Você está na página 1de 112

2 Deconvolution

Introduction The Convolutional Model The Convolutional Model in the Time Domain The Convolu-
tional Model in the Frequency Domain Inverse Filtering The Inverse of the Source Wavelet Least-Squares
Inverse Filtering Minimum Phase Optimum Wiener Filters Spiking Deconvolution Prewhitening Wavelet
Processing by Shaping Filters Predictive Deconvolution Predictive Deconvolution in Practice Operator
Length Prediction Lag Percent Prewhitening Eect of Random Noise on Deconvolution Multiple Atten-
uation Field Data Examples Prestack Deconvolution Signature Deconvolution Vibroseis Deconvolution
Poststack Deconvolution The Problem of Nonstationarity Time-Variant Deconvolution Time-Variant
Spectral Whitening Frequency-Domain Deconvolution Inverse Q Filtering Deconvolution Strategies Exer-
cises Appendix B: Mathematical Foundation of Deconvolution Synthetic Seismogram The Inverse of
the Source Wavelet The Inverse Filter Frequency-Domain Deconvolution Optimum Wiener Filters Spiking
Deconvolution Predictive Deconvolution Surface-Consistent Deconvolution Inverse Q Filtering References
2.0 INTRODUCTION
Deconvolution compresses the basic wavelet in the
recorded seismogram, attenuates reverberations and
short-period multiples, thus increases temporal resolu-
tion and yields a representation of subsurface reec-
tivity. The process normally is applied before stack;
however, it also is common to apply deconvolution
to stacked data. Figure 2.0-1 shows a stacked section
with and without deconvolution. Deconvolution has
produced a section with a much higher temporal res-
olution. The ringy character of the stack without de-
convolution limits resolution, considerably.
Figure 2.0-2 shows selected common-midpoint
(CMP) gathers from a marine line before and after de-
convolution. Note that the prominent reections stand
out more distinctly on the deconvolved gathers. Decon-
volution has removed a considerable amount of ringy-
ness, while it has compressed the waveform at each of
the prominent reections. The stacked sections associ-
ated with these CMP gathers are shown in Figure 2.0-
3. The improvement observed on the deconvolved CMP
gathers also are noted on the corresponding stacked sec-
tion.
Figure 2.0-4 shows some NMO-corrected CMP
gathers from a land line with and without deconvo-
lution. Corresponding stacked sections are shown in
Figure 2.0-5. Again, note that deconvolution has com-
pressed the wavelet and removed much of the reverber-
ating energy.
Deconvolution sometimes does more than just
wavelet compression; it can remove a signicant part
of the multiple energy from the section. Note that the
stacked section in Figure 2.0-6 shows a marked improve-
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
160 Seismic Data Analysis
FIG. 2.0-1. Interpreters prefer the crisp, nely detailed appearance of the deconvolved section (right) as opposed to the
blurred, ringy appearance of the section without deconvolution (left). (Data courtesy Enterprise Oil.)
ment between 2 and 4 s after deconvolution.
To understand deconvolution, rst we need to ex-
amine the constituent elements of a recorded seismic
trace (Section 2.1). The earth is composed of layers of
rocks with dierent lithology and physical properties.
Seismically, rock layers are dened by the densities and
velocities with which seismic waves propagate through
them. The product of density and velocity is called seis-
mic impedance. The impedance contrast between adja-
cent rock layers causes the reections that are recorded
along a surface prole. The recorded seismogram can
be modeled as a convolution of the earths impulse re-
sponse with the seismic wavelet. This wavelet has many
components, including source signature, recording l-
ter, surface reections, and receiver-array response. The
earths impulse response is what would be recorded if
the wavelet were just a spike. The impulse response
comprises primary reections (reectivity series) and all
possible multiples.
Ideally, deconvolution should compress the wavelet
components and eliminate multiples, leaving only the
earths reectivity in the seismic trace. Wavelet com-
pression can be done using an inverse lter as a de-
convolution operator. An inverse lter, when convolved
with the seismic wavelet, converts it to a spike (Section
2.2). When applied to a seismogram, the inverse lter
should yield the earths impulse response. An accurate
inverse lter design is achieved using the least-squares
method (Section 2.2).
The fundamental assumption underlying the de-
convolution process (with the usual case of unknown
source wavelet) is that of minimum phase. This issue is
dealt with also in Section 2.2.
The optimum Wiener lter, which has a wide
range of applications, is discussed in Section 2.3. The
Wiener lter converts the seismic wavelet into any de-
sired shape. For example, much like the inverse lter,
a Wiener lter can be designed to convert the seismic
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
Deconvolution 161
FIG. 2.0-2. Note the prominent reections on the deconvolved gathers (b). The reverberations would make it dicult to
distinguish prominent reections on the undeconvolved gathers (a).
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
162 Seismic Data Analysis
FIG. 2.0-3. (a) The section obtained from the undeconvolved gathers of Figure 2.0-2a, and (b) the section obtained from
the deconvolved gathers of Figure 2.0-2b.
wavelet into a spike. However, the Wiener lter diers
from the inverse lter in that it is optimal in the least-
squares sense. Also, the resolution (spikiness) of the
output can be controlled by designing a Wiener pre-
diction error lter the basis for predictive deconvo-
lution (Section 2.3). Converting the seismic wavelet into
a spike is like asking for a perfect resolution. In prac-
tice, because of noise in the seismogram and assump-
tions made about the seismic wavelet and the recorded
seismogram, spiking deconvolution is not always desir-
able. Finally, the prediction error lter can be used to
remove periodic components multiples, from the seis-
mogram. Practical aspects of predictive deconvolution
are presented in Section 2.4, and eld data examples are
provided in Section 2.5. Finally, time-varying aspects of
the source waveform nonstationarity, are discussed in
Section 2.6.
The mathematical treatment of deconvolution is
found in Appendix B. However, several numerical ex-
amples, which provide the theoretical groundwork from
a heuristic viewpoint, are given in the text. Much of the
early theoretical work on deconvolution came from the
MIT Geophysical Analysis Group, which was formed in
the mid-1950s.
2.1 THE CONVOLUTIONAL MODEL
A sonic log segment is shown in Figure 2.1-1a. The
sonic log is a plot of interval velocity as a function of
depth based on downhole measurement using logging
tools. Here, velocities were measured between the 1000-
to 5400-ft depth interval at 2-ft intervals. The veloc-
ity function was extrapolated to the surface by a lin-
ear ramp. The sonic log exhibits a strong low-frequency
component with a distinct blocky character represent-
ing gross velocity variations. Actually, it is this low-
frequency component that normally is estimated by ve-
locity analysis of CMP gathers (Section 3.2).
In many sonic logs, the low-frequency component
is an expression of the general increase of velocity with
depth due to compaction. In some sonic logs, however,
the low-frequency component exhibits a blocky charac-
ter (Figure 2.1-1a), which is due to large-scale lithologic
variations. Based on this blocky character, we may de-
ne layers of constant interval velocity (Table 2-1), each
of which can be associated with a geologic formation
(Table 2-2).
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
Deconvolution 163
FIG. 2.0-4. Some NMO-corrected gathers associated with the stacked sections in Figure 2.0-5, (a) before, (b) after deconvo-
lution. Deconvolution has removed the ringy character from the data.
Table 2-1. The interval velocity trend obtained from
the sonic log in Figure 2.1-1a.
Layer
Number Interval Velocity,* ft/s Depth Range, ft
1 21 000 1000 2000
2 19 000 2000 2250
3 18 750 2250 2500
4 12 650 2500 3775
5 19 650 3775 5400
*The velocity in Layer 2 gradually decreases from the
top of the layer to the bottom.
Table 2-2. Stratigraphic identication associated with
layering described in Table 2-1.
Layer Number Lithologic Unit
1 Limestone
2 Shaly limestone with gradual
increase in shale content
3 Shaly limestone
4 Sandstone
5 Dolomite
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
164 Seismic Data Analysis
FIG. 2.0-5. Deconvolution helps distinguish prominent reections with ease (b). However, on a section without deconvolution
(a), reections are buried in reverberating energy. Selected CMP gathers for both sections are shown in Figure 2.0-4.
The sonic log also has a high-frequency component
superimposed on the low-frequency component. These
rapid uctuations can be attributed to changes in rock
properties that are local in nature. For example, the
limestone layer can have interbeddings of shale and
sand. Porosity changes also can aect interval velocities
within a rock layer. Note that well-log measurements
have a limited accuracy; therefore, some of the high-
frequency variations, particularly those associated with
a rst arrival that is strong enough to trigger one re-
ceiver but not the other in the log tool (cycle skips), are
not due to changes in lithology.
Well-log measurements of velocity and density pro-
vide a link between seismic data and the geology of the
substrata. We now explain the link between log mea-
surements and the recorded seismic trace provided by
seismic impedance the product of density and veloc-
ity. The rst set of assumptions that is used to build
the forward model for the seismic trace follows:
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
Deconvolution 165
FIG. 2.0-6. CMP stacks with (a) no deconvolution before
stack, (b) spiking deconvolution before stack. Deconvolution
can remove a signicant amount of multiple energy from
seismic data. (Data courtesy Elf Aquitane and partners.)
Assumption 1. The earth is made up of horizontal
layers of constant velocity.
Assumption 2. The source generates a compres-
sional plane wave that impinges on layer bound-
aries at normal incidence. Under such circum-
stances, no shear waves are generated.
Assumption 1 is violated in both structurally com-
plex areas and in areas with gross lateral facies changes.
Assumption 2 implies that our forward model for the
seismic trace is based on zero-oset recording an un-
realizable experiment. Nevertheless, if the layer bound-
aries were deep in relation to cable length, we assume
that the angle of incidence at a given boundary is small
and ignore the angle dependence of reection coe-
cients. Combination of the two assumptions thus imply
a normal-incidence one-dimensional (1-D) seismogram.
Based on assumptions 1 and 2, the reection coe-
cient c (for pressure or stress), which is associated with
the boundary between, say, layers 1 and 2, is dened as
c =
I
2
I
1
I
2
+ I
1
, (2 1a)
where I is the seismic impedance associated with each
layer given by the product of density and compres-
sional velocity v.
From well-log measurements, we nd that the ver-
tical density gradient often is much smaller than the
vertical velocity gradient. Therefore, we often assume
that the impedance contrast between rock layers is es-
sentially due to the velocity contrast, only. Equation
(2-1a) then takes the form:
c =
v
2
v
1
v
2
+ v
1
. (2 1b)
If v
2
is greater than v
1
, the reection coecient would
be positive. If v
2
is less than v
1
, then the reection
coecient would be negative.
The assumption that density is invariant with
depth or that it does not vary as much as velocity is
not always valid. The reason we can get away with it
is that the density gradient usually has the same sign
as the velocity gradient. Hence, the impedance function
derived from the velocity function only should be cor-
rect within a scale factor.
For vertical incidence, the reection coecient is
the ratio of the reected wave amplitude to the incident
wave amplitude. Moreover, from its denition (equation
2-1a), the reection coecient is seen as the ratio of
the change in acoustic impedance to twice the average
acoustic impedance. Therefore, seismic amplitudes as-
sociated with earth models with horizontal layers and
vertical incidence (assumptions 1 and 2) are related to
acoustic impedance variations.
The reection coecient series c(z), where z is the
depth variable, is derived from sonic log v(z) and is
shown in Figure 2.1-1b. We note the following:
(a) The position of each spike gives the depth of the
layer boundary, and
(b) the magnitude of each spike corresponds to the
fraction of a unit-amplitude downward-traveling
incident plane wave that would be reected from
the layer boundary.
To convert the reection coecient series c(z) (Fig-
ure 2.1-1b) derived from the sonic log into a time series
c(t), select a sampling interval, say 2 ms. Then use the
velocity information in the log (Figure 2.1-1a) to con-
vert the depth axis to a two-way vertical time axis. The
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
166 Seismic Data Analysis
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
Deconvolution 167
FIG. 2.1-2. A seismic source wavelet after onset takes the
form shown at top left. As the wavelet travels into the earth,
the amplitude level drops (geometric spreading) and a loss
of high frequencies occurs (frequency absorption).
result of this conversion is shown in Figure 2.1-1c, both
as a conventional wiggle trace and as a variable area
and wiggle trace (the same trace repeated six times to
highlight strong reections). The reection coecient
series c(t) (Figure 2.1-1c) represents the reectivity of a
series of ctitious layer boundaries that are separated by
an equal time interval the sampling rate (Goupillaud,
1961). The major events in this reectivity series are
from the boundary between layers 2 and 3 located at
about 0.3 s, and the boundary between layers 4 and 5
located at about 0.5 s.
The reection coecient series (Figure 2.1-1c) that
was constructed is composed only of primary reec-
tions (energy that was reected only once). To get a
complete 1-D response of the horizontally-layered earth
model (assumption 1), multiple reections of all types
(surface, intrabed and interbed multiples) must be in-
cluded. If the source were unit-amplitude spike, then the
recorded zero-oset seismogram would be the impulse
response of the earth, which includes primary and mul-
tiple reections. Here, the Kunetz method (Claerbout,
1976) is used to obtain such an impulse response. The
impulse response derived from the reection coecient
series in Figure 2.1-1c is shown in Figure 2.1-1d with
the variable area and wiggle display.
The characteristic pressure wave created by an im-
pulsive source, such as dynamite or air gun, is called the
signature of the source. All signatures can be described
as band-limited wavelets of nite duration for exam-
ple, the measured signature of an Aquapulse source in
Figure 2.1-2. As this waveform travels into the earth,
its overall amplitude decays because of wavefront diver-
gence. Additionally, frequencies are attenuated because
of the absorption eects of rocks (see Section 1.4). The
progressive change of the source wavelet in time and
depth also is shown in Figure 2.1-2. At any given time,
the wavelet is not the same as it was at the onset of
source excitation. This time-dependent change in wave-
form is called nonstationarity.
Wavefront divergence is removed by applying a
spherical spreading function (Section 1.2). Frequency
attenuation is compensated for by the processing tech-
niques discussed in Section 2.6. Nevertheless, the simple
convolutional model discussed here does not incorporate
nonstationarity. This leads to the following assumption:
Assumption 3. The source waveform does not
change as it travels in the subsurface it is sta-
tionary.
The Convolutional Model in
the Time Domain
A convolutional model for the recorded seismogram
now can be proposed. Suppose a vertically propagat-
ing downgoing plane wave with source signature (Figure
2.1-3a) travels in depth and encounters a layer bound-
ary at 0.2-s two-way time. The reection coecient as-
sociated with the boundary is represented by the spike
in Figure 2.1-3b. As a result of reection, the source
wavelet replicates itself such that it is scaled by the re-
ection coecient. If we have a number of layer bound-
aries represented by the individual spikes in Figures 2.1-
3b through 2.1-3f, then the wavelet replicates itself at
those boundaries in the same manner. If the reection
coecient is negative, then the wavelet replicates itself
with its polarity reversed, as in Figure 2.1-3c.
Now consider the ensemble of the reection coe-
cients in Figure 2.1-3g. The response of this sparse spike
series to the basic wavelet is a superposition of the in-
dividual impulse responses. This linear process is called
the principle of superposition. It is achieved computa-
tionally by convolving the basic wavelet with the reec-
tivity series (Figure 2.1-3g). The convolutional process
already was demonstrated by the numerical example in
Section 1.1.
The response of the sparse spike series to the basic
wavelet in Figure 2.1-3g has some important character-
istics. Note that for events at 0.2 and 0.35 s, we iden-
tify two layer boundaries. However, to identify the three
closely spaced reecting boundaries from the composite
response (at around 0.6 s), the source waveform must be
removed to obtain the sparse spike series. This removal
process is just the opposite of the convolutional process
used to obtain the response of the reectivity series to
the basic wavelet. The reverse process appropriately is
called deconvolution.
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
168 Seismic Data Analysis
FIG. 2.1-3. A wavelet (a) traveling in the earth repeats itself when it encounters a reector along its path (b, c, d, e, f). The
left column represents the reection coecients, while the right column represents the response to the wavelet. Amplitudes
of the response are scaled by the reection coecient. The resulting seismogram (bottom right) represents the composite
response of the earths reectivity (bottom left) to the wavelet (top right).
The principle of superposition now is applied to the
impulse response derived from the sonic log in Figure
2.1-1d. Convolution of a source signature with the im-
pulse response yields the synthetic seismogram shown
in Figure 2.1-4. The synthetic seismogram also is shown
in Figure 2.1-1e. This 1-D zero-oset seismogram is free
of random ambient noise. For a more realistic represen-
tation of a recorded seismogram, noise is added (Figure
2.1-4).
The convolutional model of the recorded seismo-
gram now is complete. Mathematically, the convolu-
tional model illustrated in Figure 2.1-4 is given by
x(t) = w(t) e(t) + n(t), (2 2a)
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
Deconvolution 169
FIG. 2.1-4. The top frame is the same as in Figure 2.1-1d.
The asterisk denotes convolution. The recorded seismogram
(bottom frame) is the sum of the noise-free seismogram and
the noise trace. This gure is equivalent to equation (2-2a).
where x(t) is the recorded seismogram, w(t) is the ba-
sic seismic wavelet, e(t) is the earths impulse response,
n(t) is the random ambient noise, and denotes con-
volution. Deconvolution tries to recover the reectivity
series (strictly speaking, the impulse response) from the
recorded seismogram.
An alternative to the convolutional model given by
equation (2-2a) is based on a surface-consistent spec-
tral decomposition (Taner and Coburn, 1981). In such
a formulation, the seismic trace is decomposed into the
convolutional eects of source, receiver, oset, and the
earths impulse response, thus explicitly accounting for
variations in wavelet shape caused by near-source and
near-receiver conditions and source-receiver separation.
The following equation describes the surface-consistent
convolutional model (Section B.8):
x

