exponents:
ζ
Hip) = -ψ- (13)
By "hierarchy" we mean a monotonie function, in this case non-increasing.
T h e reader is referred to Parisi and Frisch's [48] original paper on the mul-
tifractality of turbulent velocity signals/fields for t h e interpretation of the
ζ(ρ) and H (p) functions in terms of variable orders of singularity. ( "Sin-
gularity" is used here in the sense of Hölder-Lipshitz heuristics: how does
260 A. Davis et al.
Ηι=0
Η,-1/2
Hi=2/3
(a)
10 12 14
η=20 m R-5km |og (r/ |)
A Ψ 2
= -5 _O-O-Ö- & --° ° ° °
O-
-Θ -o
,Q--O D D Q
V -10 τ-·β-
.D-B"0-43
^ . ^ ^ O o O
cr"
^ ^ · _ ^ * - X"* X χ X
-2 -is ,-* -X-
>^*"
— Θ- p=i
-20 - -B- p=2
• - O- p=3
X-
p=4
(b)
Αζ(ρ)
0.6
0.5
β-1=ζ(2) / ; . ·
0.4
0.3 Η=ζ(1)
\ /
0.2
Ύ
Ύ
0.1
Ύ
Ρ
0 0.5 1 1.5 2Î 2.5 3 3.5 4
velocity. Arnéodo et al. [5] and Muzy et al. [44] use recursively gener-
ated fields (akin to cascades but with negative weights as well as positive
ones) smoothed by fractional integration. Finally, Benzi et al. [11] have
recently proposed a wavelet-based technique for generating fields with any
prescribed ζ(ρ) function.
We now turn to an alternative (and currently more popular) multifrac-
tal approach to statistical data characterization, singularity analysis, and
another challenging issue in nonlinear processes, namely intermittency.
^0 L
' J
BL(r)
first introduced by Grassberger [28] and Hentschel and Procaccia [31] with
dynamical systems and strange attractors in mind. Following Parisi and
Frisch [48] who where focusing explicitly on structure functions, Halsey et
al. [30] established an interpretation for D(q) in terms of variable orders
of singularity. (In this case, "singularity" is used in the sense of measures:
how does a(x) in p(r;x) = rs(r;x) oc r _ a ( x ) vary with χΊ)
Obtaining K(q) and C(q) or D(q) is the goal of this approach, providing
us with a means to characterize intermittency, in both quantitative and
qualitative terms (examples to follow).
4.3. Geometrical, spectral and dynamical implications of
statistical singularity analysis
Measures with a variable C(q) are known as "multifractals" (one should
specify stationary). This leaves measures with non-vanishing constant C(q)
- hence K{q) oc (q—l) - to be "monofractals" (again stationarity being
implicit). In essence, a random (deterministic) fractal is a sparse subset of
space that can be described statistically (exactly) with a single exponent,
its fractal dimension. Whether viewed as a stationary random measure
Multifractal Analysis of Geophysical Signals 271
max/mean=1
c^o.o
C^O.05 max/mean=10
tA^wÄ
^=0.18 max/mean=52
C^O.72 max/mean=265
η=20 m
(a)
12
10 \ o q-1
X - -Ξ-
- - O-
q=2
8 q=3
N< -X- q=4
ω
V
CsJ 6
ö) I
o.
O
4 I X:
ώ.
o- .o
2 I S
- El· .
I <\Jog 2 (r/l)
0 -Θ—Θ—Θ—©—Θ—Θ—Θ—Θ—e-
8 10 12 14
K(q) (b)
1
0.8
0.6
0.4
1 -β = K(2) .
0.2
^ - Κ Ό ) _
0>
5 . 2 . O n Ç(p)-to-/f(çr) c o n n e c t i o n s
Recalling t h a t e{x) is derived from f(x) for a given geophysical process
(e.g., governing cloud structure), b o t h / ( # ) ' s structure functions a n d s(#)'s
singularity analysis can produce evidence of multifractality. It is therefore
natural t o ask "Is it fundamentally t h e same multifractality?" and, if so,
"Is there a connection between ζ(ρ) a n d K(q)T\ T h e answer t o t h e former
and more philosophical question is probably "yes." Davis et al. [17] address
t h e latter question from an empirical perspective, suggesting t h a t t h e fields
| Δ / ( Γ ; Χ)\ a n d e(r; x) b e studied jointly. This can of course b e done within
Multifractal Analysis of Geophysical Signals 277
with the appropriate ranges for p and q defined in Equations (11) and (19);
concerning r, one should of course remain in the scaling regime defined in
Equation (4). The new exponent function X(p,q) reverts to known cases
for q = 0 and p = 0 and will be concave (dppX < 0, dgqX < 0) on its
domain of definition.
Davis et al. [17] show how the X(p,q) can be used to test the scaling
hypothesis that there exists two exponents t > 0 and s > 0 such that
ε(Γ;ι)=Μ. (24)
This generalizes Kolmogorov's [34] relation for fully developed turbulence
corresponding to s = 1 and t = 3; more precisely, Kolmogorov argued
Equation (24) not at every x but on average only, hence Ç(£ = 3) = s = l
since we know that (s(r)) is independent of r. If however Equation (24)
applies everywhere, then it follows (by taking p/t-th powers and averaging)
that
C(p) = (f)p-if(f), (25)
as often quoted in the case of turbulence in spite of the fact that Equa-
tion (24) has not been verified beyond Kolmogorov's average at p = t = 3
([55]; and references therein). So it appears that at least two new pa-
rameters are needed to connect £(·) and ΛΓ(·), the essential one being
s = ζ(ί) which allows Equation (24) to relate a stationary field (ε) to a
non-stationary one (/). Indeed, for small enough C\ the values of K(p/t)
in Equation (25) will be small as well (except possibly when p ^> i); in
turbulence theory K(p/3) is called an "intermittency correction" (usually
at p = 2). This makes (s/t)p reflect, to first order, the non-stationarity of
f(x) in Equation (25) since we then get ζ(ρ) « (s/t)p hence Hi « s/t (1/3
for turbulence).
(Hi, Ci) G [0,1] ® [0,1], the upper limit on Ci being imposed only to ex-
clude degenerate gradient fields. Figure 9 represents an (ifi,Ci)-plot that
we have populated with empirical findings from the turbulence literature
(approximate positioning) and some of our LWC studies (precise position-
ing).
We notice on Figure 9 that the models discussed in this paper fall pri-
marily on either axis: additive processes (Section 3.3) have Ci = 0 and
Hi > 0 (they are neither intermittent nor stationary) while multiplicative
cascades (Section 4.3) have Hi = 0 and Ci > 0 (they are both stationary
and intermittent). We have added Mandelbrot's [38] "Devil's staircases"
which are simply integrals of cascade models, obeying Equations (24)-(25)
with s = t = 1 according to Equation (15). They are found at Hi = 1 and
finite Ci (they are almost everywhere differentiable but, having fully devel-
oped cascades as gradient fields, they are intermittent). A more familiar
case is found at Hι = Ci = 1: Heaviside steps, the integrals of Dirac £'s. It
seems that these standard scale-invariant models carefully avoid the locus
of the data, inside the square domain. This clearly demonstrates the need
for a new class of stochastic models that can access the whole multifrac-
tal domain and which are likely to have both additive and multiplicative
ingredients. A list of publications where such multi-affine models are de-
scribed is provided at the end of section 3; we believe this list (of five) to
be exhaustive at the time of publication but there is little doubt that many
more will follow. We also anticipate numerous geophysical applications for
(Hi, Ci )-plots. In particular, we foresee all sorts of data-to-data and data-
to-model intercomparisons for the purposes of validating numerical models
or retrievals from remote sensing for instance ([16] and [18]).
§6. Conclusions
Wavelets have been extensively used to show in graphic detail how
structures occur on all observable scales in many geophysical signals. A
natural environment for modeling such signals is provided by stochastic
processes which have no characteristic scale (sporting, e.g., power-law en-
ergy spectra) and these are best described by multiscaling or "multifractal"
statistics. Since scale-invariance prevails in nature as well, we are urged to
blend wavelet- and multifractal analysis techniques. The major benefit of
this merger for the wavelet community is the fact that multifractal statis-
tics have strong dynamical overtones; they can be used to characterize in
quantitative and qualitative terms both intermittency and non-stationarity.