ij
(t) = s
j
(t) h
l
(t) e
k
(t) g
i
(t) + n(t), (2 2b)
where x

ij
(t) is a model of the recorded seismogram,
s
j
(t) is the waveform component associated with source
location j, g
i
(t) is the component associated with re-
ceiver location i, and h
l
(t) is the component associated
with oset dependency of the waveform dened for each
oset index l = |ij|. As in equation (2-2a), e
k
(t) repre-
sents the earths impulse response at the source-receiver
midpoint location, k = (i + j)/2. By comparing equa-
tions (2-2a) and (2-2b), we infer that w(t) represents
the combined eects of s(t), h(t), and g(t).
The assumption of surface-consistency implies that
the basic wavelet shape depends only on the source
and receiver locations, not on the details of the ray-
path from source to reector to receiver. In a transition
zone, surface conditions at the source and receiver lo-
cations may vary signicantly from dry to wet surface
conditions. Hence, the most likely situation where the
surface-consistent convolutional model may be applica-
ble is with transition-zone data. Nevertheless, the for-
mulation described in this section is the most accepted
model for the 1-D seismogram.
The random noise present in the recorded seismo-
gram has several sources. External sources are wind mo-
tion, environmental noise, or a geophone loosely coupled
to the ground. Internal noise can arise from the record-
ing instruments. A pure-noise seismogram and its char-
acteristics are shown in Figure 2.1-5. A pure random-
noise series has a white spectrum it contains all the
frequencies. This means that the autocorrelation func-
tion is a spike at zero lag and zero at all other lags. From
Figure 2.1-5, note that these characteristic requirements
are reasonably satised.
Now examine the equation for the convolutional
model. All that normally is known in equation (2-2a) is
x(t) the recorded seismogram. The earths impulse
response e(t) must be estimated everywhere except at
the location of wells with good sonic logs. Also, the
source waveform w(t) normally is unknown. In certain
cases, however, the source waveform is partly known;
for example, the signature of an air-gun array can be
measured. However, what is measured is only the wave-
form at the very onset of excitation of the source array,
and not the wavelet that is recorded at the receiver.
Finally, there is no a priori knowledge of the ambient
noise n(t).
We now have three unknowns w(t), e(t), and
n(t), one known x(t), and one single equation (2-2a).
Can this problem be solved? Pessimists would say no.
However, in practice, deconvolution is applied to seismic
data as an integral part of conventional processing and
is an eective method to increase temporal resolution.
To solve for the unknown e(t) in equation (2-2a),
further assumptions must be made.
Assumption 4. The noise component n(t) is zero.
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
170 Seismic Data Analysis
FIG. 2.1-5. A random signal with innite length has a at
amplitude spectrum and an autocorrelogram that is zero
at all lags except the zero lag. The discrete random series
with nite length shown here seems to satisfy these require-
ments. What distinguishes a random signal from a spike
(1, 0, 0, . . .)?
Assumption 5. The source waveform is known.
Under these assumptions, we have one equation,
x(t) = w(t) e(t). (2 3a)
and one unknown, the reectivity series e(t). In reality,
however, neither of the above two assumptions normally
is valid. Therefore, the convolutional model is examined
further in the next section, this time in the frequency
domain, to relax assumption 5.
If the source waveform were known (such as the
recorded source signature), then the solution to the de-
convolution problem is deterministic. In Section 2.2, one
such method of solving for e(t) is considered. If the
source waveform were unknown (the usual case), then
the solution to the deconvolution problem is statistical.
The Wiener prediction theory (Section 2.3) provides one
method of statistical deconvolution.
The Convolutional Model in
the Frequency Domain
The convolutional model for the noise-free seismogram
(assumption 4) is represented by equation (2-3a). Con-
volution in the time domain is equivalent to multiplica-
tion in the frequency domain (Section A.1). This means
that the the amplitude spectrum of the seismogram
equals the product of the amplitude spectra of the seis-
mic wavelet and the earths impulse response (Section
B.1):
A
x
() = A
w
()A
e
(), (2 3b)
where A
x
(), A
w
(), and A
e
() are the amplitude spec-
tra of x(t), w(t), and e(t), respectively.
Figure 2.1-6 shows the amplitude spectra (top row)
of the impulse response e(t), the seismic wavelet w(t),
and the seismogram x(t). The impulse response is the
same as that shown in Figure 2.1-1d. The similarity in
the overall shape between the amplitude spectrum of
the wavelet and that of the seismogram is apparent. In
fact, a smoothed version of the amplitude spectrum of
the seismogram is nearly indistinguishable from the am-
plitude spectrum of the wavelet. It generally is thought
that the rapid uctuations observed in the amplitude
spectrum of a seismogram are a manifestation of the
earths impulse response, while the basic shape is asso-
ciated primarily with the source wavelet.
Mathematically, the similarity between the ampli-
tude spectra of the seismogram and the wavelet suggests
that the amplitude spectrum of the earths impulse re-
sponse must be nearly at (Section B.1). By examin-
ing the amplitude spectrum of the impulse response in
Figure 2.1-6, we see that it spans virtually the entire
spectral bandwidth. As seen in Figure 2.1-5, a time se-
ries that represents a random process has a at (white)
spectrum over the entire spectral bandwidth. From close
examination of the amplitude spectrum of the impulse
response in Figure 2.1-6, we see that it is not entirely
at the high-frequency components have a tendency
to strengthen gradually. Thus, reectivity is not entirely
a random process. In fact, this has been observed in the
spectral properties of reectivity functions derived from
a worldwide selection of sonic logs (Walden and Hosken,
1984).
We now study the autocorrelation functions (mid-
dle row, Figure 2.1-6) of the impulse response, seismic
wavelet, and synthetic seismogram. Note that the au-
tocorrelation functions of the basic wavelet and seismo-
gram also are similar. This similarity is conned to lags
for which the autocorrelation of the wavelet is nonzero.
Mathematically, the similarity between the autocorrelo-
gram of the wavelet and that of the seismogram suggests
that the impulse response has an autocorrelation func-
tion that is small at all lags except the zero lag (Section
B.1). The autocorrelation function of the random series
in Figure 2.1-5 also has similar characteristics. How-
ever, there is one subtle dierence. When compared,
Figures 2.1-5 and 2.1-6 show that autocorrelation of
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
Deconvolution 171
FIG. 2.1-6. Convolution of the earths impulse response (a) with the wavelet (b) (equation 2-2a) yields the seismogram (c)
(bottom row). This process also is convolutional in terms of their autocorrelograms (middle row) and multiplicative in terms
of their amplitude spectra (top row). Assumption 6 (white reectivity) is based on the similarity between autocorrelograms
and amplitude spectra of the impulse response and wavelet.
the impulse response has a signicantly large negative
lag value following the zero lag. This is not the case
for the autocorrelation of random noise. The positive
peak (zero lag) followed by the smaller negative peak in
the autocorrelogram of the impulse response arises from
the spectral behavior discussed above. In particular, the
positive peak and the adjacent, smaller negative peak of
the autocorrelogram together nearly act as a fractional
derivative operator (Section A.1), which has a ramp ef-
fect on the amplitude spectrum of the impulse response
as seen in Figure 2.1-6.
The above observations made on the amplitude
spectra and autocorrelation functions (Figure 2.1-6) im-
ply that reectivity is not entirely a random process.
Nonetheless, the following assumption almost always is
made about reectivity to replace the statement made
in assumption 5.
Assumption 6. Reectivity is a random process.
This implies that the seismogram has the charac-
teristics of the seismic wavelet in that their auto-
correlations and amplitude spectra are similar.
This assumption is the key to implementing the pre-
dictive deconvolution. It allows the autocorrelation of
the seismogram, which is known, to be substituted for
the autocorrelation of the seismic wavelet, which is un-
known. In Section 2.3, we shall see that as a result of
assumption 6, an inverse lter can be estimated directly
from the autocorrelation of the seismogram. For this
type of deconvolution, Assumption 5, which is almost
never met in reality, is not required. But rst, we need
to review the fundamentals of inverse ltering.
2.2 INVERSE FILTERING
If a lter operator f(t) were dened such that convolu-
tion of f(t) with the known seismogram x(t) yields an
estimate of the earths impulse response e(t), then
e(t) = f(t) x(t). (2 4)
By substituting equation (2-4) into equation (2-3a), we
get
x(t) = w(t) f(t) x(t). (2 5)
When x(t) is eliminated from both sides of the equation,
the following expression results:
(t) = w(t) f(t), (2 6)
where (t) represents the Kronecker delta function:
(t) =
_
1, t = 0,
0, otherwise.
(2 7)
By solving equation (2-6) for the lter operator
f(t), we obtain
f(t) = (t)
1
w(t)
. (2 8)
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
172 Seismic Data Analysis
FIG. 2.2-1. A owchart for inverse ltering.
Therefore, the lter operator f(t) needed to compute
the earths impulse response from the recorded seismo-
gram turns out to be the mathematical inverse of the
seismic wavelet w(t). Equation (2-8) implies that the in-
verse lter converts the basic wavelet to a spike at t = 0.
Likewise, the inverse lter converts the seismogram to a
series of spikes that denes the earths impulse response.
Therefore, inverse ltering is a method of deconvolution,
provided the source waveform is known (deterministic
deconvolution). The procedure for inverse ltering is de-
scribed in Figure 2.2-1.
The Inverse of the Source Wavelet
Computation of the inverse of the source wavelet is ac-
complished mathematically by using the z-transform
(Section A.2). For example, let the basic wavelet be
a two-point time series given by w(t) : (1,
1
2
). The
z-transform of this wavelet is dened by the following
polynomial:
W(z) = 1
1
2
z. (2 9)
The power of variable z is the number of unit time de-
lays associated with each sample in the series. The rst
term has zero delay, so z is raised to zero power. The
second term has unit delay, so z is raised to rst power.
Hence, the z-transform of a time series is a polynomial
in z, whose coecients are the values of the time sam-
ples.
A relationship exists between the z-transform and
the Fourier transform (Section A.2). The z-variable is
dened as
z = exp
_
it
_
, (2 10)
where is angular frequency and t is sampling inter-
val.
The convolutional relation in the time domain
given by equation (2-8) means that the z-transform of
the inverse lter, F(z), is obtained by polynomial di-
vision of the z-transform of the input wavelet, W(z),
given by equation (2-9) (Section A.2):
F(z) =
1
1
1
2
z
= 1 +
1
2
z +
1
4
z
2
+ (2 11)
The coecients of F(z) : (1,
1
2
,
1
4
, . . .) represent the
time series associated with the lter operator f(t). Note
that the series has an innite number of coecients, al-
though they decay rapidly. As in any ltering process,
in practice the operator is truncated.
First consider the rst two terms in equation (2-
11) which yield a two-point lter operator (1,
1
2
). The
design and application of this operator is summarized in
Table 2-3. The actual output is (1, 0,
1
4
), whereas the
ideal result is a zero-delay spike (1, 0, 0). Although not
ideal, the actual result is spikier than the input wavelet,
(1,
1
2
).
Can the result be improved by including one more
coecient in the inverse lter? As shown in Table
2-4, the actual output from the three-point lter is
(1, 0, 0,
1
8
). This is a more accurate representation of
the desired output (1, 0, 0, 0) than that achieved with
the output from the two-point lter (Table 2-3). Note
that there is less energy leaking into the nonzero lags of
the output from the three-point lter. Therefore, it is
spikier. As more terms are included in the inverse lter,
the output is closer to being a spike at zero lag. Since
the number of points allowed in the operator length is
limited, in practice the result never is a perfect spike.
Table 2-3. Design and application of the truncated in-
verse lter (1,
1
2
) with the input wavelet (1,
1
2
).
Filter Design
Input Wavelet w(t) : (1,
1
2
)
The zTransform W(z) = 1
1
2
z
The Inverse F(z) = 1 +
1
2
z +
1
4
z
2
+
The Inverse Filter f(t) : (1,
1
2
,
1
4
, )
Filter Application
Truncated Inverse Filter (1,
1
2
)
Input Wavelet (1,
1
2
)
Actual Output (1, 0,
1
4
)
Desired Output (1, 0, 0)
Convolution Table:
1
1
2
Output
1
2
1 1
1
2
1 0
1
2
1
1
4
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
Deconvolution 173
Table 2-4. Design and application of the truncated in-
verse lter (1,
1
2
,
1
4
) with the input wavelet (1,
1
2
).
Filter Design
Input Wavelet w(t) : (1,
1
2
)
The zTransform W(z) = 1
1
2
z
The Inverse F(z) = 1 +
1
2
z +
1
4
z
2
+
The Inverse Filter f(t) : (1,
1
2
,
1
4
, )
Filter Application
Truncated Inverse Filter (1,
1
2
,
1
4
)
Input Wavelet (1,
1
2
)
Actual Output (1, 0, 0,
1
8
)
Desired Output (1, 0, 0, 0)
Convolution Table:
1
1
2
Output
1
4
1
2
1 1
1
4
1
2
1 0
1
4
1
2
1 0
1
4
1
2
1
1
8
The inverse of the input wavelet w(t) : (1,
1
2
)
has coecients that rapidly decay to zero (equation
2-11). What about the inverse of the input wavelet
w(t) : (
1
2
, 1)? Again, dene the ztransform:
W(z) =
1
2
+ z. (2 12)
The ztransform of its inverse is given by the polyno-
mial division:
F(z) =
1