The benefit for the multifractal community is a somewhat more unified view
of their two main tools, singularity analysis (targeting the intermittency)
and structure functions (targeting the non-stationarity). This connection
is timely because many important questions are still open about the inter-
Multifractal Analysis of Geophysical Signals 279
C1 = 1
• .—. . . « « ^ • • •
' ^degenerate gradient fields y / / '
/ / / / / / / / ' </'
1.0 < /. /. <. £ £ / /....J. <, /X...M
Dirac δ-functions Heaviside step functions
-» more integration -»
\
« °- 3 1\ Turbulence
Rj <- dissipation field ...
CO & velocity field -»
•Ό
C
CO
PURELY ϊ
c MULTIPLICATIVE
Φ MODELS J
ts
1
| 30/06/87 FIRE LWC dataseT
£
c
ω 0.1 PURELY
o ADDITIVE Averages for FIRE-87 and
E MODELS; ASTEX-92 LWC data bases
2-D„
o.o — r 0.2 0.3 0.4 0.5 1.0
0.0 0.1
white noise fractional standard almost everywhere
to Brownian Brownian differentiate
Ί/Ρ noise motion motion functions
<0*fte1) (β=5/3) (ß=2) (ß*3)
less stationarity, more smoothness =§>
defined on [0,L], we must require G(r) = 0 for r > L. There is also an in-
ner scale in most situations of practical interest; for 0 < r < /, we can take
f(x + r) « / ( # ) hence G(r) « G(0). Between these limits (with / <^ £ ) , we
assume Equation (A.3) to apply; more precisely, we take G(r) « G(0)[l/ry.
It follows t h a t
y « (l) m i n [M. (A.4)
So, for all practical purposes, the exponent μ in Equation (A.3) cannot ex-
ceed unity without leading to fields t h a t are uncorrelated from one pixel to
the next. If μ -> 0 (hence R& L), there is no trend towards decorrelation;
we view this as a symptom of non-stationarity and in Section A.5 will intro-
duce a 2-point statistic better adapted to this situation t h a n ( / ( # 4 - r ) / ( # ) ) .
β + μ = 1, (Α.6)
β < 1. (A.7)
K(x)
ff(x)
< R^-
have been investigated primarily in the turbulence literature ([35], [47] and
[37]), as a model for the intermittency of the dissipation field in very high
Reynolds number flows. These processes, known as "multifractals" [48],
are now widely applied in deterministic chaos [30].
Stationarity in the above sense [Equations (A.3)-(A.7)] follows from
[43]
(ε(χ + r)e(x)) ~ r~K^, K(2) = logm<W?> (A.10)
nection between ( Δ / ( Γ ) 2 ) and E(k) [43]. For their scaling versions, Equa-
tions (A. 14) and (5), this yields
β-ζ(2) = 1, (Α.15)
in analogy with Equations (A.6) and (A. 11). Notice that the spectral crite-
rion for non-stationarity in Equation (A. 12) is retrieved from (A14-A15).
We must furthermore require β < 3 in order to have stationary incre-
ments by requiring the gradient field (E\r/(k) = k2E/(k)) to be stationary
(/?-2<l).
A.6. Additive scale-invariant models, non-intermittent and
non-stationary
The classic example of a non-stationary stochastic process is Brownian
motion (often referred to as a Wiener-Lévy process):
A . 7 . S t o c h a s t i c c o n t i n u i t y w i t h s t a t i o n a r i t y or scale-invariance
but not both
If ( Δ / ( Γ ) 2 ) -+ 0 when r -» 0, then / ( · ) is said to be "stochastically"
continuous and this is t h e case for all non-stationary scaling processes obey-
ing Equation (A. 14). Of course processes t h a t are stationary per se have
stationary increments (if scaling prevails, then β < 1 implies β < 3). It is
easy to see t h a t
<A/(r)2) = 2 [ G ( 0 ) - G ( r ) ] . (A.17)
function
GR(r) = GR(0) exp(-r/R). (A.18)
In Figure Ale, we took R = L/8 = 512. For the same process the structure
function is therefore
(Δ/Λ(Γ)2) = 2G ß (0)[l - exp(-r/i?)]. (A.19)
Notice the cross-over from non-stationary and continuous behavior (i.e.,
2
( Δ / Λ ( Γ ) 2 ) OC r) to stationary and discontinuous behavior (i.e., (A/#(r) ) «
constant) as r goes from <C R to ^> R. In essence, we are dealing with
Brownian motion (on scales r <& R) that is forced to come back to the
origin on a regular basis (at intervals of length « R). In the limit R —> 0
(vanishing correlation length) white noise is retrieved. For R —> oo (diverg-
ing correlation length) "unconstrained" Brownian motion is retrieved and
the fact that (//*(#)/#(#+ r)) —> constant > 0 in this case is characteristic
of non-stationary behavior, as anticipated in Section A.2.
LIST OF SYMBOLS
Scale parameters:
• general
r temporal or spatial scale on which a statistic is to be
conditioned
k « 1/r, wavenumber (in absolute value)
• instrumental and/or computational
Multifractal Analysis of Geophysical Signals 291
• for either
£(r;x) = \&f(r;x)\,-e(r;x) or p(r;x)
((£(r;x)z) scales like rA^)
Statistical properties:
E{k) energy (or power) spectrum, spectral density
G(r) = ( / ( r + x)f(x)), auto-correlation function (of a
stationary zero-mean process)
( | Δ / ( Γ ; Χ)\Ρ) p-th moment of stationary increments | Δ / ( Γ ; Χ)|,
structure function of order p
292 A. Davis et al.
Φσ(*) = -Vi/20O
Φ#(#) Haar wavelet, see Equation (9b)
Φ Μ (Χ) Mexican hat, see Equation (17a)
Φ^(χ) French top hat, see Equation (17b)
-oil(a?) = -I[ (x) = S(x - r / 2 ) - i(a? + r / 2 ) ,
"poor man's" wavelet [45]
• wavelet transforms of random processes
ϊ φ [ / ] ( α ; b) for structure functions of / ( · ) : Φ = -dh, Φ#, ^ G ;
a = r and 6 = x
Τφ[ε] (α; 6) for singularity analysis of ε(·) :
0 = h, Ψ\, ΨΗ, ΨΜ, VF\ α = r and b = x
Exponents
Mathematical functions:
References
1. Anseimet, F., Y. Gagne, E. J. Hopfinger, and R. A. Antonia, High-
order Velocity Structure Functions in Turbulent Shear Flows, Journal
of Fluid Mechanics, 140, 63-80, 1984.
2. Argoul, F., A. Arnéodo, G. Grasseau, Y. Gagne, E. J. Hopfinger, and
U. Frisch, Wavelet analysis of turbulence reveals the multifractal n a t u r e
of the Richardson cascade, Nature, 338, 51-53, 1989.
3. Arnéodo, A., G. Grasseau, and M. Holschneider, Wavelet transform of
multifractals, Physical Review Letters, 6 1 , 2281-2284, 1988.
4. Arnéodo, A., G. Grasseau, and M. Holschneider, Wavelet transform
analysis of some dynamical systems, in Wavelets, J. M. Combes, A.
Grossmann and P. Tchamitchian, (eds.), Springer, Berlin, 1989.
5. Arnéodo, A., E. Bacry, and J. F. Muzy, Wavelet analysis of fractal
signals, Direct determination of the singularity spectrum of fully de-
294 A. Davis et al.
Anthony Davis
Universities Space Research Association
NASA-GSFC (Code 913)
Greenbelt
MD 20771
USA
e-mail: davisuclimate.gsfc.nasa.gov
298 A. Davis et al.
Alexander Marshak
Science Systems and Applications, Inc.
5900 Princess Garden Parkway
Lanham
MD 20706
USA
Warren Wiscombe
NASA Goddard Space Flight Center
Climate and Radiation Branch
Greenbelt
MD 20771
USA
Simultaneous Noise Suppression and Signal Compression
Using a Library of Orthonormal Bases
and the Minimum Description Length Criterion
Naoki Saito
§1. Introduction
Wavelet transforms and their relatives such as wavelet packet trans-
forms and local trigonometric transforms are becoming increasingly popular
in many fields of applied sciences. So far their most successful application
area seems to be data compression; see e.g., [14], [6], [35], [30]. Meanwhile,
several researchers claimed that wavelets and these transforms are also use-
ful for reducing noise in (or denoising) signals/images [16], [7], [10], [21].