1
2
+ z
= 2 4z 8z
2
(2 13)
As a result, the inverse lter coecients are given by
the divergent series f(t) : (2, 4, 8, ). Truncate
this series and convolve the two-point operator with the
input wavelet (
1
2
, 1) as shown in Table 2-5. The actual
output is (1, 0, 4), while the desired output is (1, 0, 0).
Not only is the result far from the desired output, but
also it is less spiky than the input wavelet (
1
2
, 1). The
reason for this poor result is that the inverse lter co-
ecients increase in time rather than decay (equation
2-13). When truncated, the larger coecients actually
are excluded from the computation.
If we kept the third coecient of the inverse lter
in the above example (equation 2-13), then the actual
output (Table 2-6) would be (1, 0, 0, 8), which also is
a bad approximation to the desired output (1, 0, 0, 0).
Table 2-5. Design and application of the truncated in-
verse lter (2, 4) with input wavelet (
1
2
, 1).
Filter Design
Input Wavelet w(t) : (
1
2
, 1)
The zTransform W(z) =
1
2
+ z
The Inverse F(z) = 2 4z 8z
2

The Inverse Filter f(t) : (2, 4, 8, )
Filter Application
Truncated Inverse Filter (2, 4)
Input Wavelet (
1
2
, 1)
Actual Output (1, 0, 4)
Desired Output (1, 0, 0)
Convolution Table:

1
2
1 Output
4 2 1
4 2 0
4 2 4
Table 2-6. Design and application of the truncated in-
verse lter (2, 4, 8) with input wavelet (
1
2
, 1).
Filter Design
Input Wavelet w(t) : (
1
2
, 1)
The zTransform W(z) =
1
2
+ z
The Inverse F(z) = 2 4z 8z
2

The Inverse Filter f(t) : (2, 4, 8, )
Filter Application
Truncated Inverse Filter (2, 4, 8)
Input Wavelet (
1
2
, 1)
Actual Output (1, 0, 0, 8)
Desired Output (1, 0, 0, 0)
Convolution Table:

1
2
1 Output
8 4 2 1
8 4 2 0
8 4 2 0
8 4 2 8
Least-Squares Inverse Filtering
A well-behaved input wavelet, such as (1,
1
2
) as op-
posed to (
1
2
, 1), has a z-transform whose inverse can
be represented by a convergent series. Then the inverse
ltering described above yields a good approximation to
a zero-lag spike output (1, 0, 0). Can we do even better
than that?
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
174 Seismic Data Analysis
Formulate the following problem: Given the input
wavelet (1,
1
2
), nd a two-term lter (a, b) such that
the error between the actual output and the desired
output (1, 0, 0) is minimum in the least-squares sense.
Compute the actual output by convolving the lter
(a, b) with the input wavelet (1,
1
2
) (Table 2-7). The
cumulative energy of the error L is dened as the sum
of the squares of the dierences between the coecients
of the actual and desired outputs:
L =
_
a 1
_
2
+
_
b
a
2
_
2
+
_

b
2
_
2
. (2 14)
The task is to nd coecients (a, b) so that L takes
its minimum value. This requires variation of L with
respect to the coecients (a, b) to vanish (Section
B.5). By simplifying equation (2-14), taking the par-
tial derivatives of quantity L with respect to a and b,
and setting the results to zero, we get
5
2
a b = 2, (2 15a)
and
a +
5
2
b = 0. (2 15b)
We have two equations and two unknowns; namely, the
lter coecients (a, b). The so-called normal set of equa-
tions (2-15a) and (2-15b) can be put into the following
convenient matrix form
_
5/2 1
1 5/2
__
a
b
_
=
_
2
0
_
. (2 16)
By solving for the lter coecients, we obtain (a, b) :
(0.95, 0.38). Design and application of this least-squares
inverse lter are summarized in Table 2-7.
To quantify the spikiness of this result and compare
it with the result from the inverse lter in Table 2-3,
compute the energy of the errors made in both (Table
2-8). Note that the least-squares lter yields less error
when trying to convert the input wavelet (1,
1
2
) to a
spike at zero lag (1, 0, 0).
We now examine the performance of the least-
squares lter with the input wavelet (
1
2
, 1). Note that
the inverse lter produced unstable results for this
wavelet (Table 2-5). We want to nd a two-term l-
ter (a, b) that, when convolved with the input wavelet
(
1
2
, 1), yields an estimate of the desired spike out-
put (1, 0, 0) (Table 2-9). As before, the least-squares
error between the actual output and the desired output
should be minimal.
The cumulative energy of the error is given by
L =
_

a
2
1
_
2
+
_

b
2
+ a
_
2
+ b
2
. (2 17)
Table 2-7. Design and application of a two-term least-
squares inverse lter (a, b).
Filter Design
Convolution of the lter (a, b) with input wavelet
(1,
1
2
):
1
1
2
Actual Output Desired Output
b a a 1
b a b a/2 0
b a b/2 0
Filter Application
Least-Squares Filter (0.95, 0.38)
Input Wavelet (1, 0.5)
Actual Output (0.95, 0.09, 0.19)
Desired Output (1, 0, 0)
Table 2-8. Error in two-term inverse and least-squares
ltering.
Input: (1,
1
2
)
Desired Output: (1, 0, 0)
Actual Output Error
Energy
Inverse Filter (1, 0, 0.25) 0.063
Least-Squares Filter (0.95, 0.09, 0.19) 0.048
Table 2-9. Design and application of a two-term least-
squares inverse lter (a, b).
Filter Design
Convolution of lter (a, b) with input wavelet (
1
2
, 1):
Actual Output Desired

1
2
1 Output
b a a/2 1
b a b/2 + a 0
b a b 0
Filter Application
Least-Squares Filter (0.95, 0.19)
Input Wavelet (0.5, 1)
Actual Output (0.24, 0.38, 0.19)
Desired Output (1, 0, 0)
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
Deconvolution 175
Table 2-10. Error in two-term inverse and least-
squares ltering.
Input: (
1
2
, 1)
Desired Output: (1, 0, 0)
Actual Output Error
Energy
Inverse Filter (1, 0, 4) 16
Least-Squares Filter (0.24, 0.38, 0.19) 0.762
By simplifying equation (2-17), taking the partial
derivatives of quantity L with respect to a and b, and
setting the results to zero, we obtain
5
2
a b = 1, (2 18a)
and
a +
5
2
b = 0. (2 18b)
Combine equations (2-18a,b) into a matrix form
_
5/2 1
1 5/2
__
a
b
_
=
_
1
0
_
. (2 19)
By solving for the lter coecients, we obtain (a, b) :
(0.95, 0.19). The design and application of this lter
are summarized in Table 2-9.
Table 2-10 shows the results from the inverse lter
and least-squares lter quantied. The error made by
the least-squares lter is, again, much less than the error
made by the truncated inverse lter. However, both l-
ters yield larger errors for input wavelet (
1
2
, 1) (Table
2-10) as compared to errors for wavelet (1,
1
2
) (Table
2-8). The reason for this is discussed next.
Minimum Phase
Two input wavelets, wavelet 1: (1,
1
2
) and wavelet 2:
(
1
2
, 1), were used for numerical analyses of the inverse
lter and least-squares inverse lter in this section. The
results indicate that the error in converting wavelet 1
to a zero-lag spike is less than the error in converting
wavelet 2 (Tables 2-8 and 2-10).
Is this also true when the desired ouput is a de-
layed spike (0, 1, 0)? The cumulative energy of the error
L associated with the application of a two-term least-
squares lter (a, b) (Table 2-11) to convert the input
wavelet (1,
1
2
) to a delayed spike (0, 1, 0) is
L = a
2
+
__
b
a
2
_
1
_
2
+
_

b
2
_
2
. (2 20)
Table 2-11. Design and application of a two-term least-
squares inverse lter (a, b).
Filter Design
Convolution of lter (a, b) with input wavelet (1,
1
2
):
Actual Output Desired
1
1
2
Output
b a a 0
b a b a/2 1
b a b/2 0
Filter Application
Least-Squares Filter (0.09, 0.76)
Input Wavelet (1,
1
2
)
Actual Output (0.09, 0.81, 0.38)
Desired Output (0, 1, 0)
Table 2-12. Error in least-squares ltering.
Input Wavelet: (1,
1
2
)
Desired Output Actual Output Error
Energy
(1, 0, 0) (0.95, 0.09, 0.19) 0.048
(0, 1, 0) (0.09, 0.81, 0.38) 0.190
By simplifying equation (2-20), taking the partial
derivatives of quantity L with respect to a and b, and
setting the results to zero, we obtain
5
2
a b = 1, (2 21a)
and
a +
5
2
b = 2. (2 21b)
Combine equations (2-21a,b) into a matrix form
_
5/2 1
1 5/2
__
a
b
_
=
_
1
2
_
. (2 22)
By solving for the lter coecients, we obtain (a, b)
: (0.09, 0.76). The design and application of this lter
are summarized in Table 2-11.
Table 2-12 shows the results of the least-sqaures
ltering to convert the input wavelet (1,
1
2
) to zero-
lag (Table 2-7) and delayed spikes (Table 2-11). Note
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
176 Seismic Data Analysis
Table 2-13. Design and application of a two-term least-
squares inverse lter (a, b).
Filter Design
Convolution of lter (a, b) with input wavelet (
1
2
, 1):
Actual Output Desired

1
2
1 Output
b a a/2 0
b a b/2 + a 1
b a b 0
Filter Application
Least-Squares Filter (0.76, 0.09)
Input Wavelet (0.5, 1)
Actual Output (0.38, 0.81, 0.09)
Desired Output (0, 1, 0)
that the input wavelet is converted to a zero-lag spike
with less error, and the corresponding actual output
more closely resembles a zero-lag spike desired output.
We now examine the performance of the least-
squares lter with the input wavelet(
1
2
, 1). The cu-
mulative energy of the error L associated with the ap-
plication of a two-term least-squares lter (a, b) (Table
2-13) to convert the input wavelet (
1
2
, 1) to a delayed
spike (0, 1, 0) is
L =
_

a
2
_
2
+
__

b
2
+ a
_
1
_
2
+ b
2
. (2 23)
By simplifying equation (2-23), taking the partial
derivatives of quantity L with respect to a and b, and
setting the results to zero, we obtain
5
2
a b = 2, (2 24a)
and
a +
5
2
b = 1. (2 24b)
Combine equations (2-24a,b) into a matrix form
_
5/2 1
1 5/2
__
a
b
_
=
_
2
1
_
. (2 25)
By solving for the lter coecients, we obtain (a, b)
: (0.76, 0.09). The design and application of this lter
are summarized in Table 2-13.
Table 2-14 shows the results of the least-squares
ltering to convert the input wavelet (
1
2
, 1) to zero-
lag (Table 2-9) and delayed spikes (Table 2-13). Note
that the input wavelet is converted to a delayed spike
with less error, and the corresponding actual output
more closely resembles a delayed spike desired output.
Table 2-14. Error in least-squares ltering.
Input Wavelet: (
1
2
, 1)
Desired Output Actual Output Error
Energy
(1, 0, 0) (0.24, 0.38, 0.19) 0.762
(0, 1, 0) (0.38, 0.81, 0.09) 0.190
Now, evaluate the results of the least-squares in-
verse ltering summarized in Tables 2-12 and 2-14.
Wavelet 1: (1,
1
2
) is closer to being a zero-delay spike
(1, 0, 0) than wavelet 2: (
1
2
, 1). On the other hand,
wavelet 2 is closer to being a delayed spike (0, 1, 0) than
wavelet 1. We conclude that the error is reduced if the
desired output closely resembles the energy distribution
in the input series. Wavelet 1 has more energy at the
onset, while wavelet 2 has more energy concentrated at
the end.
Figure 2.2-2 shows three wavelets with the same
amplitude spectrum, but with dierent phase-lag spec-
tra. As a result, their shapes dier. (From Section 1.1,
we know that the shape of a wavelet can be altered
by changing the phase spectrum without modifying the
amplitude spectrum.) The wavelet on top has more en-
ergy concentrated at the onset, the wavelet in the mid-
dle has its energy concentrated at the center, and the
wavelet at the bottom has most of its energy concen-
trated at the end.
We say that a wavelet is minimum phase if its en-
ergy is maximally concentrated at its onset. Similarly,
a wavelet is maximum phase if its energy is maximally
concentrated at its end. Finally, in all in-between situa-
tions, the wavelet is mixed phase. Note that a wavelet is
dened as a transient waveform with a nite duration
it is realizable. A minimum-phase wavelet is one-sided
it is zero before t = 0. A wavelet that is zero for
t < 0 is called causal. These denitions are consistent
with intuition physical systems respond to an excita-
tion only after that excitation. Their response also is of
nite duration. In summary, a minimum-phase wavelet
is realizable and causal.
These observations are quantied by consider-
ing the following four, three-point wavelets (Robinson,
1966):
Wavelet A : (4, 0, 1)
Wavelet B : (2, 3, 2)
Wavelet C : (2, 3, 2)
Wavelet D : (1, 0, 4)
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
Deconvolution 177
FIG. 2.2-2. A wavelet has a nite duration. If its energy
is maximally front-loaded, then it is minimum-phase (top).
If its energy is concentrated mostly in the middle, then it
is mixed-phase (middle). Finally, if its energy is maximally
end-loaded, then the wavelet is maximum-phase. A quanti-
tative analysis of this phase concept is provided in Figure
2.2-3.
Compute the cumulative energy of each wavelet
at any one time. Cumulative energy is computed by
adding squared amplitudes as shown in Table 2-15.
These values are plotted in Figure 2.2-3. Note that all
four wavelets have the same amount of total energy
17 units. However, the rate at which the energy builds
Table 2-15. Cumulative energy of wavelets A, B, C,
and D at time samples 0, 1 and 2.
Wavelet 0 1 2
A 16 16 17
B 4 13 17
C 4 13 17
D 1 1 17
FIG. 2.2-3. A quantitative analysis of the minimum- and
maximum-phase concept. The fastest rate of energy build-
up in time occurs when the wavelet is minimum-phase (A).
The slowest rate occurs when the wavelet is maximum-phase
(D).
up is signicantly dierent for each wavelet. For exam-
ple, with wavelet A, the energy builds up rapidly close
to its total value at the very rst time lag. The energy
for wavelets B and C builds up relatively slowly. Finally,
the energy accumulates at the slowest rate for wavelet
D. From Figure 2.2-3, note that the energy curves for
wavelets A and D form the upper and lower boundaries.
Wavelet A has the least energy delay, while wavelet D
has the largest energy delay.
FIG. 2.2-4. All wavelets referred to in Figure 2.2-3 (A, B,
C, and D ) have the same amplitude spectrum as shown
above (Adapted from Robinson, 1966).
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
178 Seismic Data Analysis
FIG. 2.2-5. Phase-lag spectra of the wavelets referred to in
Figure 2.2-3. They have the common amplitude spectrum of
Figure 2.2-4 (Adapted from Robinson, 1966).
Given a xed amplitude spectrum as in Figure 2.2-
4, the wavelet with the least energy delay is called min-
imum delay, while the wavelet with the most energy de-
lay is called maximum delay. This is the basis for Robin-
sons energy delay theorem: A minimum-phase wavelet
has the least energy delay.
Time delay is equivalent to a phase-lag. Figure 2.2-
5 shows the phase spectra of the four wavelets. Note
that wavelet A has the least phase change across the
frequency axis; we say it is minimum phase. Wavelet
D has the largest phase change; we say it is maximum
phase. Finally, wavelets B and C have phase changes
between the two extremes; hence, they are mixed phase.
Since all four wavelets have the same amplitude
spectrum (Figure 2.2-4) and the same power spectrum,
they should have the same autocorrelation. This is ver-
ied as shown in Table 2-16, where only one side of the
autocorrelation is tabulated, since a real time series has
a symmetric autocorrelation (Section 1.1).
Note that zero lag of the autocorrelation (Table 2-
16) is equal to the total energy (Table 2-15) contained
in each wavelet 17 units. This is true for any wavelet.
In fact, Parsevals theorem states that the area under
the power spectrum is equal to the zero-lag value of the
autocorrelation function (Section A.1).
The process by which the seismic wavelet is com-
pressed to a zero-lag spike is called spiking deconvolu-
tion. In this section, lters that achieve this goal were
studied the inverse and the least-squares inverse
lters. Their performance depends not only on lter
length, but also on whether the input wavelet is mini-
mum phase.
Table 2-16. Autocorrelation lags of wavelets A, B, C,
and D.
Wavelet A
4 0 1 Output
4 0 1 17
4 0 1 0
4 0 1 4
Wavelet B
2 3 2 Output
2 3 2 17
2 3 2 0
2 3 2 4
Wavelet C
2 3 2 Output
2 3 2 17
2 3 2 0
2 3 2 4
Wavelet D
1 0 4 Output
1 0 4 17
1 0 4 0
1 0 4 4
The spiking deconvolution operator is strictly the
inverse of the wavelet. If the wavelet were minimum
phase, then we would get a stable inverse, which also is
minimum phase. The term stable means that the lter
coecients form a convergent series. Specically, the
coecients decrease in time (and vanish at t = );
therefore, the lter has nite energy. This is the case
for the wavelet (1,
1
2
) with an inverse (1,
1
2
,
1
4
, . . .). The
inverse is a stable spiking deconvolution lter. On the
other hand, if the wavelet were maximum phase, then
it does not have a stable inverse. This is the case for the
wavelet (
1
2
, 1), whose inverse is given by the divergent
series (2, 4, 8, . . .). Finally, a mixed-phase wavelet
does not have a stable inverse. This discussion leads us
to assumption 7.
Assumption 7. The seismic wavelet is minimum
phase. Therefore, it has a minimum-phase inverse.
Now, a summary of the implications of the under-
lying assumptions for deconvolution stated in Sections
2.1 and 2.2 is appropriate.
(a) Assumptions 1, 2, and 3 allow formulating the con-
volutional model of the 1-D seismogram by equa-
tion (2-2).
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
Deconvolution 179
(b) Assumption 4 eliminates the unknown noise term
in equation (2-2a) and reduces it to equation (2-
3a).
(c) Assumption 5 is the basis for deterministic decon-
volution it allows estimation of the earths re-
ectivity series directly from the 1-D seismogram
described by equation (2-3a).
(d) Assumption 6 is the basis for statistical deconvolu-
tion it allows estimates for the autocorrelogram
and amplitude spectrum of the normally unknown
wavelet in equation (2-3a) from the known recorded
1-D seismogram.
(e) Finally, assumption 7 provides a minimum-phase
estimate of the phase spectrum of the seismic
wavelet from its amplitude spectrum, which is es-
timated from the recorded seismogram by way of
assumption 6.
Once the amplitude and phase spectra of the
seismic wavelet are statistically estimated from the
recorded seismogram, its least-squares inverse spik-
ing deconvolution operator, is computed using opti-
mum Wiener lters (Section 2.3). When applied to the
wavelet, the lter converts it to a zero-delay spike.
When applied to the seismogram, the lter yields the
earths impulse response (equation 2-4). In Section 2.3,
we show that a known wavelet can be converted into a
delayed spike even if it is not minimum phase.
2.3 OPTIMUM WIENER FILTERS
Return to the desired output the zero-delay spike
(1, 0, 0), that was considered when studying inverse and
least-squares lters (Section 2.2). Rewrite equation (2-
16), which we solved to obtain the least-squares inverse
lter, as follows:
2
_
5/4 1/2
1/2 5/4
__
a
b
_
=
_
2
0
_
. (2 26)
Divide both sides by 2 to obtain
_
5/4 1/2
1/2 5/4
__
a
b
_
=
_
1
0
_
. (2 27)
The autocorrelation of the input wavelet (1,
1
2
) is
shown in Table 2-17. Note that the autocorrelation lags
are the same as the rst column of the 2 2 matrix on
the left side of equation (2-27).
Now compute the crosscorrelation of the desired
output (1, 0, 0) with the input wavelet (1,
1
2
) (Table
2-18). The crosscorrelation lags are the same as the col-
umn matrix on the right side of equation (2-27).
Table 2-17. Autocorrelation lags of input wavelet
(1,
1
2
).
1
1
2
Output
1
1
2
5
4
1
1
2

1
2
Table 2-18. Crosscorrelation lags of desired output
(1, 0, 0) with input wavelet (1,
1
2
).
1 0 0 Output
1
1
2
1
1
1
2
0
In general, the elements of the matrix on the left
side of equation (2-27) are the lags of the autocorre-
lation of the input wavelet, while the elements of the
column matrix on the right side are the lags of the
crosscorrelation of the desired output with the input
wavelet.
Now perform similar operations for wavelet
(
1
2
, 1). By rewriting the matrix equation (2-19), we
obtain
2
_
5/4 1/2
1/2 5/4
__
a
b
_
=
_
1
0
_
. (2 28)
Divide both sides by 2 to obtain
_
5/4 1/2
1/2 5/4
__
a
b
_
=
_
1/2
0
_
. (2 29)
The autocorrelation of wavelet (
1
2
, 1) is given in
Table 2-19. The elements of the matrix on the left side of
equation (2-29) are the autocorrelation lags of the input
wavelet. Note that autocorrelation of wavelet (
1
2
, 1) is
identical to that of wavelet (1,
1
2
) (Table 2-17). As
discussed in Section 2.2, an important property of a
group of wavelets with the same amplitude spectrum is
that they also have the same autocorrelation.
The crosscorrelation of the desired output (1, 0, 0)
with input wavelet (
1
2
, 1) is given in Table 2-20. Note
that the right side of equation (2-29) is the same as the
crosscorrelation lags.
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
180 Seismic Data Analysis
Table 2-19. Autocorrelation lags of input wavelet
(
1
2
, 1).

1
2
1 Output

1
2
1
5
4

1
2
1
1
2
Table 2-20. Crosscorrelation lags of desired output
(1, 0, 0) with input wavelet (
1
2
, 1).
1 0 0 Output

1
2
1
1
2

1
2
1 0
Matrix equations (2-27) and (2-29) were used to de-
rive the least-squares inverse lters (Section 2.2). These
lters then were applied to the input wavelets to com-
press them to zero-lag spike. The matrices on the left
in equations (2-27) and (2-29) are made up of the au-
tocorrelation lags of the input wavelets. Additionally,
the column matrices on the right are made up of lags
of the crosscorrelation of the desired output a zero-
lag spike, with the input wavelets. These observations
were generalized by Wiener to derive lters that convert
the input to any desired output (Robinson and Treitel,
1980).
The general form of the matrix equation such as
equation (2-29) for a lter of length n is (Section B.5):
_
_
_
_
_
_
r
0
r
1
r
2
r
n1
r
1
r
0
r
1
r
n2
r
2
r
1
r
0
r
n3
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
r
n1
r
n2
r
n3
r
0
_
_
_
_
_
_
_
_
_
_
_
_
a
0
a
1
a
2
.
.
.
a
n1
_
_
_
_
_
_
=
_
_
_
_
_
_
g
0
g
1
g
2
.
.
.
g
n1
_
_
_
_
_
_
(2 30)
Here r
i
, a
i
, and g
i
, i = 0, 1, 2, . . . , n 1 are the auto-
correlation lags of the input wavelet, the Wiener lter
coecients, and the crosscorrelation lags of the desired
output with the input wavelet, respectively.
The optimum Wiener lter (a
0
, a
1
, a
2
, . . . , a
n1
) is
optimum in that the least-squares error between the ac-
tual and desired outputs is minimum. When the de-
sired output is the zero-lag spike (1, 0, 0, . . . , 0), then
the Wiener lter is identical to the least-squares inverse
lter. In other words, the least-squares inverse lter re-
ally is a special case of the Wiener lter.
The Wiener lter applies to a large class of prob-
lems in which any desired output can be considered,
not just the zero-lag spike. Five choices for the desired
output are:
Type 1: Zero-lag spike,
Type 2: Spike at arbitrary lag,
Type 3: Time-advanced form of input series,
Type 4: Zero-phase wavelet,
Type 5: Any desired arbitrary shape.
These desired output forms will be discussed in the fol-
lowing sections.
The general form of the normal equations (2-30)
was arrived at through numerical examples for the spe-
cial case where the desired output was a zero-lag spike.
Section B.5 provides a concise mathematical treatment
of the optimum Wiener lters. Figure 2.3-1 outlines the
design and application of a Wiener lter.
Determination of the Wiener lter coecients re-
quires solution of the so-called normal equations (2-
30). From equation (2-30), note that the autocorrela-
tion matrix is symmetric. This special matrix, called
the Toeplitz matrix, can be solved by Levinson recur-
sion, a computationally ecient scheme (Section B.6).
To do this, compute a two-point lter, derive from it
a three-point lter, and so on, until the n-point lter
is derived (Claerbout, 1976). In practice, ltering algo-
rithms based on the optimum Wiener lter theory are
known as Wiener-Levinson algorithms.
Spiking Deconvolution
The process with type 1 desired output (zero-lag
spike) is called spiking deconvolution. Crosscorrelation
of the desired spike (1, 0, 0, . . . , 0) with input wavelet
(x
0
, x
1
, x
2
, . . . , x
n1
) yields the series (x
0
, 0, 0, . . . , 0).
FIG. 2.3-1. A owchart for Wiener lter design and appli-
cation.
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
Deconvolution 181
The generalized form of the normal equations (2-30)
takes the special form:
_
_
_
_
_
_
r
0
r
1
r
2
r
n1
r
1
r
0
r
1
r
n2
r
2
r
1
r
0
r
n3
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
r
n1
r
n2
r
n3
r
0
_
_
_
_
_
_
_
_
_
_
_
_
a
0
a
1
a
2
.
.
.
a
n1
_
_
_
_
_
_
=
_
_
_
_
_
_
1
0
0
.
.
.
0
_
_
_
_
_
_
(2 31)
Equation (2-31) was scaled by (1/x
0
). The least-
squares inverse lter, which was discussed in Section 2.2,
has the same form as the matrix equation (2-31). There-
fore, spiking deconvolution is mathematically identical
to least-squares inverse ltering. A distinction, however,
is made in practice between the two types of ltering.
The autocorrelation matrix on the left side of equation
(2-31) is computed from the input seismogram (assump-
tion 6) in the case of spiking deconvolution (statistical
deconvolution), whereas it is computed directly from the
known source wavelet in the case of least-squares inverse
ltering (deterministic deconvolution).
Figure 2.3-2 is a summary of spiking deconvolution
based on the Wiener-Levinson algorithm. Frame (a) is
the input mixed-phase wavelet. Its amplitude spectrum
shown in frame (b) indicates that the wavelet has most
of its energy conned to a 10- to 50-Hz range. The au-
tocorrelation function shown in frame (d) is used in
equation (2-31) to compute the spiking deconvolution
operator shown in frame (e). The amplitude spectrum
of the operator shown in frame (f) is approximately the
inverse of the amplitude spectrum of the input wavelet
shown in frame (b). (The approximation improves as op-
erator length increases.) This should be expected, since
the goal of spiking deconvolution is to atten the out-
put spectrum. Application of this operator to the input
wavelet gives the result shown in frame (k).
Ideally, we would like to get a zero-lag spike, as
shown in frame (n). What went wrong? Assumption 7
was violated by the mixed-phase input wavelet shown in
frame (a). Frame (h) shows the inverse of the deconvo-
lution operator. This is the minimum-phase equivalent
of the input mixed-phase wavelet in frame (a). Both
wavelets have the same amplitude spectrum shown in
frames (b) and (i), but their phase spectra are signif-
icantly dierent as shown in frames (c) and (j). Since
spiking deconvolution is equivalent to least-squares in-
verse ltering, the minimum-phase equivalent is merely
the inverse of the deconvolution operator. Therefore, the
amplitude spectrum of the operator is the inverse of the
amplitude spectrum of the minimum-phase equivalent
as shown in frames (f) and (i), and the phase spectrum
of the operator is the negative of the phase spectrum
of the minimum-phase wavelet as shown in frames (g)
and (j). One way to extract the seismic wavelet, pro-
vided it is minimum phase, is to compute the spiking
deconvolution operator and nd its inverse.
In conclusion, if the input wavelet is not minimum
phase, then spiking deconvolution cannot convert it to a
perfect zero-lag spike as in frame (k). Although the am-
plitude spectrum is virtually at as shown in frame (l),
the phase spectrum of the output is not minimum phase
as shown in frame (m). Finally, note that the spiking
deconvolution operator is the inverse of the minimum-
phase equivalent of the input wavelet. This wavelet may
or may not be minimum phase.
Prewhitening
From the preceding section, we know that the ampli-
tude spectrum of the spiking deconvolution operator is
(approximately) the inverse of the amplitude spectrum
of the input wavelet. This is sketched in Figure 2.3-3.
What if we had zeroes in the amplitude spectrum of the
input wavelet? To study this, apply a minimum-phase
band-pass lter (Exercise 2-10) with a wide passband
(3-108 Hz) to the minimum-phase wavelet of Figure
2.3-2, as shown in frame (h). Deconvolution of the l-
tered wavelet does not produce a perfect spike; instead,
a spike accompanied by a high-frequency pre-and post-
cursor results (Figure 2.3-4). This poor result occurs be-
cause the deconvolution operator tries to boost the ab-
sent frequencies, as seen from the amplitude spectrum
of the output. Can this problem occur in a recorded seis-
mogram? Situations in which the input amplitude spec-
trum has zeroes rarely occur. There is always noise in
the seismogram and it is additive in both the time and
frequency domains. Moreover, numerical noise, which
also is additive in the frequency domain, is generated
during processing. However, to ensure numerical sta-
bility, an articial level of white noise is added to the
amplitude spectrum of the input seismogram before de-
convolution. This is called prewhitening and is referred
to in Figure 2.3-3.
If the percent prewhitening is given by a scalar, 0
< 1, then the normal equations (2-31) are modied as
follows:
_
_
_
_
_
_
r
0
r
1
r
2
r
n1
r
1
r
0
r
1
r
n2
r
2
r
1
r
0
r
n3
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
r
n1
r
n2
r
n3
r
0
_
_
_
_
_
_
_
_
_
_
_
_
a
0
a
1
a
2
.
.
.
a
n1
_
_
_
_
_
_
=
_
_
_
_
_
_
1
0
0
.
.
.
0
_
_
_
_
_
_
,
(2 32)
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
182 Seismic Data Analysis
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
Deconvolution 183
FIG. 2.3-3. Prewhitening amounts to adding a bias to the amplitude spectrum of the seismogram to be deconvolved. This
prevents dividing by zero since the amplitude spectrum of the inverse lter (middle) is the inverse of that of the seismogram
(left). Convolution of the lter with the seismogram is equivalent to multiplying their respective amplitude spectra this
yields nearly a white spectrum (right).
where = 1 +. Adding a constant r
0
to the zero lag
of the autocorrelation function is the same as adding
white noise to the spectrum, with its total energy equal
to that constant. The eect of the prewhitening level
on performance of deconvolution is discussed in Section
2.4.
Wavelet Processing by Shaping Filters
Spiking deconvolution had trouble compressing wavelet
(
1
2
, 1) to a zero-lag spike (1, 0, 0) (Table 2-14). In terms
of energy distribution, this input wavelet is more sim-
ilar to a delayed spike, such as (0, 1, 0), than it is to a
zero-lag spike, (1, 0, 0). Therefore, a lter that converts
wavelet (
1
2
, 1) to a delayed spike would yield less error
than the lter that shapes it to a zero-lag spike (Table
2-14).
Recast the lter design and application outlined
in Table 2-13 in terms of optimum Wiener lters by
following the owchart in Figure 2.3-1. First, compute
the crosscorrelation (Table 2-21). From Table 2-19, we
know the autocorrelation of the input wavelet. By sub-
stituting the results from Tables 2-19 and 2-21 into the
matrix equation (2-30), we get