In this paper, we take advantage of both sides: we propose an algorithm
for simultaneously suppressing random noise in data and compressing the
signal, i.e., we try to "kill two birds with one stone."
§3. P r o b l e m F o r m u l a t i o n
Let us consider a discrete degradation model
d = / + n,
where d, / , n G R and N = 2 n . T h e vector d represents the noisy
observed d a t a and / is the unknown true signal to b e estimated. T h e
vector n is white Gaussian noise (WGN), i.e., n ~ Λ/^Ο, σ2Ι). Let us
assume t h a t σ 2 is unknown.
We now consider an algorithm to estimate / from the noisy observation
d. First, we prepare the library of orthonormal bases mentioned in the
previous section. This library consists of the standard Euclidean basis
of R ^ , the Haar-Walsh bases, various wavelet bases and wavelet packet
best-bases generated by Daubechies' QMFs, their less asymmetric versions
(i.e., coiflets), and local trigonometric best-bases. This collection of bases
is highly adaptable and versatile for representing various transient signals
[7]. For example, if the signal consists of blocky functions such as acoustic
impedance profiles of subsurface structure, the Haar-Walsh bases capture
those discontinuous features b o t h accurately and efficiently. If the signal
consists of piecewise polynomial functions of order p , then the Daubechies
wavelets/wavelet packets with filter length L > 2(p + 1) or the coiflets
with filter length L > 3(p + 1) would be efficient because of the vanishing
moment property. If the signal has a sinusoidal shape or highly oscillating
characteristics, the local trigonometric bases would do the job. Moreover,
computational efficiency of this library is also attractive; the most expensive
expansion in this library, i.e., the local trigonometric expansion, costs about
0(iV[log 2 N]2) as explained in the previous section.
Let us denote this library by C = {#i,#2> · · · , ^ M } 5 where Bm repre-
sents one of t h e orthonormal bases in the library, and M (typically 5 to
20) is the number of bases in this library. If we want, we can add other
orthonormal bases in this library such as the Karhunen-Loève basis [1] or
the prolate spheroidal wave functions [13], [36]. However, normally, the
above-mentioned multiresolution bases are more t h a n enough, considering
their versatility and computational efficiency [7].
Since the bases in the library C compress signals/images very well, we
make a strong assumption here: we suppose the unknown signal / can be
completely represented by k ( < N) elements of a basis # m , i.e.,
/ = W ra a£>, (1)
where W m G R X is an orthogonal matrix whose column vectors are the
basis elements of ß m , and a m G R is the vector of expansion coefficients
of / with only k non-zero coefficients. At this point, we do not know the
actual value of k and the basis Bm. We would like to emphasize t h a t in
Noise Suppression and Signal Compression 305
reality the signal / might not be strictly represented by (1). We regard (1)
as a model at hand rather than a rigid physical model exactly explaining f
and we will try our best under this assumption. (This is often the case if we
want to fit polynomials to some data.) Now the problem of simultaneous
noise suppression and signal compression can be stated as follows: find
the "best" k and m given the library C. In other words, we translate the
estimation problem into a model selection problem where models are the
bases Bm and the number of terms k under the additive WGN assumption.
For the purpose of data compression, we want to have k as small as
possible. At the same time, we want to minimize the distortion between
the estimate and the true signal by choosing the most suitable basis ß m ,
keeping in mind that the larger k normally gives smaller value of error.
How can we satisfy these seemingly conflicting demands?
(Prom now on, we denote the logarithm of base 2 by "log", and the nat-
ural logarithm, i.e., base e by "In".) The Shannon code has the shortest
codelength on the average, and satisfies the so-called Kraft inequality [11]:
sum of the codelength of its integer part [v] and the number of fractional
binary digits of the truncation precision J, i.e.,
L(t,,) = L * ( H ) + l o g ( l / i ) . (4)
Having gone through the above examples, now we can state the MDL
principle more clearly. Let M = {0 m : m = l , 2 , . . . } b e a class or collection
of models at hand. The integer m is simply an index of a model in the list.
Let x be a sequence of observed data. Assume that we do not know the
true model 0 generating the data x. As in [29], [24], given the index m, we
can write the codelength for the whole process as
L(x, 0 m , m) = L(m) + L(0 m \ m) + L(x | 0 m , m). (5)
This equation says that the codelength to rewrite the data is the sum of the
codelengths to describe: (i) the index m, (ii) the model 0 m given ra, and
(iii) the data x using the model 0 m . The MDL criterion suggests picking
the model 0 m * which gives the minimum of the total description length
(5).
The last term of the right-hand side (RHS) of (5) is the length of the
Shannon code of the data assuming the model 0 m is the true model, i.e.,
L(x | 0 m , m ) = - l o g p ( x | 0 m ,ra), (6)
and the maximum likelihood (ML) estimate 0 m minimizes (6) by the defi-
nition:
L(x | 0 m , m ) = - l o g p ( x | 0 m , m ) < - l o g p ( x | 0 m ,ra). (7)
However, we should consider a further truncation of 0 m as shown in Ex-
ample 4.4 above to check that additional savings in the description length
is possible. The finer truncation precision we use, the smaller the term
(7), but the larger the term L(0 m | m) becomes. Suppose that the model
0 m has km real-valued parameters, i.e., 0 m = (0 m > i,... ,0m,fcm)· Rissanen
showed in [27], [29] that the optimized truncation precision (δ*) is of order
1/v/ÏVand
min L(x, 0m,<$, ra, δ)
δ
R e m a r k . Even though the list of models M does not include the true
model, the MDL method achieves the best result among the available mod-
els. See Barron and Cover [4] for detailed information on the error between
the MDL estimate and the true model.
We also would like to note that the MDL principle does not attempt
to find the absolutely minimum description of the data. The MDL always
requires an available collection of models and simply suggests picking the
best model from that collection. In other words, the MDL can be considered
as an "oracle" for model selection [24]. This contrasts with the algorithmic
complexities such as the Kolmogorov complexity which gives the absolutely
minimum description of the data, however, in general, is impossible to
obtain [27].
that the data is contaminated by the additive WGN with known variance
σ 2 , i.e.,
Vi = /(«,·) +e<»
where /(·) is an unknown function to be estimated by the polynomial mod-
els, and e,· ~ JV(0, σ 2 ). To invoke the MDL formalism, we pose this ques-
tion in the information transmission setting. First we prepare an encoder
which computes the ML estimate of the coefficients of the polynomial,
(3o,... , S m ) , of the given degree m from the data. (In the additive WGN
assumption the ML estimate coincides with the least squares estimate.)
This encoder transmits these m coefficients as well as the estimation errors.
We also prepare a decoder which receives the coefficients of the polynomial
and residual errors and reconstruct the data. (We assume that the abscis-
sas {xi}iLi and the noise variance σ2 are known to both the encoder and
the decoder.) Then we ask how many bits of information should be trans-
mitted to reconstruct the data. If we used polynomials of degree N — 1,
we could find a polynomial passing through all N points. In this case, we
could describe the data extremely well. In fact, there is no error between
the observed data and those reconstructed by the decoder. However, we
do not gain anything in terms of data compression/transmission since we
also have to encode the model which requires N coefficients of the polyno-
mial. In some sense, we did not "learn" anything in this case. If we used
the polynomial of degree 0, i.e., a constant, then it would be an extremely
efficient model, but we would need many bits to describe the deviations
from that constant. (Of course, if the underlying data is really a constant,
then the deviation would be 0.)
Let us assume there is no prior preference on the order m. Then we
can easily see that the total codelength (9) in this case becomes
m -
t=l \ j=0 /
The MDL criterion has been successfully used in various fields such
as signal detection [32], image segmentation [19], and cluster analysis [31]
where the optimal number of signals, regions, and clusters, respectively,
should be determined. If one knows a priori the physical model to explain
the observed data, that model should definitely be used, e.g., the complex
310 N. Saito
MDL(d,a^\a2,k,m)
= L(fc,ra) + L ( a ^ } , 5 2 \k,m) + L(d | a ^ , 5 2 , f c , m ) , (11)
Κ,-α^ΙΙ2· (17)
Considering that the vector a™ only contains k nonzero elements, we can
easily conclude that the minimum of (17) is achieved by taking the largest
k coefficients in magnitudes of d m as the ML estimate of ctm\ i.e.,
S « = 0(l)dm = 0(t)(Wlrf), (18)
where Θ^ ^ is a thresholding operation which keeps the k largest elements in
absolute value intact and sets all other elements to zero. Finally, inserting
(18) into (14), we obtain
L(k,m) = [ L ( m ) + log(fc2
- kl + X) if 1
ί ^ k ^ *2' (20)
v v
' ^ +oo otherwise. '
312 N. Saito
As for the second term of (11), which is critical for our algorithm, we
have to encode k expansion coefficients â^J and σ 2 , i.e., (fc+1) real-valued
parameters. However, in this case, by normalizing the whole sequence by
||d||, we can safely assume that the magnitude of each coefficient in a ' '
is strictly less than one; in other words, the integer part of each coefficient
is simply zero. Hence we do not need to encode the integer part as in (9)
if we transmit the real-valued parameter ||d||. Now the description length
of (œm ,σ 2 ) given (fc, m) becomes approximately ^ ^ l o g i V - h L*([S2]) +
L*([||d||]) bits since there are k + 2 real-valued parameters: k nonzero
coefficients, σ 2 , and ||d||. After normalizing by ||d||, we clearly have σ 2 < 1
(see (19)), so that L*([52]) = 1 (see (3)). For each expansion coefficient,
however, we still need to specify the index of the coefficient, i.e., where the
k non-zero elements are in the vector orm . This requires klogN bits. As
a result, we have
curve which typically decreases for the small fc, then starts increasing be-
cause of the penalty term, then finally decreases again at some large k near
from k = N because the residual error becomes very small. Now what we
really want is the value of k achieving the minimum at the beginning of
the fc-axis, and we want to avoid searching for k beyond the maximum oc-
curring for k near N. So, we can safely assume that k\ = 0 and &2 = N/2
in (20) to avoid searching more than necessary. (In fact, setting &2 > N/2
does not make much sense in terms of data compression either.)
We briefly examine below the computational complexity of our algo-
rithm. To obtain (fc*,m*), we proceed as follows:
Step 1: Expand the data d into bases B\,..., BM · Each expansion (includ-
ing the best-basis selection procedure) costs O(N) for wavelets, 0(iV log N)
for wavelet packet best-bases, and 0(N[\ogN]2) for local trigonometric
best-bases.
Step2: Let K(= &2 — &i + 1) denote the length of the search range for k.
For k\ < k < k*i, 1 < m < M, compute the expression in the parenthesis of
the RHS in (23). This costs approximately 0(N + 3MK) multiplications
and MK calls to the log function.
Step 3: Search the minimum entry in this table, which costs MK compar-
314 N. Saito
isons.
Step 4: Reconstruct the signal estimate (24), which costs O(N) for wavelets,
0(ΛΓ log N) for wavelet packet best-bases, and 0(7V[log ΛΓ]2) for local trigono-
metric best-bases.
purpose.
As for the 2D version of the wavelet packet best-basis, the sequen-
tial method may be generalized, but it is not easily interpreted; the ID
best-bases may be different from column to column so that the resultant
coefficients viewing along the row direction may not share the same fre-
quency bands and scales unlike the 2D wavelet bases. This also makes the
reconstruction algorithm complicated. Therefore, we should use the other
tensor-product 2D wavelet approach for the construction of the 2D wavelet
packet best-basis: we recursively decompose not only the "low-low" com-
ponents but also the other three components. This process produces the
"quad-tree" structure of wavelet packet coefficients instead of the "binary-
tree" structure for ID wavelet packets. Finally the 2D wavelet packet
best-basis coefficients are selected using the entropy criterion [33].
The 2D version of the local trigonometric transforms can be con-
structed using the quad-tree structure again: the original image is smoothly
folded and segmented into 4 subimages, 16 subimages, . . . , and in each
subimage the separable DCT/DST is applied, and then the quad-tree struc-
ture of the coefficients is constructed. Finally, the local trigonometric best-
basis is selected using the entropy criterion [33].
For an image of N = Ni x ΛΓ2 pixels, the computational costs are
approximately O(N), 0(N\og4 JV), 0(iV[log4 iV]2) for a 2D wavelet, a 2D
wavelet packet best-basis, a 2D local trigonometric best-basis, respectively.
§7. Examples
In this section, we give several examples to show the usefulness of our
algorithm.
x 10
-1
-2
Rs (25)
~—W^b}—-7il +
b}>>
where r is a compression ratio. The original data precision was bf = 8
(bytes) in this case. Since it is enough to use b{ = 2 (bytes) for indices and
r = 13.3%, we have Rs « 9.40%, i.e., 90.60% of the original data can be
318 N. Saito
KJLi
. l*t ^ψΥ^/^
' ' \ * ^ , Λ^ * « Γ % Α ^ ' l/'vw*vv^^ww^j» \wj i^j
t i l ^i
IfV ;
llfS /^AM/WV/X^J
Ιν ν / ^^-\^^/ντ
k/#^w^
discarded.
§8. D i s c u s s i o n s
Our algorithm is intimately connected to the "denoising" algorithm
of Coifman and Majid [7], [10]. Their algorithm first picks t h e best-basis
from the collection of bases and sorts the best-basis coefficients in order of
decreasing magnitude. Then they use t h e "theoretical compression r a t e " of
the sorted best-basis coefficients { α ι · } ^ : 1 as a key criterion for separating
a signal component from noise. T h e theoretical compression r a t e of a unit
vector u is defined as c(u) = 2H^/N(u), where H(u) is the ^ 2 -entropy of
u w
TA, i.e., H(u) = — Σί=ι l l ° S i -> ^ d N(u) is the length of u. We note
t h a t 0 < c(u) < 1 for any real unit vector u , and c(u) = 0 implies u =
{Sifi0} for some z'o (the best possible compression), and c(u) = 1 implies
u = ( 1 , . . . , I)/ y/N(u) (the worst compression). T h e n t o decide how many
coefficients to keep as a signal component, they compare c{{ai]f_k+l)^ the
320 N. Saito
the inverse wavelet transform. Donoho claimed informally in [15] that the
reason why their method works is the ability of wavelets to compress the
signal energy into a few coefficients. The main differences between our
algorithm and that of Donoho and Johnstone are:
• Their method requires the user to set the coarsest scale parameter
J <n and a good estimate of σ 2 , and the resulting quality depends
on these parameters. On the other hand, our method does not
require any such parameter setting.
§9. C o n c l u s i o n s
We have described an algorithm for simultaneously suppressing the ad-
ditive WGN component and compressing the signal component in a dataset.
One or more of the bases in the library, consisting of wavelets, wavelet pack-
ets, and local trigonometric bases, compress the signal component quite
well, whereas the WGN component cannot be compressed efficiently by any
basis in the library. Based on this observation, we have tried to estimate
the "best" basis and the "best" number of terms to retain for estimating the
signal component in the data using the MDL criterion. Both synthetic and
real field data examples have shown the wide applicability and usefulness
of this algorithm.
322 N. Saito
Acknowledgements
The author would like to thank Prof. R. Coifman and Prof. A. Barron
of Yale University for fruitful discussions.
References
1. Ahmed, N. and K. R. Rao, Orthogonal Transforms for Digital Signal
Processing, Springer-Verlag, New York, 1975.
2. Auscher, P., G. Weiss, and M. V. Wickerhauser, Local sine and cosine
bases of Coifman and Meyer and the construction of smooth wavelets,
in Wavelets: A Tutorial in Theory and Applications, C. K. Chui (ed.),
Academic Press, San Diego, 237-256, 1992.
3. Barron, A. R., Complexity regularization with application to artificial
neural networks, in Proceeding NATO A SI on Nonparametric Func-
tional Estimation, Kluwer, 1991.
4. Barron, A. R. and T. M. Cover, Minimum complexity density estima-
tion, IEEE Trans. Inform. Theory, 37 (4), 1034-1054, 1991.
5. Beylkin, G., R. Coifman, and V. Rokhlin, Fast wavelet transforms and
numerical algorithms I, Comm. Pure Appl. Math., 44, 141-183, 1991.
6. Bradley, J. N. and C. M. Brislawn, Image compression by vector quan-
tization of multiresolution decompositions, Physica D, 60, 245-258,
1992.