5/4 1/2
1/2 5/4

a
b

1
1/2

. (2 33)
By solving for the lter coecients, we obtain (a, b) :
(
16
21
,
2
21
). This lter is applied to the input wavelet as
shown in Table 2-22. As we would expect, the output is
the same as that of the least-squares lter (Table 2-13).
Note that, from Table 2-14, the energy of the least-
squares error between the actual and desired outputs
was 0.190 and 0.762 for a delayed spike and a zero-
lag spike desired output, respectively. This shows that
there is less error when converting wavelet (
1
2
, 1) to the
delayed spike (0, 1, 0) than to zero-lag spike (1, 0, 0).
In general, for any given input wavelet, a series of
desired outputs can be dened as delayed spikes. The
least-squares errors then can be plotted as a function of
delay. The delay (lag) that corresponds to the least er-
ror is chosen to dene the desired delayed spike output.
The actual output from the Wiener lter using this opti-
mum delayed spike should be the most compact possible
result.
Table 2-21. Crosscorrelation lags of desired output
(0, 1, 0) with input wavelet (
1
2
, 1).
0 1 0 Output

1
2
1 1

1
2
1
1
2
Table 2-22. Convolution of input wavelet (
1
2
, 1) with
lter coecients (
16
21
,
2
21
).

1
2
1 Output

2
21

16
21
0.38

2
21

16
21
0.81

2
21

16
21
0.09
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
184 Seismic Data Analysis
FIG. 2.3-4. (a) Minimum-phase wavelet, (b) after band-pass ltering, (c) followed by deconvolution. The amplitude spectrum
of the band-pass ltered wavelet is zero above 108 Hz (middle row); therefore, the inverse lter derived from it yields unstable
results (bottom row). The time delays on the wavelets in the left frames of the middle and bottom rows are for display
purposes only.
The process that has a type 5 desired output (any
desired arbitrary shape) is called wavelet shaping. The
lter that does this is called a Wiener shaping lter.
In fact, type 2 (delayed spike) and type 4 (zero-phase
wavelet) desired outputs are special cases of the more
general wavelet shaping.
Figure 2.3-5 shows a series of wavelet shapings that
use delayed spikes as desired outputs. The input is a
mixed-phase wavelet. Filter length was held constant
in all eight cases. Note that the zero-delay spike case
(spiking deconvolution) does not always yield the best
result (Figure 2.3-5a). A delay in the neighborhood of
60 ms (Figure 2.3-5e) seems to yield an output that is
closest to being a perfect spike. Typically, the process
is not very sensitive to the amount of delay once it is
close to the optimum delay. If the input wavelet were
minimum-phase, then the optimum delay of the desired
output spike generally is zero. On the other hand, if the
input wavelet were mixed-phase, as illustrated in Fig-
ure 2.3-5, then the optimum delay is nonzero. Finally, if
the input wavelet were maximum-phase, then the opti-
mum delay is the length of that wavelet (Robinson and
Treitel, 1980).
Can we not delay the desired spike output (Fig-
ure 2.3-5) and obtain a better result than we obtained
from spiking deconvolution? This goal is achieved by
applying a constant-time shift (60 ms in Figure 2.3-5)
to a delayed spike result. Better yet, the same result
can be obtained by shifting the shaping lter operator
as much as the delay in the spike and applying it to the
input wavelet. Such a lter operator is two-sided (non-
causal), since it has coecients for negative and positive
time values. The one-sided lter dened along the posi-
tive time axis has an anticipation component, while the
lter dened along the negative time axis has a mem-
ory component (Robinson and Treitel, 1980). The two-
sided lter has an anticipation component and a mem-
ory component. Figure 2.3-6 shows a series of shaping
lterings with two-sided Wiener lters for various spike
delay values.
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
Deconvolution 185
FIG. 2.3-5. Shaping ltering. (0) Input wavelet, (1) desired
output, (2) shaping lter operator, (3) actual output. Here,
the purpose is to convert the mixed-phased wavelet (0) to
a series of delayed spikes as shown in (a) through (h) by
using a one-sided operator (anticipation component only).
The best result is with a 60-ms delay (e).
FIG. 2.3-6. Shaping ltering. (0) Input wavelet, (1) desired
output, (2) shaping lter operator, (3) actual output. Here,
the purpose is to convert the mixed-phase wavelet (0) to a
series of delayed spikes as shown in (a) through (h) using a
two-sided operator (with memory and anticipation compo-
nents). The best result is obtained with a zero-delay spike
using a two-sided lter (a).
Figure 2.3-7 shows examples of wavelet shaping.
The input wavelet represented by trace (b) is the same
mixed-phase wavelet as in Figure 2.3-6 (top left frame).
This wavelet is shaped into zero-phase wavelets with
three dierent bandwidths represented by traces (c),
(d) and (e). This process commonly is referred to as de-
phasing. Figure 2.3-7 shows another wavelet shaping in
which the input wavelet is converted to its minimum-
phase equivalent represented by trace (f). This conver-
sion is often applied to recorded air-gun signatures.
Figure 2.3-8 shows examples of a recorded air-
gun signature that was shaped into its minimum-phase
equivalent and into a spike. When the input is the
recorded signature, then the wavelet shapings in Fig-
ure 2.3-8 are called signature processing.
Wavelet shaping requires knowledge of the input
wavelet to compute the crosscorrelation column on the
right side of equation (2-30). If it is unknown, which is
the case in reality, then the minimum-phase equivalent
of the input wavelet can be estimated statistically from
the data. This minimum-phase estimate then is shaped
to a zero-phase wavelet.
Wavelet processing is a term that is used with ex-
ibility. The most common meaning refers to estimat-
ing (somehow) the basic wavelet embedded in the seis-
mogram, designing a shaping lter to convert the esti-
mated wavelet to a desired form, usually a broad-band
zero-phase wavelet (Figure 2.3-8), and nally, applying
the shaping lter to the seismogram. Another type of
wavelet processing involves wavelet shaping in which
the desired output is the zero-phase wavelet with the
same amplitude spectrum as that of the input wavelet
(Figure 2.3-9). Note that this type of wavelet processing
does not try to atten the spectrum, but only tries to
correct for the phase of the input wavelet, which some-
times is assumed to be minimum-phase.
Predictive Deconvolution
The type 3 desired output, a time-advanced form of the
input series, suggests a prediction process. Given the
input x(t), we want to predict its value at some future
time (t + ), where is prediction lag. Wiener showed
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
186 Seismic Data Analysis
FIG. 2.3-7. Shaping ltering with various desired outputs. (a) Impulse response, (b) input seismogram. Here, (c), (d) and
(e) show three possible desired outputs that are band-limited zero-phase wavelets, while (f) shows a desired output that is
the minimum-phase equivalent of the input wavelet (b). Finally, (g) and (h) are desired outputs that are band-pass ltered
versions of (f).
that the lter used to estimate x(t+) can be computed
by using a special form of the matrix equation (2-30)
(Robinson and Treitel, 1980). Since the desired output
x(t +) is the time-advanced version of the input x(t),
we need to specialize the right side of equation (2-30)
for the prediction problem.
Consider a ve-point input time series x(t) :
(x
0
, x
1
, x
2
, x
3
, x
4
), and set = 2. The autocorrelation
of the input series is computed in Table 2-23, and the
crosscorrelation between the desired output x(t+2) and
the input x(t) is computed in Table 2-24. Compare the
results in Tables 2-23 and 2-24, and note that g
i
= r
i+
for = 2 and i = 0, 1, 2, 3, 4.
Equation (2-30), for this special case, is rewritten
as follows:

r
0
r
1
r
2
r
3
r
4
r
1
r
0
r
1
r
2
r
3
r
2
r
1
r
0
r
1
r
2
r
3
r
2
r
1
r
0
r
1
r
4
r
3
r
2
r
1
r
0

a
0
a
1
a
2
a
3
a
4

r
2
r
3
r
4
r
5
r
6

. (2 34)
The prediction lter coecients a(t) : (a
0
, a
1
,
a
2
, a
3
, a
4
) can be computed from equation (2-34) and
applied to the input series x(t) : (x
0
, x
1
, x
2
, x
3
, x
4
) to
compute the actual output y(t) : (y
0
, y
1
, y
2
, y
3
, y
4
) (Ta-
ble 2-25). We want to predict the time-advanced form of
the input; hence, the actual output is an estimate of the
series x(t + ) : (x
2
, x
3
, x
4
), where = 2. The predic-
tion error series e(t) = x(t +) y(t) : (e
2
, e
3
, e
4
, e
5
, e
6
)
is given in Table 2-26.
The results in Table 2-26 suggest that the error se-
ries can be obtained more directly by convolving the
input series x(t) : (x
0
, x
1
, x
2
, x
3
, x
4
) with a lter with
coecients (1, 0, a
0
, a
1
, a
2
, a
3
, a
4
) (Table 2-27).
The results for (e
2
, e
3
, e
4
, e
5
, e
6
) are identical (Tables
2-26 and 2-27). Since the series (a
0
, a
1
, a
2
, a
3
, a
4
) is
called the prediction lter, it is natural to call the se-
ries (1, 0, a
0
, a
1
, a
2
, a
3
, a
4
) the prediction error
lter. When applied to the input series, this lter yields
the error series in the prediction process (Table 2-27).
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
Deconvolution 187
FIG. 2.3-8. Signature processing: (a) Recorded signature,
(b) desired output, (c) shaping operator, (d) shaped signa-
ture. The desired output is a zero-delay spike (top) and the
minimum-phase equivalent of the recorded signature (bot-
tom).
Table 2-23. Autocorrelation lags of input series x(t) :
(x
0
, x
1
, x
2
, x
3
, x
4
).
r
0
= x
2
0
+ x
2
1
+ x
2
2
+ x
2
3
+ x
2
4
r
1
= x
0
x
1
+ x
1
x
2
+ x
2
x
3
+ x
3
x
4
r
2
= x
0
x
2
+ x
1
x
3
+ x
2
x
4
r
3
= x
0
x
3
+ x
1
x
4
r
4
= x
0
x
4
r
5
= 0
r
6
= 0
Table 2-24. Crosscorrelation of desired output x(t +
) : (x
2
, x
3
, x
4
), = 2, with input x(t) : (x
0
, x
1
, x
2
,
x
3
, x
4
).
g
0
= x
0
x
2
+ x
1
x
3
+ x
2
x
4
g
1
= x
0
x
3
+ x
1
x
4
g
2
= x
0
x
4
g
3
= 0
g
4
= 0
Table 2-25. Convolution of prediction lter a(t) : (a
0
,
a
1
, a
2
, a
3
, a
4
) with input series x(t) : (x
0
, x
1
, x
2
, x
3
, x
4
)
to compute actual output y(t) : (y
0
, y
1
, y
2
, y
3
, y
4
).
y
0
= a
0
x
0
y
1
= a
1
x
0
+ a
0
x
1
y
2
= a
2
x
0
+ a
1
x
1
+ a
0
x
2
y
3
= a
3
x
0
+ a
2
x
1
+ a
1
x
2
+ a
0
x
3
y
4
= a
4
x
0
+ a
3
x
1
+ a
2
x
2
+ a
1
x
3
+ a
0
x
4
Table 2-26. The error series e(t) = x(t + ) y(t) :
(e
2
, e
3
, e
4
, e
5
, e
6
), = 2. For y(t), see Table 2-25.
e
2
= x
2
a
0
x
0
e
3
= x
3
a
1
x
0
a
0
x
1
e
4
= x
4
a
2
x
0
a
1
x
1
a
0
x
2
e
5
= 0 a
3
x
0
a
2
x
1
a
1
x
2
a
0
x
3
e
6
= 0 a
4
x
0
a
3
x
1
a
2
x
2
a
1
x
3
a
0
x
4
Table 2-27. Convolution of prediction error lter coef-
cients (1, 0, a
0
, a
1
, a
2
, a
3
, a
4
) with input series
x(t) : (x
0
, x
1
, x
2
, x
3
, x
4
).
e
0
= x
0
e
1
= x
1
e
2
= x
2
a
0
x
0
e
3
= x
3
a
1
x
0
a
0
x
1
e
4
= x
4
a
2
x
0
a
1
x
1
a
0
x
2
e
5
= 0 a
3
x
0
a
2
x
1
a
1
x
2
a
0
x
3
e
6
= 0 a
4
x
0
a
3
x
1
a
2
x
2
a
1
x
3
a
0
x
4
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
188 Seismic Data Analysis
FIG. 2.3-9. Wavelet processing. An autocorrelogram (a),
estimated from the seismic trace, is used after smoothing (b)
to compute the spiking deconvolution operator (d). Here (c)
is just a one-sided version of (b). The inverse of the operator
(d) is the minimum-phase wavelet (e), which is sometimes
assumed to be the basic wavelet contained in the original
seismic trace. It is easy to compute its zero-phase equiv-
alent (f) and design a shaping lter (g) that converts the
minimum-phase wavelet (e) to the zero-phase wavelet (f).
The actual output is (h), which should be compared with
(f). The zero-phase equivalent (f) has the same amplitude
spectrum as the minimum-phase wavelet (e).
Why place so much emphasis on the error series?
Consider the prediction process as it relates to a seis-
mic trace. From the past values of a time series up to
time t, a future value can be predicted at time t + ,
where is the prediction lag. A seismic trace often has a
predictable component (multiples) with a periodic rate
FIG. 2.3-10. A owchart for predictive deconvolution using
a prediction lter.
of occurrence. According to assumption 6, anything else,
such as primary reections, is unpredictable.
Some may claim that reections are predictable as
well; this may be the case if deposition is cyclic. How-
ever, this type of deposition is not often encountered.
While the prediction lter yields the predictable com-
ponent (the multiples) of a seismic trace, the remaining
unpredictable part, the error series, is essentially the
reection series.
Equation (2-34) can be generalized for the case of
an n-long prediction lter and an -long prediction lag.