7. Coifman, R. R. and F. Majid, Adapted waveform analysis and de-
noising, in Progress in Wavelet Analysis and Applications, Y. Meyer
and S. Roques (eds.), Editions Frontieres, B.P.33, 91192 Gif-sur-Yvette
Cedex, France, 63-76, 1993.
8. Coifman, R. R. and Y. Meyer. Remarques sur l'analyse de fourier à
fenêtre, Comptes Rendus Acad. Sei. Paris, Série I, 312, 259-261, 1991.
9. Coifman, R. R. and M. V. Wickerhauser, Entropy-based algorithms
for best basis selection, IEEE Trans. Inform. Theory, 38 (2), 713-719,
1992.
10. Coifman, R. R. and M. V. Wickerhauser, Wavelets and adapted
waveform analysis, in Wavelets: Mathematics and Applications,
J. Benedetto and M. Frazier (eds.), chapter 10, CRC Press, Boca Ra-
ton, Florida, 1993.
11. Cover, T. M. and J. A. Thomas, Elements of Information Theory,
Wiley Interscience, New York, 1991.
12. Daubechies, L, Orthonormal bases of compactly supported wavelets,
Comm. Pure Appl. Math., 4 1 , 909-996, 1988.
13. Daubechies, L, Ten Lectures on Wavelets, volume 61 of CBMS-NSF
Regional Conference Series in Applied Mathematics, SIAM, Philadel-
phia, 1992.
14. DeVore, R. A., B. Jawerth, and B. J. Lucier, Image compression
Noise Suppression and Signal Compression 323
Naoki Saito
Schlumberger-Doll Research
Old Quarry Road, Ridgefield, C T 06877
and
Department of Mathematics
Yale University
10 Hillhouse Avenue, New Haven, C T 06520
e-mail: saito @ridgefield, sdr.sib. com
L o n g - M e m o r y P r o c e s s e s , t h e A l l a n Variance a n d W a v e l e t s
Abstract. Long term memory has frequently been observed in physical time
series. Statistical theory for long-memory stochastic processes is radically different
from the standard time series analysis, which assumes short term memory. The
Allan variance is a particular measure of variability developed for long-memory
processes. This variance can be interpreted as a Haar wavelet coefficient variance,
suggesting an approach towards assessing the variability of general wavelet classes.
The theory is applied to a 'time' series of vertical ocean shear measurements for
which some drawbacks with the Haar wavelet are observed.
§1. I n t r o d u c t i o n
In a variety of applications, time series analysts have noticed t h a t
the estimated autocovariance sequence for their d a t a tends to decrease
rather slowly, indicating t h a t the series has 'long memory' in the sense t h a t
changes in t h e remote past continue to affect the present value of the se-
ries. Beran (1992) gives a good review of statistical and historical aspects
of long-memory processes. Time series t h a t are well modelled by long-
memory processes have been observed, for example, by Newcomb (1895) in
astronomy, by Gösset (Student, 1927) in chemistry, and by Smith (1938) in
agriculture. In geophysics a famous early example is Hurst's (1951) study of
the minimum annual height of the river Nile. This series has a sample auto-
covariance sequence (acvs) sT t h a t is approximately proportional to | τ | - 0 · 3 ;
i.e., the sequence decays hyperbolically, thus ruling out such s t a n d a r d time
series models as ARMA models t h a t have exponentially decaying autoco-
variance sequences (here sT is an estimate of sT = cov {Xt,Xt+r} when
{Xt} is a stationary process). T h e applicability of long-memory processes
to climate d a t a has been recently discussed by Raftery and Haslett (1989)
and Smith (1992).
,τ = !(|τ+1|2"-2|τ|2" + | τ - ΐ Γ ) , r = ±l,±2,...,
where so = var{X*} (note t h a t the case H = 1/2 corresponds t o t h e iid
case because then sT = 0 for τ φ 0, whereas {Xt} has long-memory if
1/2 < H < 1). Then var {X} = s0N2H~2, where X is t h e sample mean of
N observations (i.e., Σ * = ι Xt/N). Note t h a t , for values of H close t o 1,
the rate of decrease of variability in X is markedly different from t h e 1/N
rate of t h e iid case. Naive application of iid statistics t o t h e sample mean
of a long-memory process can thus be very misleading. For example, if we
fit a fractional Gaussian process to the Nile data, we obtain an estimate
of H = 0.85, implying t h a t the variance of t h e sample mean decreases
like 1/N03 instead of t h e usual 1/N rate. Numerically, obtaining 100
observations from this fractional Gaussian process is equivalent t o obtaining
only 4 observations from an iid process!
In spite of its slow rate of decay, t h e sample average of a long-memory
process is still a surprisingly efficient estimator of t h e mean level of t h e
process (Beran and Künsch, 1985; Percival, 1985). Unfortunately, t h e same
cannot b e said for other standard statistics. In particular, t h e sample
variance is a poor estimator of t h e process variance for a long-memory
process because it has b o t h severe bias and low efficiency (Beran, 1992).
Due t o these problems in estimation, Allan (1966) criticised t h e use of
the sample variance as a meaningful estimator of variability for stationary
processes with long memory and for nonstationary processes with infinite
variance. He proposed an alternative theoretical measure of variability
t h a t is now known as t h e Allan variance (see Section 2 below). In terms
of filtering theory, t h e Allan variance can b e interpreted as t h e variance of
a process after it has been subjected t o an approximate band-pass filter of
'constant Q ' (i.e., t h e ratio of t h e center frequency of t h e pass-band a n d t h e
width of t h e pass-band is a constant). T h e chief advantages of t h e Allan
Long-Memory Processes and Wavelets 327
variance for long-memory processes are two-fold: first, this variance can b e
estimated without bias and with good efficiency for such processes, and,
second, estimates of the Allan variance can in t u r n be used to estimate
t h e parameters of t h e long-memory process (as we note in Section 2, these
statements also hold for certain nonstationary power-law processes). T h e
Allan variance has been applied for over twenty-five years as t h e routine
time domain measure of frequency stability in high-performance oscillators.
In a recent article Flandrin (1992) briefly noted t h a t the Allan variance
can b e interpreted in terms of the coefficients of a Haar wavelet transform
of a time series. We explore this connection in detail in Section 2. As
we note in Section 3, the notion of the Allan variance can be generalized
to other wavelets to define a wavelet variance. T h e wavelet variance is
a useful way of summarizing the properties of the wavelet transform for
certain processes. In particular, the parameters of long-memory processes
can b e deduced from the wavelet variance. We also note in Section 3 t h a t
s t a n d a r d estimators for the Allan variance can be easily generalized to these
other wavelet variances.
In terms of computational complexity, the Allan variance is t h e sim-
plest of the wavelet variances. Unfortunately, it can be misleading for
certain processes of interest in geophysics. In the particular example we
consider in Section 4, we find t h a t the wavelet variance for Daubechies's
'least asymmetric' wavelet of order 8 (Daubechies, 1992) yields markedly
b e t t e r results (hereinafter we refer to this wavelet as the LA(8) wavelet).
This example demonstrates the usefulness for d a t a analysis of classes of
wavelets beyond the simple Haar wavelet. Our analysis also demonstrates
the importance of conducting a parallel spectral analysis to validate t h e
interpretation of a wavelet variance in terms of the parameters for a long-
memory process. Finally, in Section 5 we discuss some possible extensions,
including some general thoughts on assessing the variability of wavelet co-
efficients.
§2. T h e A l l a n Variance a n d t h e H a a r W a v e l e t
Suppose we have a time series of length N t h a t can b e regarded as one
portion of one realization from the stochastic process {YJ,i = 0 , ± 1 , . . . }
(for convenience, we assume t h a t the sampling interval between consecutive
observations is unity). Let
^ « ^ Σ ^ Η (1)
i=o
represent t h e sample average of k consecutive observations, the latest one of
which is Y%. T h e Allan variance at scale k is denoted by σ\ (k) and is defined
to be half t h e mean square difference between adjacent nonoverlapping
328 D. Percival and P. Guttorp
Yt(kys; i.e.,
al(k)=\E{[Yt{k)-Yt-k{k)f}.
In order for 0"y(fc) to be independent of the index £, we must impose a
stationarity condition on the process {!*}, namely, t h a t its first backward
difference Zt = Yt — Yt-\ is a stationary process. Note t h a t the Allan
variance at scale k is a measure of how much averages of length k change
from one time period of length k to the next.