r
0
r
1
r
2
r
n1
r
1
r
0
r
1
r
n2
r
2
r
1
r
0
r
n3
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
r
n1
r
n2
r
n3
r
0

a
0
a
1
a
2
.
.
.
a
n1

r
+1
r
+2
.
.
.
r
+n1

(2 35)
Note that design of the prediction lters requires only
autocorrelation of the input series.
There are two approaches to predictive deconvolu-
tion:
(1) The prediction lter (a
0
, a
1
, a
2
, . . . , a
n1
) may be
designed using equation (2-35) and applied on in-
put series as described in Figure 2.3-10.
(2) Alternatively, the prediction error lter (1, 0, 0, . . . ,
0, a
0
, a
1
, a
2
, . . . , a
n1
) can be designed and
convolved with the input series as described in Fig-
ure 2.3-11.
Now consider the special case of unit prediction lag,
= 1. For n = 5, equation (2-35) takes the following
form:
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
Deconvolution 189
FIG. 2.3-11. A owchart for predictive deconvolution using
a prediction error lter.

r
0
r
1
r
2
r
3
r
4
r
1
r
0
r
1
r
2
r
3
r
2
r
1
r
0
r
1
r
2
r
3
r
2
r
1
r
0
r
1
r
4
r
3
r
2
r
1
r
0

a
0
a
1
a
2
a
3
a
4

r
1
r
2
r
3
r
4
r
5

. (2 36)
By augmenting the right side to the left side, we obtain:

r
1
r
0
r
1
r
2
r
3
r
4
r
2
r
1
r
0
r
1
r
2
r
3
r
3
r
2
r
1
r
0
r
1
r
2
r
4
r
3
r
2
r
1
r
0
r
1
r
5
r
4
r
3
r
2
r
1
r
0

1
a
0
a
1
a
2
a
3
a
4

0
0
0
0
0
0

.
(2 37)
Add one row and move the negative sign to the column
matrix that represents the lter coecients to get:

r
0
r
1
r
2
r
3
r
4
r
5
r
1
r
0
r
1
r
2
r
3
r
4
r
2
r
1
r
0
r
1
r
2
r
3
r
3
r
2
r
1
r
0
r
1
r
2
r
4
r
3
r
2
r
1
r
0
r
1
r
5
r
4
r
3
r
2
r
1
r
0

1
a
0
a
1
a
2
a
3
a
4

L
0
0
0
0
0

.
(2 38)
where L = r
0
r
1
a
0
r
2
a
1
r
3
a
2
r
4
a
3
r
5
a
4
.
Note that there are six unknowns, (a
0
, a
1
, a
2
, a
4
, a
5
, L),
and six equations. Solution of these equations yields
the unit-delay prediction error lter (1, a
0
, a
1
,
a
2
, a
4
, a
5
), and the quantity L the error in the
ltering process (Section B.5). We can rewrite equation
(2-38) as follows:

r
0
r
1
r
2
r
3
r
4
r
5
r
1
r
0
r
1
r
2
r
3
r
4
r
2
r
1
r
0
r
1
r
2
r
3
r
3
r
2
r
1
r
0
r
1
r
2
r
4
r
3
r
2
r
1
r
0
r
1
r
5
r
4
r
3
r
2
r
1
r
0

b
0
b
1
b
2
b
3
b
4
b
5

L
0
0
0
0
0

.
(2 39)
where b
0
= 1, b
i
= a
i
, and i = 1, 2, 3, 4, 5. This equa-
tion has a familiar structure. In fact, except for the scale
factor L, it has the same form as equation (2-31), which
yields the coecients for the least-squares zero-delay in-
verse lter. This inverse lter is therefore the same as
the prediction error lter with unit prediction lag, ex-
cept for a scale factor. Hence, spiking deconvolution ac-
tually is a special case of predictive deconvolution with
unit prediction lag.
We now know that predictive deconvolution is a
general process that encompasses spiking deconvolu-
tion. In general, the following statement can be made:
Given an input wavelet of length (n+), the prediction
error lter contracts it to an -long wavelet, where
is the prediction lag (Peacock and Treitel, 1969). When
= 1, the procedure is called spiking deconvolution.
Figure 2.3-12 interrelates the various lters dis-
cussed in this chapter and indicates the kind of process
they imply. From Figure 2.3-12, note that Wiener l-
ters can be used to solve a wide range of problems. In
particular, predictive deconvolution is an integral part
of seismic data processing that is aimed at compressing
the seismic wavelet, thereby increasing temporal reso-
lution. In the limit, it can be used to spike the seismic
wavelet and obtain an estimate for reectivity.
2.3-12. A owchart for interrelations between various de-
convolution lters.
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
190 Seismic Data Analysis
2.4 PREDICTIVE DECONVOLUTION
IN PRACTICE
It now is appropriate to review the implications of the
assumptions stated in Sections 2.1 and 2.2 that under-
lie the process of deconvolution within the context of
predictive deconvolution.
(a) Assumptions 1, 2, and 3 are the basis for the convo-
lutional model of the recorded seismogram (Section
2.1). In practice, deconvolution often yields good
results in areas where these three assumptions are
not strictly valid.
(b) Assumption 3 can be relaxed in practice by consid-
ering a time-variant deconvolution (Section 2.6).
In this technique, a seismogram is divided into a
number of time gates, typically three or more. De-
convolution operators then are designed from each
gate and convolved with data within that gate. Al-
ternatively, time-variant spectral whitening can be
used to account for nonstationarity (Section 2.6).
(c) Not much can be done about assumption 4. How-
ever, noise can be minimized in the recording pro-
cess. Deconvolution operators can be designed us-
ing time gates and frequency bands with low noise
levels. Poststack deconvolution can be used in an
eort to take advantage of the noise reduction in-
herent in the stacking process.
(d) If the source wavelet were minimum-phase and
known (assumption 5), then a perfect result could
be obtained from deconvolution in the noise-free
case as in trace (c) of Figures 2.4-1 and 2.4-2.
(e) If assumption 6 were violated and if the source
waveform were not known, then you would have
problems as in trace (d) of Figures 2.4-1 and 2.4-2.
(f) The quality of the output from spiking deconvolu-
tion is degraded further when the source wavelet is
not minimum-phase as in Figures 2.4-3 and 2.4-4;
that is, when assumption 7 is violated.
(g) Finally, in addition to violating assumptions 5 and
7, if there were noise in the data, that is, when as-
sumption 4 is violated, then the result of the decon-
volution would be unacceptable as in Figure 2.4-5.
Figures 2.4-1 through 2.4-5 test our condence in
the usefulness of predictive deconvolution. In reality,
deconvolution has been applied to billions of seismic
traces; most of the time it has yielded satisfactory re-
sults. Figures 2.4-1 through 2.4-5 emphasize the criti-
cal assumptions that underlie predictive deconvolution.
When deconvolution does not work on some data, the
most probable reason is that one or more of the above
assumptions has been violated. In the remaining part
of this section, a series of numerical experiments will be
performed to examine the validity of these assumptions.
The purpose of these experiments is to gain a basic un-
derstanding of deconvolution from a practical point of
view.
Operator Length
We start with a single, isolated minimum-phase wavelet
as in trace (b) of Figure 2.4-6. Assumptions 1 through 5
are satised for this wavelet. The ideal result of spiking
deconvolution is a zero-lag spike, as indicated by trace
(a). In this and the following numerical analyses, we
refer to the autocorrelogram and amplitude spectrum
(plotted with linear scale) of the output from each de-
convolution test to better evaluate the results. In Figure
2.4-6 and the following gures, n, , and refer to oper-
ator length of the prediction lter, prediction lag, and
percent prewhitening, respectively. The length of the
prediction error lter then is n + .
In Figure 2.4-6, prediction lag is unity and equal
to the 2-ms sampling rate, prewhitening is 0%, and op-
erator length varies as indicated in the gure. Short
operators yield spikes with small-amplitude and rela-
tively high-frequency tails. The 128-ms-long operator
gives an almost perfect spike output. Longer operators
whiten the spectrum further, bringing it closer to the
spectrum of the impulse response.
The action of spiking deconvolution on the seismo-
gram derived by convolving the minimum-phase wavelet
with a sparse-spike series is similar (Figure 2.4-7) to the
case of the single isolated wavelet (Figure 2.4-6). Recall
that spiking deconvolution basically is inverse ltering
where the operator is the least-squares inverse of the
seismic wavelet. Therefore, an increasingly better result
should be obtained when more and more coecients are
included in the inverse lter.
Now consider the real situation of an unknown
source wavelet. Based on assumption 6, autocorrelation
of the input seismogram rather than that of the seismic
wavelet is used to design the deconvolution operator.
The result of using the trace rather than the wavelet
autocorrelation is shown in Figure 2.4-8. Deconvolution
recovers the gross aspects of the spike series, trace (a).
However, note that the deconvolved traces have spu-
rious small-amplitude spikes trailing each of the real
spikes. We see that increasing operator length does not
indenitely improve the results; on the contrary, more
and more spurious spikes are introduced.
Very short operators produce the same type of
noise spikes as in Figures 2.4-7 and 2.4-8. Examine the
series of deconvolution tests in Figure 2.4-8 and note
that the 94-ms operator does the best job. Compare
the autocorrelogram of trace (b) in Figure 2.4-8 with
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
Deconvolution 191
2.4-1. (a) Impulse response, (b) seismogram, (c) spiking deconvolution using known, minimum-phase wavelet, (d) deconvo-
lution assuming an unknown, minimum-phase source wavelet. Impulse response (a) is a sparse-spike series. For an unknown
source wavelet (in violation of assumption 4), spiking deconvolution yields a less than perfect result (compare (c) and (d)).
2.4-2. (a) Impulse response, (b) seismogram, (c) spiking deconvolution using known, minimum-phase source wavelet, (d)
deconvolution assuming an unknown, minimum-phase source wavelet. Impulse response (a) is based on a sonic log (Figure
2.1-1a). For the unknown source wavelet (in violation of assumption 4), spiking deconvolution yields a less than perfect result.
(Compare (c) and (d).
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
192 Seismic Data Analysis
2.4-3. (a) Impulse response, (b) seismogram, (c) deconvolution using a known, mixed-phase source wavelet, (d) deconvolution
assuming an unknown, mixed-phase source wavelet. Impulse response (a) is a sparse-spike series. For a mixed-phase source
wavelet (in violation of assumption 5), spiking deconvolution yields a degraded output (d), even when the wavelet is known
(c).
2.4-4. (a) Impulse response, (b) seismogram, (c) deconvolution using a known, mixed-phase source wavelet, (d) deconvolution
assuming an unknown, mixed-phase source. Impulse response (a) is based on the sonic log of Figure 2.1-1a. For the mixed-
phase source wavelet (in violation of assumption 5), spiking deconvolution yields a degraded output (d) even when the wavelet
is known (c).
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
Deconvolution 193
2.4-5. (a) Impulse response, (b) seismogram with noise, (c) deconvolution assuming an unknown, mixed-phase source wavelet.
Impulse response (a) is based on the sonic log of Figure 2.1-1a. In the presence of random noise (in violation of assumption
3), spiking deconvolution can produce a result with nelation to the earths reectivity (compare (a) to (c)).
that of trace (b) in Figure 2.4-6. Note that only the
rst 100-ms portion represents the autocorrelation of
the source wavelet. This explains why the 94-ms oper-
ator worked best; that is, the autocorrelation lags of
trace (b) in Figure 2.4-8 beyond 94 ms do not represent
the seismic wavelet.
Consider the seismogram in Figure 2.4-9, where the
wavelet is assumed to be unknown. Deconvolution has
restored the spikes that correspond to major reections
in the impulse response as in trace (b) with some suc-
cess. The 64-ms operator is a good choice.
The mixed-phase wavelet in Figure 2.4-10 shows
what can happen when assumption 7 is violated. The
wavelet in Figure 2.4-6 is the minimum-phase equiva-
lent of the mixed-phase wavelet in Figure 2.4-10. Both
wavelets have the same autocorrelograms and ampli-
tude spectra. Hence, the deconvolution operators for
both wavelets are identical. Because the minimum-
phase assumption was violated, deconvolution does not
convert the mixed-phase wavelet to a perfect spike. In-
stead, the deconvolved output is a complicated high-
frequency wavelet. Also note that the dominant peak in
the output is negative, while the impulse response has
a positive spike. This dierence in the sign can happen
when a mixed-phase wavelet is deconvolved. Increasing
the operator length further whitens the spectrum; how-
ever, the 128-ms operator yields a result that cannot be
improved further by longer operators.
The seismogram obtained from the mixed-phase
wavelet and the sparse-spike series (used in the pre-
ceding gures) is shown in Figure 2.4-11. The 94-ms
operator gives the best result. This also is the case in
Figure 2.4-12, where both assumptions 6 and 7 are vi-
olated. The situation with the seismogram in Figure
2.4-13 is not very good. The spikes that correspond to
major reections in the impulse response were restored;
however, there are some timing errors and polarity re-
versals. (Compare these results with those in Figure 2.4-
9 for the events between 0.2 and 0.3 s and 0.6 and 0.7
s.) The 64-ms operator gives an output that cannot be
improved by longer operators.
What kind of operator length should be used for
spiking deconvolution? To select an operator length,
ideally we want to use the autocorrelation of the un-
known seismic wavelet. Fortunately, the autocorrelation
of the input seismogram has the characteristics of the
wavelet autocorrelation (assumption 6). Therefore, it
seems appropriate that we should use part of the au-
tocorrelation obtained from the input seismogram that
most resembles the autocorrelation of the unknown seis-
mic wavelet. That part is the rst transient zone in the
autocorrelation, as seen by comparing the autocorrela-
tions of trace (b) in Figure 2.4-6 and trace (c) in Figure
2.4-9. The autocorrelations of trace (b) in Figure 2.4-10
and trace (c) in Figure 2.4-13 suggest the same princi-
ple.
Prediction Lag
So far, we have learned that predictive deconvolution
has two uses: (a) spiking deconvolution the case
of unit prediction lag, and (b) predicting the input
(text continues on p. 198)
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
194 Seismic Data Analysis
2.4-6. Test of operator length for a single, isolated input wavelet, where n = operator length, = prediction lag, and =
percent prewhitening. (a) Impulse response, (b) seismogram with minimum-phase source wavelet.
2.4-7. Test of operator length where n = operator length, = prediction lag, and = percent prewhitening. (a) Impulse
response, (b) seismogram with known, minimum-phase source wavelet.
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
Deconvolution 195
2.4-8. Test of operator length where n = operator length, = prediction lag, and = percent prewhitening. (a) Impulse
response, (b) seismogram with known, minimum-phase source wavelet.
2.4-9. Test of operator length where n = operator length, = prediction lag, and = percent prewhitening. (a) Reectivity,
(b) impulse response, (c) seismogram with unknown, minimum-phase source wavelet.
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
196 Seismic Data Analysis
2.4-10. Test of operator length for a single, isolated input wavelet where n= operator length, = prediction lag, and =
percent prewhitening. (a) Impulse response, (b) seismogram with mixed-phase source wavelet.
2.4-11. Test of operator length where n= operator length, = prediction lag, and = percent prewhitening. (a) Impulse
response, (b) seismogram with known, mixed-phase source wavelet.
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
Deconvolution 197
2.4-12. Test of operator length where n= operator length, = prediction lag, and = percent prewhitening. (a) Impulse
response, (b) seismogram with unknown, mixed-phase source wavelet.
2.4-13. Test of operator length where n= operator length, = prediction lag, and = percent prewhitening. (a) Reectivity,
(b) impulse response, (c) seismogram with unknown, mixed-phase source wavelet.
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
198 Seismic Data Analysis
seismogram at a future time dened by the prediction
lag. Case (b) is used to predict and attenuate multiples.
Now, the eect of the prediction lag parameter is
examined from an interpretive point of view. Consider
the single, isolated minimum-phase wavelet in Figure
2.4-14. Here, operator length and percent prewhiten-
ing are kept constant, while prediction lag is varied.
When prediction lag is equal to the sampling rate, then
the result is equivalent to spiking deconvolution. Predic-
tive deconvolution using a prediction lag greater than
unity yields a wavelet of nite duration instead of a
spike. Given an input wavelet of +n samples, predic-
tive deconvolution using prediction lter with length n
and prediction lag converts this wavelet into another
wavelet that is samples long. The rst lags of the
autocorrelation are preserved, while the next n lags are
zeroed out. Additionally, the amplitude spectrum of the
output increasingly resembles that of the input wavelet
as prediction lag is increased (Figure 2.4-14). At a 94-
ms prediction lag, predictive deconvolution does noth-
ing to the input wavelet because almost all the lags of
its autocorrelation have been left untouched. This ex-
periment has an important practical implication: Under
the ideal, noise-free conditions, resolution on the output
from predictive deconvolution can be controlled by ad-
justing the prediction lag. Unit prediction lag implies
the highest resolution, while a larger prediction lag im-
plies less than full resolution. However, in reality, these
assessments are dictated by the signal-to-noise ratio.
The deconvolved output using a unit prediction lag con-
tains high frequencies; nevertheless, resolution may be
degraded if the high-frequency energy is mostly noise,
not signal.
In Figure 2.4-14, prediction lags of 8 and 22 ms
correspond to the rst and second zero crossings on au-
tocorrelation of the input wavelet, respectively. The rst
zero crossing produces a spike with some width, while
the second zero crossing lag produces a wavelet with a
positive and negative lobe.
The relationship between prediction lag and
whitening also holds for the sparse-spike series in Figure
2.4-15 and when the input wavelet is unknown (Figure
2.4-16).
The eect of prediction lag on the output from
predictive deconvolution of a synthetic seismogram,
which was obtained from the sonic log (Figure 2.1-1a),
is demonstrated in Figures 2.4-17 and 2.4-18. As the
prediction lag is increased, the output spectrum be-
comes increasingly less broadband. Predictive decon-
volution of seismograms constructed from the mixed-
phase wavelet again demonstrates that output resolu-
tion can be controlled by adjusting prediction lag (Fig-
ures 2.4-19 through 2.4-23).
If prediction lag is increased, then the output am-
plitude spectrum becomes increasingly band-limited.
The output also can be band-limited by applying a
band-pass lter on the spiking deconvolution output.
(text continues on p. 203)
2.4-14. Test of prediction lag for a single, isolated input wavelet where n= operator length, = prediction lag, and =
percent prewhitening. (a) Impulse response, (b) seismogram with minimum-phase source wavelet.
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
Deconvolution 199
2.4-15. Test of prediction lag where n= operator length, = prediction lag, and = percent prewhitening. (a) Impulse
response, (b) seismogram with known, minimum-phase source wavelet.
2.4-16. Test of prediction lag where n= operator length, = prediction lag, and = percent prewhitening. (a) Impulse
response, (b) seismogram with unknown, minimum-phase source wavelet.
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
200 Seismic Data Analysis
2.4-17. Test of prediction lag where n= operator length, = prediction lag, and = percent prewhitening. (a) Reectivity,
(b) impulse response, (c) seismogram with known minimum-phase source wavelet.
2.4-18. Test of prediction lag where n= operator length, = prediction lag, and = percent prewhitening. (a) Reectivity,
(b) impulse response, (c) seismogram with unknown, minimum-phase source wavelet.
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
Deconvolution 201
2.4-19. Test of prediction lag for a single, isolated input wavelet where n = operator length, = prediction lag, and =
percent prewhitening. (a) Impulse response, (b) seismogram with mixed-phase source wavelet.
2.4-20. Test of prediction lag where n = operator length, = prediction lag, and = percent prewhitening. (a) Impulse
response, (b) seismogram with known, mixed-phase source wavelet.
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
202 Seismic Data Analysis
2.4-21. Test of prediction lag where n = operator length, = prediction lag, and = percent prewhitening. (a) Impulse
response, (b) seismogram with unknown, mixed-phase source wavelet.
2.4-22. Test of prediction lag where n = operator length, = prediction lag, and = percent prewhitening. (a) Reectivity,
(b) impulse response, (c) seismogram with known, mixed-phase source wavelet.
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
Deconvolution 203
2.4-23. Test of prediction lag where n = operator length, = prediction lag, and = percent prewhitening. (a) Reectivity,
(b) impulse response, (c) seismogram with unknown, mixed-phase source wavelet.
Are these two ways of band-limiting equivalent? Refer
to the results from both the minimum- and mixed-phase
wavelets in Figures 2.4-14 and 2.4-19, respectively. Note
that the output of the 22-ms prediction lag has an am-
plitude spectrum that is band-limited to approximately
0 to 100 Hz. However, the spectral shape within this
bandwidth is not a boxcar, but rather similar to that of
the input wavelet. The boxcar shape would be the case if
a band-pass lter (0 to 100 Hz) were applied to the out-
put of the spiking deconvolution (2-ms prediction lag).
Hence, spiking deconvolution followed by band-pass l-
tering is not equivalent to predictive deconvolution with
a prediction lag greater than unity.
In conclusion, if prediction lag is increased, the out-
put from predictive deconvolution becomes less spiky.
This eect can be used to our advantage, since it allows
the bandwidth of deconvolved output to be controlled
by adjusting the prediction lag. The application of spik-
ing deconvolution to eld data is not always desirable,
since it boosts high-frequency noise in the data. The
most prominent eect of the nonunity prediction lag is
suppression of the high-frequency end of the spectrum
and preservation of the overall spectral shape of the
input data. This eect is seen in Figures 2.4-18 and 2.4-
23, which correspond to the minimum-and mixed-phase
seismic wavelets. If prediction lag is increased further,
then the low-frequency end of the spectrum is aected
as well, making the output more band-limited.
Percent Prewhitening
The reasons for prewhitening were discussed in Sec-
tion 2.3. Consider the single, isolated minimum-phase
wavelet in Figure 2.4-24. Keep the operator length and
prediction lag constant and vary the percent prewhiten-
ing. Note that the eect of varying prewhitening is
similar to that of varying the prediction lag; that is,
the spectrum increasingly becomes less broadband as
the percent prewhitening is increased. Compare Figure
2.4-14 with Figure 2.4-24. Note that prewhitening nar-
rows the spectrum without changing much of the at-
ness character, while larger prediction lag narrows the
spectrum and alters its shape, making it look more like
the spectrum of the input seismic wavelet. These char-
acteristics also can be inferred from the shapes of the
output wavelets. Prewhitening preserves the spiky char-
acter of the output, although it adds a low-amplitude,
high-frequency tail (Figure 2.4-24). On the other hand,
increasing prediction lag produces a wavelet with a du-
ration equal to the prediction lag (Figure 2.4-14).
The eect of prewhitening on the sparse-spike train
seismogram with a known and unknown minimum-
phase wavelet is shown in Figures 2.4-25 and 2.4-26,
respectively. The eect of prewhitening on deconvo-
lution of the synthetic seismogram obtained from the
sonic log (Figure 2.1-1a) is shown in Figures 2.4-27
and 2.4-28 for known and unknown minimum-phase
wavelets. Prewhitening tests using the mixed-phase
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
204 Seismic Data Analysis
2.4-24. Test of percent prewhitening for a single, isolated input wavelet where n = operator length, = prediction lag, and
= percent prewhitening. (a) Impulse response, (b) seismogram with minimum-phase source wavelet.
2.4-25. Test of percent prewhitening where n = operator length, = prediction lag, and = percent prewhitening. (a)
Impulse response, (b) seismogram with known, minimum-phase source wavelet.
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
Deconvolution 205
2.4-26. Test of percent prewhitening where n = operator length, = prediction lag, and = percent prewhitening. (a)
Impulse response, (b) seismogram with unknown, minimum-phase source wavelet.
2.4-27. Test of percent prewhitening where n = operator length, = prediction lag, and = percent prewhitening. (a)
Reectivity, (b) impulse response, (c) seismogram with known, minimum-phase source wavelet.
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
206 Seismic Data Analysis
2.4-28. Test of percent prewhitening where n = operator length, = prediction lag, and = percent prewhitening. (a)
Reectivity, (b) impulse response, (c) seismogram with unknown, minimum-phase source wavelet.
2.4-29. Test of percent prewhitening for a single, isolated input wavelet where n = operator length, = prediction lag, and
= percent prewhitening. (a) Impulse response, (b) seismogram with mixed-phase source wavelet.
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
Deconvolution 207
2.4-30. Test of percent prewhitening for a single, isolated input wavelet where n = operator length, = prediction lag, and
= percent prewhitening. (a) Impulse response, (b) seismogram with minimum-phase source wavelet.
wavelet are shown in Figure 2.4-29. Finally, the com-
bined eects of a prediction lag that is greater than
unity and prewhitening for the single, isolated wavelet
are shown in Figure 2.4-30. These gures demonstrate
that prewhitening narrows the output spectrum, mak-
ing it band-limited. In particular, the tests in Figures
2.4-24 and 2.4-29 using the single, isolated minimum-
and mixed-phase wavelets suggest that spiking decon-
volution with some prewhitening is somewhat equiva-
lent to spiking deconvolution without prewhitening fol-
lowed by post-deconvolution broad band-pass ltering.
However, this is not exactly true, for prewhitening still
leaves some relatively suppressed energy at the high-
frequency end of the spectrum. From Figure 2.4-30, we
infer that predictive deconvolution with a prediction lag
greater than unity and with some prewhitening yields a
result somewhat equivalent to a spiking deconvolution
followed by band-pass ltering.
In conclusion, we can say that prewhitening yields
a band-limited output. However, the eect is less con-
trollable when compared to varying the prediction lag.
By varying prediction lag, we have some idea of the out-
put bandwidth, since it is related to prediction lag. The
smaller the prediction lag, the broader the output band-
width. Prewhitening is used only to ensure that numer-
ical instability in solving for the deconvolution operator
(equation 2-32) is avoided. In practice, typically 0.1 to
1% prewhitening is standard.
Eect of Random Noise on Deconvolution
We assume that the noise component in the recorded
seismogram is zero (assumption 4). The autocorrelation
of ideal random noise is zero at all lags except the zero
lag (Figure 2.1-5). Therefore, the eect of random noise
on deconvolution operators should be somewhat simi-
lar to the eect of prewhitening. Both eects modify the
diagonal of the autocorrelation matrix, making it more
dominant [equation (2-32)]. However, the noise compo-
nent also slightly modies the nonzero lags of the au-
tocorrelation. Compare the autocorrelograms of traces
(b) in Figures 2.4-24 and 2.4-31. In Figure 2.4-24, an
isolated minimum-phase wavelet was considered, while
in Figure 2.4-31, random noise was added to the same
wavelet. The output wavelet shape from spiking decon-
volution of the noisy wavelet using a 128-ms operator is
similar to the output from spiking deconvolution of the
wavelet without noise, using the same operator length
but with, say, 20 percent prewhitening. This result has
practical importance prewhitening is equivalent to
adding perfect random noise to the system. Since a
recorded seismogram always contains some amount of
random noise, only a minute amount, say 0.1 percent,
of the white noise needs to be added to the seismogram
for numerical stability.
The eect of random noise on the performance
of deconvolution is examined further in Figures 2.4-
32 and 2.4-33. These results should be compared with
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
208 Seismic Data Analysis
2.4-31. Eect of random noise on deconvolution performance. Input seismogram (b) associated with reectivity (a) contains
a single, isolated wavelet (at around 0.2 s) buried in random noise. Here, n = operator length, = prediction lag, and =
percent prewhitening.
2.4-32. Eect of random noise on deconvolution performance. (a) Reectivity, (b) impulse response. Input seismogram (c)
with unknown, minimum-phase source wavelet is noise contaminated. Here, n = operator length, = prediction lag, and
= percent prewhitening.
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
Deconvolution 209
2.4-33. Eect of random noise on deconvolution performance. (a) Reectivity, (b) impulse response. Input seismogram (c)
with unknown, mixed-phase source wavelet is noise contaminated. Here, n = operator length, = prediction lag, and =
percent prewhitening.
their noiseless counterparts in Figures 2.4-9 and 2.4-
13, respectively. Observe that the noise component has
a harmful eect on deconvolution. For example, when
comparing Figures 2.4-9 and 2.4-32, note that the de-
convolution result from the noisy seismogram has spu-
rious spikes (for instance, between 0.5 and 0.6 s), which
could be interpreted as genuine reections. Noisy eld
data, which yield better stack when not treated by de-
convolution, have been noted. Only by testing can we
determine whether deconvolution performs satisfacto-
rily on data with a severe noise problem.
Multiple Attenuation
We have learned that a prediction lter predicts peri-
odic events, like multiples, in the seismogram. The pre-
diction error lter yields the unpredictable component
of the seismogram the reectivity series. For exam-
ple, consider the simple case of water-bottom multiples.
If the reection coecient of the water bottom is c
w
and
if water depth is equivalent to a two-way time t
w
, then
the time series is