To see why the Allan variance might be of interest for long-memory
processes, suppose momentarily t h a t {Yt} is a fractional Gaussian process
with self-similarity parameter 1/2 < H < 1. If we let 5 y ( · ) denote the
spectral density function (sdf) for {Yt} defined for frequencies / between
— 1/2 and 1/2, then we have
Μ/) = Μ/)|/Γ-2H
where Li(·) is a slowly varying function for | / | —>> 0 (Beran, 1992). Thus,
if we plot log(5y ( / ) ) versus log(/) for positive values of / close to zero, we
will observe (to a good approximation) a line with a slope of 1 — 2H. A
similar result holds for the Allan variance: we have
a2Y(k) = L2(k)k2H~2,
Sy(f) Ξ 4sin
JgP-2
(7r/)
(this definition is motivated by the theory of linear filters). Suppose mo-
mentarily t h a t
Sy(f) = L3(f)\f\°,
where Ls(·) is a slowly varying function for | / | -> 0, and —3 < a < 0; i.e.,
SY(-) is a 'red' power-law process with exponent a greater t h a n —3. T h e n
we have
al(k) = L4(k)k~a-\
Long-Memory Processes and Wavelets 329
<***=(£) [Y2jk(k)-Y2jk-k(k)].
If the first backward difference process {Zt} for {Yt} is stationary with zero
mean, we then have
the time series contributes to exactly one d2- k in the summation in Equa-
tion 4. T h e estimator <Ty(&) is not commonly used because of the superior
statistical properties of a closely related estimator known as the maximal-
overlap estimator (Greenhall, 1991). This estimator is defined as
7 (l/y/2k, Z = 0,...,k-1;
Z,fc
~ \-l/>/2fc, Z = fe, . . . , 2 f e - l
then we have djtk — W2jk,kl i-e-> the drib's are obtained by appropriately
subsampling every 2/cth value of the Wife's. We refer to the Wife's as
the maximal-overlap Haar wavelet coefficients. We can now easily see the
difference between the non-overlapped and maximal-overlap estimators:
N / 2 k N
9 ~ 1 ~
? whereas fc w
v(fc) = Ü Σ ^ Μ > *y( ) = fe(^_2fc + 1) Σ *%; (7)
i.e., ôr y (k) makes use of just the N/2k subsampled Wt,k% whereas <5"y(fc)
makes use of all N — 2k -\- 1 of the Wtfk's. Thus the maximal-overlap
estimator does not use the usual decimated subseries of the discrete Haar
wavelet transform, but it can be said to use a uniformly sampled version
of the corresponding continuous transform. Note t h a t use of σ\ (k) r a t h e r
t h a n ^(k) imparts a certain degree of independence in the choice of the
origin (i.e., if we form a new time series by discarding Y\ and adding Y/v+i,
only one of the Wife's in the maximal-overlap estimator changes whereas
all of t h e m change in the nonoverlapped estimator). Note also t h a t the
definition of σ\(k) in Equation 5 holds for all sample sizes N (i.e., N need
not be a power of 2).
For computational purposes, it is more efficient to evaluate the right-
h a n d side of Equation 5 by computing the summations needed to obtain
t h e y t (fc)'s just once. To do so, let XQ = 0, and let Xt = Xt-i + Y% for
Long-Memory Processes and Wavelets 331
and
Wtik = (Xt - 2Xt-k + Xt-2k)/V2k.
We can thus rewrite Equation 5 as
§3. T h e W a v e l e t Variance
It is easy t o extend t h e ideas of t h e previous section t o wavelets other
t h a n t h e Haar wavelet. There are two approaches t h a t we can take t o gen-
erate t h e maximal-overlap wavelet coefficients Wt,k based upon a wavelet
of unit scale specified by t h e L\ filter coefficients /ΐο,ι, /ΐι,ι, . . . , /ΐχ,ι-ι,ι-
In t h e first approach, we start by generating t h e appropriate filter {h^k}
for each scale k based upon t h e wavelet for unit scale. This can b e done
easily by taking t h e inverse discrete wavelet transform of a properly placed
pulse (for details, see t h e subsection ' W h a t Do Wavelets Look Like?' in
Section 13.10 of Press et al., 1992). Let Lk b e t h e length of t h e wavelet
of scale k so t h a t , for example, L\ = 2 for t h e Haar wavelet while L\ = 8
for t h e LA(8) wavelet (the values of Lk for scales k = 2, 4, . . . , obey t h e
relationship Lk = (2k — l ) ( I q — 1) + 1 and can also b e obtained via t h e
recursive formula L2k — 2Lk + L\ — 2). For each scale k we then filter our
time series t o obtain
Lk-\
Wt,k= Σ Kkyt-h t= Lk,...,N
1=0
T h e usual pyramid algorithm uses the same scaling and wavelet filters at
each scale; i.e., at scale 2fc, these two filters are applied to the subsampled
o u t p u t from t h e scaling (low-pass) filter from scale k. For the maximal-
overlap estimator we must eliminate all subsampling, so formally we must
use different filters as we move from scale k to scale 2k. Forming these new
filters, however, is quite simple: the wavelet and scaling filters we need for
scale k in t h e maximal-overlap pyramid algorithm are obtained by insert-
ing k — 1 zeros between each of the coefficients in the wavelet and scaling
filters for unit scale (in effect, these zeros compensate for the elimination of
subsampling). T h e maximal-overlap pyramid algorithm avoids all multipli-
cations involving coefficients equal to zero by keeping track of the indices
of elements of the series t h a t need to be multiplied by the nonzero filter
coefficients.
Given the wavelet filter {7i/,i} and corresponding scaling filter {37,1}
for unit scale, the basic step of the maximal-overlap pyramid algorithm
takes as input
• a scale fc, which must be an integer power of 2, and
SetXt(Îîm+1=flb,iXt.
Set XN_Mh+t — ho,ixt·
Set u = t.
For / = 1 to L\ — 1, do the inner loop of 2:
Decrement u by k.
Increment Xt_m+i by gi,iXu>
Increment Xx_Mk+t by hlAXu.
E n d of t h e outer loop of 2.
Long-Memory Processes and Wavelets 333
N
1
ù W
r(k) = 771Γ7-Τ TT Y h
YK J
k(N - Lk + 1) £f *'*
(cf. Equation 7). A plot of t h e square root of this quantity versus scale
A: on a log/log scale yields a generalization of t h e σ-τ curve, from which
we can infer from regions of linearity t h a t a power law process might b e a
good model for our data.
334 D. Percival and P. Guttorp
1h
functions for the LA(8) wavelet, which has the following filter coefficients:
ho.i = 0.03222310060407815 h4A = 0.80373875180538600
hiji = 0.01260396726226383 h 5 |i = -0.49761866763256290
h2[i = -0.09921954357695636 h6[i = -0.02963552764596039
Η3Λ = -0.29785779560560505 h7[i = 0.07576571478935668
(the source of these coefficients is the 'N = 4' entry of Table 6.3, p. 198,
Daubechies, 1992, which gives the LA(8) scaling filter coefficients normal-
ized to sum to 2 - the above hi y s were obtained via the 'quadrature mirror'
relationship between scaling and wavelet filters with a renormalization so
t h a t Σ / = ο ^ ? ι = -0· Note t h a t the transfer functions for b o t h wavelets
roughly define a set of octave band filters; i.e., the transfer functions for
scale k are approximately concentrated between frequencies and l/4k and
l/2k (the spacing between the minor tick marks on the frequency axis is
1/16). T h e plots show t h a t the transfer functions for the LA(8) wavelet
are a b e t t e r approximation to a set of octave band filters t h a n those for t h e
Haar wavelet. T h e phase functions for the Haar wavelet are approximately
linear over the nominal pass-bands, whereas those for the LA(8) wavelet
are approximately constant and fairly close to zero. Thus the o u t p u t from
the Haar wavelet filters will be phase-shifted with respect to the input,
making it difficult to line up events at various scales with the original time
series. In contrast, because the filters for the LA(8) wavelet are approxi-
mately zero phase, we can more easily line up events at various scales with
the original time series. Note, however, t h a t the spans of Haar wavelet
filters for scales 1, 2, 4 and 8 are, respectively, 2, 4, 8 and 16, whereas
the corresponding spans for the LA(8) wavelet are 8, 22, 50 and 106. If
these wavelets are used as noncircular filters as discussed in Section 3, the
o u t p u t from the LA(8) wavelet will become increasingly shorter compared
to t h a t of the Haar wavelet as the scale increases. In fact, for the larger
scales the length of the LA(8) wavelet will exceed the length of the time
series, whereas the reverse will be true for the Haar wavelet.
Figure 3 shows (from b o t t o m to top) the renormalized o u t p u t s from
the Haar wavelet filters for physical scales 0.1, 0.2, 0.4, 0.8, 1.6 and 3.2
meters (because the distance between adjacent observations is 0.1 meters,
these physical scales correspond to, respectively, the unitless scales 1, 2,
4, 8, 16 and 32). In order to obtain physically meaningful units, it is
necessary to renormalize the Wife's by dividing by (2/c) 1 / 2 . Each of these
renormalized filtered series is drawn with the same vertical scale, so we
see t h a t the variability gets progressively larger as we move from shorter
to longer scales (the distance between minor tick marks on the vertical
scale is 1/s). T h e usual Haar wavelet transform for these scales can be
obtained by appropriately subsampling these filtered series. Note t h a t all
six of these filtered series appear to be approximately rescaled versions of
Long-Memory Processes and Wavelets 337
ΙΙ^ηΚΚ^
yvM^Jw^^ yVv#-
TO^^^^*>iH^ r*> » » « ^ » W ^ »»w"^Wa"WllVVv *WMh|'v|l'«'«%p>Ίν V
_L X I X J
350 450 550 650 750 850 950 1050
meters
Figure 3. Renormalized outputs of Haar wavelet filters for physical
scales 0.1, 0.2, 0.4, 0.8, 1.6 and 3.2 meters (bottom to top).
\^f^^jt^Hêjjl^
^'■"»"«flNi»
Tf|r~
■ »#» >WH"* iH ♦»« H"»fr»»M i»«»1
J_ _]_ X X _L X J
350 450 550 650 750 850 950 1050
meters
Figure 4. Renormalized outputs of LA(8) wavelet filters for physical
scales 0.1, 0.2, 0.4, 0.8, 1.6 and 3.2 meters (bottom to top).
that there is much less variability at the smaller scales for the LA(8) wavelet
than for the Haar wavelet. While there is some correspondence between the
filtered series at different scales, the correspondence is much less marked
for the LA(8) wavelet than for the Haar wavelet. Each of the filtered series
in this figure is slightly shorter than the corresponding one in Figure 3 due
to the longer span of the filters for the LA(8) wavelet. Note that only part
of the usual LA(8) wavelet transform for these scales can be obtained by
subsampling these filtered series because we are not filtering the data in
Figure 1 as if it were circular (an assumption that would make no sense
Long-Memory Processes and Wavelets 339
10 1
10°
io- 1
■ ■
io- 2 J I I I ' i l l
2 1 1 2 3 2 1 1
10- 10- 10° 10 10 10 10" 10" 10° 10 10 2 1 0 3
scale (meters) scale (meters)
F i g u r e 5. 'σ-τ' curves for the Haar wavelet (left-hand plot) and the
LA(8) wavelet (right-hand).
at all for a series of ocean depth measurements!). Again we can see the
bursts in t h e filtered series for the smaller scales near 450 and 1000 meters.
An examination of the renormalized filtered series for physical scales longer
t h a n 3.2 meters shows no indication of these bursts. In the region marked
by t h e thin vertical lines between the two bursts, the filtered series for all
scales appear to be consistent with a stationarity assumption, so we have
chosen this subseries of 4096 values as a candidate for analysis using the
ideas discussed in the previous two sections.
T h e left-hand plot of Figure 5 shows the ' σ - τ ' curve, a popular analysis
tool in the frequency stability literature. This plot shows the square root
of t h e Allan variance (estimated using the maximal-overlap estimator) at
different physical scales plotted versus scale on a log-log plot. Theory
suggests t h a t regions of linearity correspond to a power-law process over a
particular region of frequencies, with the exponent of the power-law process
being related to the slope of the line (in log-log space). T h e first 7 values of
the σ-τ curve fall on such a line almost perfectly. The line drawn through
t h e m on t h e plot was calculated via least squares and has a slope of 0.83.
Since ay{k) (the square root of the Allan variance) varies approximately as
£-(c*+i)/2 for a p 0 w e r _ i a w process with exponent a (see Section 2), the σ-τ
curve strongly suggests the presence of a power-law process over scales of
0.1 to 6.4 meters with an exponent of a = —2.66 = —8/3. T h e right-hand
plot is t h e σ-τ curve corresponding to the LA(8) wavelet for the first 9
scales (recall t h a t filtered series for the longer scales cannot be obtained
due to t h e span of the filters for this wavelet), and it tells quite a different
story. T h e values are not aligned in an obvious straight line as in t h e case
of t h e Allan variance, and the slope t h a t we found using the first 7 scales
of t h e Allan variance certainly does not look reasonable for portions of t h e
340 1D. Pe rciva / and ' P . (jUttOJ
io 3 r
102[
101 I
10" h
10-4
10-4
<, ' "■
10-4 ■Λν.
• 1 .
i 1 1
10-4
io-7 _ L _ L J
10": 10"2 10"1 10° 101
f (cycles/meter)
F i g u r e 6. Comparison of spectral estimates implied by σ-τ curves
with a WOSA spectral estimate.
§5. C o n c l u d i n g R e m a r k s
Here we make several remarks concerning the results of the previous
sections. First, a wavelet analysis of a time series t h a t can be modelled as
a power-law process can be used to deduce the properties of the underlying
process, b u t it is dangerous to do so without a careful look at a traditional
sdf estimate with good prevention against leakage. For example, if we had
computed just the left-hand σ-τ curve of Figure 5, we might have been
badly fooled by the degree to which the points line up as theory suggests
they should in the presence of a power-law process.
Second, as others have noted, the wavelet variance is a tool t h a t is
well-adapted for studying power-law processes; however, the wavelet vari-
ance can lead us to find 'power laws' (or 'fractal behavior') in d a t a t h a t
might be best modelled in other ways. Consider, for example, the octave
b a n d centered at 1 0 _ 1 cycles/meter in Figure 6. T h e wavelet-based esti-
mates of t h e power in this band are b o t h somewhat higher t h a n what the
WOSA sdf estimate suggests is reasonable (in fact, the deficiency of power
here can be a t t r i b u t e d to the preprocessing operation in which t h e veloc-
ity series was first differenced over an interval of 10 meters). This octave
band corresponds to a scale of 3.2 meters in the σ-τ curves of Figure 5.
For b o t h of these curves, the value at 3.2 meters can be obtained almost
exactly by linearly interpolating (on a log/log scale) between the points at
scales 1.6 and 6.4 meters. This suggests t h a t the wavelet-based estimates
might b e biased in the sense t h a t regions t h a t do not agree with a nominal
power-law behavior will tend to be 'filled in' in a manner consistent with
the power-law assumption. W h a t is vitally needed is a careful study of the
bias and variance of wavelet-based estimates of the sdf for processes t h a t
342 D. Percival and P. Guttorp
References
1. Allan, D. W., Statistics of atomic frequency clocks, Proc. IEEE, 31,
221-30, 1966.
2. Beran, J., Statistical methods for data with long-range dependence,
Statist. Sei., 7, 404-16, 1992.
3. Beran, J. and H. R. Künsch, Location estimators for processes with
long-range dependence, Research Report 40, Seminar für Statistik,
ETH, Zürich, 1985.
4. Daubechies, I., Ten lectures on wavelets, SI AM, Philadelphia, 1992.
5. Flandrin, P., Wavelet analysis and synthesis of fractional brownian
motion, IEEE Trans. Info. Theo., 38, 910-7, 1992.
6. Greenhall, C. A., Recipes for degrees of freedom of frequency stability
estimators, IEEE Trans. Instr. Meas., 40, 994-9, 1991.
7. Hurst, H. E., Long-term storage capacity of reservoirs, Trans. Amer.
Soc. Civ. Eng., 116, 770-9, 1951.
8. Mandelbrot, B. B. and J. W. Van Ness, Fractional Brownian motions,
fractional noises and applications, SI AM Rev., 10, 422-37, 1968.
9. Mandelbrot, B. B. and J. R. Wallis, Computer experiments with frac-
tional Gaussian noises. Part 1, averages and variances, Water Resour.
Resear., 5, 228-41, 1969.
10. Newcomb, S., Astronomical constants (the elements of the four inner
planets and the fundamental constants of astronomy), Supplement to
the American Ephemeris and Nautical Almanac for 1897, U.S. Govern-
ment Printing Office, Washington, 1895.
11. Percival, D. B., The statistics of long memory processes, unpublished
Ph.D. dissertation, Department of Statistics, University of Washington,
1983.
12. Percival, D. B., On the sample mean and variance of a long memory
process, Technical Report 69, Department of Statistics, University of
Washington, 1985.
13. Percival, D. B. and A. T. Waiden, Spectral analysis for physical appli-
cations: multitaper and conventional univariate techniques, Cambridge
University Press, Cambridge, 1993.
14. Press, W. H., B. P. Flannery, S. A. Teukolsky and W. T. Vetterling, Nu-
merical recipes: the art of scientific computing (second edition), Cam-
bridge University Press, Cambridge, 1992.
15. Raftery, A. E. and J. Haslett, Space-time modelling with long-memory
dependence: assessing Ireland's wind power resource (with discussion),
J. Roy. Statist. Soc, Ser. C, 38, 1-21, 1989.
16. Smith, H. F., An empirical law describing heterogeneity in the yields
of agricultural crops, J. Agri. Sei., 28, 1-23, 1938.
17. Smith, R. L., Comment [on Beran, 1992], Statist. Sei., 7, 422-5, 1992.
344 D. Percival and P. Guttorp
Robert Spindel and John Harlett of the Applied Physics Laboratory, Uni-
versity of Washington, graciously provided discretionary funding to support
Percival's work on this manuscript. Guttorp's work was partially funded
by NSF grant DM S-9115756. The authors wish to thank Mike Gregg for
supplying the data used in Section 4 and also to thank Chuck Greenhall
and Emma McCoy for their helpful critiques.
Donald B. Percival
Applied Physics Laboratory
HN-10
University of Washington
Seattle, WA 98195
e-mail: [email protected]
Peter Guttorp
Department of Statistics
GN-22
University of Washington
Seattle, WA 98195
e-mail: peterùstat. Washington, edu
Bibliography
776, 1990.
66. Farge, M., E. Goirand, and V. Wickerhauser, Wavelet packets analysis
compression and filtering of two-dimensional turbulent flows, LMD,
Ecole Normale Supérieure, Paris, 1991, Preprint.
67. Farge, M., E. Goirand, Y. Meyer, F. Pascal, and M. V. Wicker-
hauser, Improved predictability of two-dimensional turbulent flows
using wavelet packet compression, Fluid Dyn. Res., 1 0 ( 4 - 6 ) , 2 2 9 -
250, 1992.
68. Field, D. J., Scale-invariance and self-similar 'Wavelet' transforms: an
analysis of natural scenes and mammalian visual systems, in Wavelets,
Fractals, and Fourier Transforms, M. Farge et al. (eds.), Clarendon
Press, Oxford, 1993.
69. Flandrin, P., Time-frequency and time-scale, Proc. of J^th Acoust.,
Speech and Signal Processing Workshop on Spectrum Estimation Mod-
eling, 77-80, 1988.
70. Flandrin, P., Fractional Brownian motion and wavelets, in Wavelets,
Fractals, and Fourier Transforms, M. Farge et al. (eds.), Clarendon
Press, Oxford, 1993.
71. Frick, P. and V. Zimin, Hierarchical models of turbulence, in Wavelets,
Fractals, and Fourier Transforms, M. Farge et al. (eds.), Clarendon
Press, Oxford, 1993.
72. Gamage, N. K. K., Modeling and analysis of geophysical turbulence:
Use of optimal transforms and basis sets, Ph.D. Thesis, Oregon State
University, 135 pp., 1989.
73. Gamage, N. and W. Blumen, Comparative analysis of low level cold
fronts: wavelet, fourier, and empirical orthogonal function decompo-
sitions, Monthly Weather Rev., 1 2 1 , 2867-2878, 1993.
74. Gamage, N. K. K. and C. Hagelberg, Detection and analysis of mi-
crofronts and associated coherent events using localized transforms,
J. Atmos. Sei., 5 0 ( 5 ) , 750-756, 1993.
75. Gambis, D., Wavelet transform analysis of the length of the day and
the El-Nino/Southern Oscillation variations at intraseasonal and in-
terannual time scales, Ann. Geophysicae, 10, 331-371, 1992.
76. Genon-Catalot, V., C. Laredo, and D. Picard, Nonparametric esti-
mation of the diffusion coefficient by wavelet methods, Scandinavian
Jour, of Statistics, 1 9 ( 4 ) , 317-336, 1992.
77. Ghez, J. M. and S. Vaienti, On the wavelet analysis for multifractal
sets, J. Stat. Phys., 57, 415-420, 1989.
78. Giannakis, G. B., Wavelet parameter and phase estimation using
Cumulant slices, IEEE Trans, on Geoscience and Remote Sensing,
2 7 ( 4 ) , 452-455, 1989.
79. Glowinski, R., W. M. Lawton, M. Ravachol, E. Tenenbaum, Wavelet
solutions of linear and nonlinear elliptic, parabolic and hyperbolic
Bibliography 351
May 1993.
124. Otaguro, T., S. Takagi, and H. Satoy, P a t t e r n search in a turbulent
signal using wavelet analysis, Proc. 21st Japan Symp. on Turbulence,
Tokyo, J a p a n , 1989.
125. Pentland, A. P., Surface interpolation using wavelets, in Proc of 2nd
European Conference on Computer Vision, G. Sandini (ed.), Springer-
Verlag, 615-619, 1992.
126. Pentland, A. P. and B. Horowitz, A practical approach to fractal-
based image compression, in Data Compression Conference, J. A.
Storer and J. H. Reif (eds.), I E E E Comput. Soc. Press, Los Alamitos,
176-185, 1991.
127. P e r m a n n , D. and I. Hamilton, Wavelet analysis of time series for the
duffing oscillator: The detection of order within chaos, Phys. Rev.
Lett, 6 9 ( 1 8 ) , 2607-2610, 1992.
128. Persoglia, S., S. Sancin, and A. Vesnaver, Adaptive deconvolution by
lattice filters: Experience on synthetic and real data, Bollettino di
Geofisica Teorica ed Applicata, 2 7 ( 1 0 7 ) , 169-183, 1985.
129. Phuvan, S., T. K. Oh, N. P. Caviris, Y. Li, and H. H. Szu, Texture
analysis by space-filling curves and one-dimensional Haar wavelets,
Opt. Eng., 3 1 ( 9 ) , 1899-1906, 1992.
130. Pittner, S., J. Schneid, and C. W. Ueberhuber, Wavelet Literature
Survey, Institute for Applied and Numerical Mathematics, Technical
University Vienna, Wien, Austria, 1993.
131. Pouligny, B., G. Gabriel, J. F. Muzy, A Arnéodo, F. Argoul, and
E. Freysz, Optical wavelet transform and local scaling properties of
fractals, J. Appl. Cryst., 2 4 ( 5 ) , 526-530, 1991.
132. Powell, D. and C. E. Elderkin, An investigation of the application
of Taylor's hypothesis to atmospheric boundary layer turbulence, J.
Atmos. Sei., 3 1 , 990-1002, 1974.
133. R a m a n a t h a n , J. and O. Zeitouni, On the wavelet transform of frac-
tional Brownian motion, IEEE Trans, on Inform. Theory, 3 7 ( 4 ) ,
1156-1158, 1991.
134. Ranchin, T. and L. Wald, The wavelet transform for the analysis
of remotely sensed images, Int. J. Remote Sensing, 1 4 ( 3 ) , 615-619,
1993.
135. Rasmussen, H. 0 . , The wavelet Gibbs phenomenon, in Wavelets,
Fractals, and Fourier Transforms, M. Farge et al. (eds.), Clarendon
Press, Oxford, 1993.
136. Redondo, J. M., Fractal models of density interfaces, in Wavelets,
Fractals, and Fourier Transforms, M. Farge et al. (eds.), Clarendon
Press, Oxford, 1993.
137. Redondo, J. M., R. M. Gonzalez, and J. L. Cano, Fractal aggregates
in the atmosphere, in Wavelets, Fractals, and Fourier Transforms, M.
Bibliography 355
367
368 Detailed Table of Contents
Bibliography 345
373