1, 0, . . . , 0, c
w
, 0, . . . , 0, c
2
w
, 0, . . . , 0, c
3
w
, 0, . . .)
as represented by trace (b) in Figure 2.4-34. The sepa-
ration between the spikes is t
w
in trace (b). Note that
the periodicity in the time series (trace (b) or (c)) man-
ifests itself in the amplitude spectrum as periodic peaks
(or notches). The greater the spike separation in time,
the closer the peaks (or notches) in the amplitude spec-
trum.
The noise-free convolutional model for the seismo-
gram that contains the water-bottom multiples can be
written as
x(t) = w(t) m(t) e(t), (2 40)
where m(t) represents the water-layer reverberation
spike series as in trace (b) of Figure 2.4-34, and e(t) now
represents the earths impulse response excluding mul-
tiples associated with the water bottom. Predictive de-
convolution can suppress the periodic component m(t)
in the seismogram as demonstrated by trace (d) in Fig-
ure 2.4-3.
Note the two distinct goals for predictive deconvo-
lution: (a) spiking the seismic wavelet w(t), and (b) pre-
dicting and attenuating multiples m(t). The rst goal
is achieved using an operator with unit prediction lag,
while the second is achieved using an operator with a
prediction lag greater than unity.
The autocorrelation of the input trace can be used
to determine the appropriate prediction lag for multiple
suppression. Periodicity associated with multiples is ev-
ident in the autocorrelogram of trace (c) in Figure 2.4-
34, as an isolated series of energy lobes in the neighbor-
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
210 Seismic Data Analysis
2.4-34. (a) Reectivity, (b) impulse response, (c) seismogram. Two-step deconvolution aimed at attenuating the multiples,
then spiking the remaining primary wavelet (d) to (e). The process can be performed in the reverse order (f) to (g). Here, n
= operator length and = prediction lag.
2.4-35. Predictive deconvolution for multiple attenuation. (a) Reectivity, (b) impulse response, (c) seismogram. Two-step
deconvolution: Predictive deconvolution (d) followed by spiking deconvolution (e). Traces (f), (g), and (h) result when single-
step deconvolution is applied to the input trace (c), using the operator lengths n and prediction lags , as indicated.
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
Deconvolution 211
hood of 0.2 and 0.4 s. Prediction lag should be chosen
to bypass the rst part of the autocorrelogram that rep-
resents the seismic wavelet. Operator length should be
chosen to include the rst isolated energy packet in the
autocorrelogram. After applying predictive deconvolu-
tion, we are left with only the water-bottom primary
reection. Isolated bursts in the autocorrelogram have
been suppressed, while periodic peaks in the amplitude
spectrum have been eliminated as shown in trace (d). If
desired, the basic wavelet can be compressed into a spike
as shown in trace (e) by applying spiking deconvolution
to the output of predictive deconvolution as shown in
trace (d). The sequence can be interchanged by rst
applying spiking deconvolution as shown in trace (f)
followed by predictive deconvolution as shown in trace
(g).
By using a suciently long spiking deconvolution
operator, two goals are achieved in one step as seen in
trace (h). However, this approach can be dangerous if
primary reections are unintentionally suppressed. This
is the case in Figure 2.4-35. Here, the water-bottom
reection is followed by a deeper event at about 0.28
s as seen in trace (a). The impulse response contains
water-bottom multiples and the peg-leg multiples that
are associated with the deeper reector as seen in trace
(b). The amplitude spectrum has peaks that come in
pairs, indicating the presence of two dierent periodic
components in the seismogram. Careful choice of pre-
dictive deconvolution parameters yields an output with
only the wavelets associated with the water bottom and
the deeper reector as seen in trace (d). This is fol-
lowed by a spiking deconvolution that yields two spikes
representative of the water bottom and the deep pri-
mary as seen in trace (e). Spiking deconvolution alone
produces the reection coecient series and the spikes
that represent the multiples as seen in trace (f). If a
longer spiking deconvolution operator is used, then the
primary reection easily can be eliminated as in trace
(g). If a predictive deconvolution operator is used with
an improper parameter choice, then again, the primary
reection can be eliminated easily as in trace (h).
How can we ensure that no primaries are de-
stroyed by deconvolution? Examine the autocorrelo-
gram of trace (c) in Figure 2.4-35. The rst 50-ms por-
tion represents the seismic wavelet. This is followed by
a burst between 50 to 170 ms that represents the cor-
relation of the water bottom and primary. The isolated
burst between 170 to 340 ms represents the actual mul-
tiple series (both the peg-legs and water-bottom mul-
tiples). The prediction lag must be chosen to bypass
the rst part of the autocorrelogram, which represents
the seismic wavelet and possible correlation between
the primaries. The operator length must be chosen to
include the rst isolated burst, in this case between 170
to 340 ms.
It is only with vertical incidence and zero-oset
recording that periodicity of the multiples is preserved.
Therefore, predictive deconvolution aimed at multi-
ple suppression may not be entirely eective when
applied to nonzero-oset data, such as common-shot
or common-midpoint data. Figure 2.4-36a shows a
common-shot gather with its autocorrelogram and av-
erage amplitude spectrum. The eld record is prepared
for deconvolution by rst applying t
2
scaling (Figure
2.4-36b) and muting the rst arrivals associated with
largely guided waves (Figure 2.4-36c). The autocorrelo-
gram in Figure 2.4-36c indicates the presence of multi-
ples. Note on the shot record the water-bottom reec-
tion is at 0.4 s at near oset. Additionally, there are
two strong primary reectors at 0.6 and 1.4 s at near
oset; these primaries give rise to a rst-order and peg-
leg multiples. Despite the atness of the spectrum and
the attenuation of nonzero lags of the autocorrelogram
after deconvolution, the rst-order multiples associated
with the water-bottom reection and the peg-leg mul-
tiples associated with the primary reection at 0.6 s
still persist in the record (Figures 2.4-36d). This occurs
because these events have large moveout which causes
signicant departure from periodicity at nonzero osets.
On the other hand, note that the peg-leg multiples as-
sociated with the primary reection at 1.4 s have been
attenuated signicantly by deconvolution. This event
has a very small moveout; thus, its perodicity is much
preserved.
Predictive deconvolution sometimes is applied to
CMP stacked data in an eort to suppress multiples.
The performance of such an approach can be unsat-
isfactory, because the amplitude relationships between
multiples often are grossly altered by the stacking pro-
cess, primarily because of velocity dierences between
primaries and multiples. Also, geometric spreading com-
pensation by using primary velocity function adversely
aects the amplitudes of multiples on nonzero-oset
data.
There is one domain in which the periodicity and
amplitudes of multiples are preserved the slant-stack
domain. In Section 6.3, the application of predictive de-
convolution to data in the slant-stack domain for mul-
tiple suppression is discussed.
2.5 FIELD DATA EXAMPLES
The deconvolution parameters now are examined using
eld data examples. We shall discuss application of sta-
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
212 Seismic Data Analysis
2.4-36. (a) A common-shot gather, (b) after t
2
scaling and (c) after muting rst arrivals which are largely guided waves,
and (d) after spiking deconvolution. The amplitude spectra averaged over the shot record are shown at the top and the
autocorrelograms are shown at the bottom.
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
Deconvolution 213
2.5-1. An autocorrelation window test used to design de-
convolution operators. The solid bars indicate the window
boundaries. The entire 6-s length was included in (a). The
autocorrelograms are displayed beneath the records.
tistical deconvolution to pre- and poststack data. Ad-
ditionally, we shall discuss application of deterministic
deconvolution to marine data to convert the recorded
source signature to its minimum-phase equivalent, and
to land data recorded using a vibroseis source to convert
the autocorrelogram of the sweep signal to its minimum-
phase equivalent.
Prestack Deconvolution
Figure 2.5-1 shows a CMP gather that contains ve
prominent reections at around 1.1, 1.35, 1.85, 2.15,
and 3.05 s. The gather also contains strong reverber-
ations associated with these reections. The examina-
tion of the deconvolution parameters will begin with an
2.5-2. Test of operator length. The corresponding autocor-
relogram is beneath each record. The window used in auto-
correlation estimation is shown in Figure 2-5-1c. (a) Input
gather. Deconvolution using prediction lag = 4 ms (spik-
ing deconvolution), 0.1% prewhitening, and prediction lter
operator lengths (b) 40 ms, (c) 80 ms, (d) 160 ms, (e) 240
ms.
analysis of the time gate to estimate the autocorrelation
function. A rst gate selected may be the entire length
(6 s) of the record as seen in panel (a). The solid lines
on the CMP gathers refer to the gate start and end
times. The autocorrelogram of the record is shown at
the bottom of each panel. A second choice might be to
exclude the deeper part of the record where ambient
noise dominates. The start of the gate is chosen as the
rst arrival path as shown in panel (b). A third choice
may be to exclude not only the deeper portion, but
also the early part of the record that contains energy
corresponding to the guided waves as shown in panel
(c). These waves travel within the water layer and are
not part of the signal reected from the substrata.
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
214 Seismic Data Analysis
2.5-3. Test of prediction lag. The corresponding autocor-
relogram is beneath each record. The window used in auto-
correlation estimation is shown in Figure 2.5-1c. (a) Input
gather. Deconvolution using prediction lter operator length
= 160 ms, 0.1 percent prewhitening, and prediction lags (b)
12 ms, (c) 32 ms, (d) 64 ms, (e) 128 ms.
By comparing the autocorrelograms from these dif-
ferent windows, note that the third choice best repre-
sents the reverberatory character of the data as shown
in panel (c) over most of the osets. All of the traces
in the autocorrelogram within approximately the rst
150 ms have a common appearance. This early portion
of the autocorrelogram characterizes the basic seismic
wavelet contained in the data.
In general, the autocorrelation window should in-
clude part of the record that contains useful reection
signal, and should exclude coherent or incoherent noise.
An autocorrelation function contaminated by noise is
undesirable since the deconvolution process is most ef-
fective on noise-free data (assumption 4).
Another aspect of the autocorrelation window is
length. Panel (d) of Figure 2.5-1 shows the autocorrel-
2.5-4. Test of percent prewhitening. The corresponding au-
tocorrelogram is beneath each record. The window used in
autocorrelation estimation is shown in Figure 2.5-1c. (a) In-
put gather. Deconvolution using prediction lter operator
length = 160 ms, prediction lag = 4 ms (spiking deconvolu-
tion), and percent prewhitening (b) 1 percent, (c) 4 percent,
(d) 16 percent, (e) 32 percent.
ogram estimated from a narrow window. The autocor-
relogram estimated from the narrower part of the time
gate (the right side of the record) in some data cases
may lack the characteristics of the reverberations, and
even those of the basic seismic wavelet.
In general, any autocorrelation function is biased;
that is, the rst lag value is computed from, say, n
nonzero samples, the second lag value is computed from
n 1 nonzero samples, and so on. If n is not large
enough, then there can be an undesirable biasing eect.
How large should the data window be to avoid such bias-
ing? If the largest autocorrelation lag used in designing
the deconvolution operator were m, an accepted rule of
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
Deconvolution 215
2.5-5. (a) A common-shot gather, (b) after muting guided waves, (c) after t
2
-scaling, and (d) after spiking deconvolution
using an operator length of 320 ms. The amplitude spectra (top) averaged over the shot record, and the autocorrelograms
(bottom) are used to to choose deconvolution parameters and evaluate the data after the application of deconvolution.
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
216 Seismic Data Analysis
thumb is that the number of data samples should be no
less than 8m.
Now that the autocorrelation window is deter-
mined, we examine operator length. In Figure 2.5-2,
prediction lag (4 ms, the same as the sampling rate) and
percent prewhitening (0.1%) are xed. The autocorrel-
ograms (at the bottom of each gather) are displayed
for diagnostic purposes. From the analyses of the sin-
gle spike, sparse spike, and reectivity models (Section
2.4), recall that the short (40-ms) operator leaves some
residual energy that corresponds to the basic wavelet
and reverberating wavetrain in the record. For a spiking
deconvolution with a 160-ms-long operator, no remnant
of the energy is associated with the basic wavelet and
reverberations. Any operator longer than 160 ms does
not change the result, signicantly. From the output of
the 160-ms operator, note that the prominent reec-
tions (at 1.1, 1.35, 1.85, and 2.15 s at the near-oset)
have been uncovered, the seismic wavelet has been com-
pressed, and the reverberations have been signicantly
suppressed.
The eect of prediction lag now is examined. In
Figure 2.5-3, the 160-ms operator length and 0.1%
prewhitening are xed, while prediction lag is varied.
If prediction lag were increased, the deconvolution pro-
cess would be increasingly less eective in broadening
the spectrum, and the autocorrelograms would contain
increasingly more energy at nonzero lags. In the ex-
treme, the deconvolution process is ineective with a
128-ms prediction lag. In practice, common values for
the prediction lag are unity (spiking deconvolution) or
the rst or second zero crossing of the autocorrelation
function (predictive deconvolution).
Finally, the percent of prewhitening is varied, while
the 4-ms prediction lag and 160-ms operator length are
xed. These tests are shown in Figure 2.5-4. By increas-
ing the percent prewhitening, the deconvolution process
becomes less eective. The high end of the spectrum is
not attened as much as the rest of the spectrum (Fig-
ure 2.4-24). Note that the autocorrelograms contain in-
creasingly more energy at nonzero lags with increasing
percent prewhitening. In practice, it is not advisable
to assign a large percent of prewhitening. Typically, a
value between 0.1 and 1 percent is sucient to ensure
stability in designing the deconvolution operator (equa-
tion 2-32).
We now examine the eect of operator length and
prediction lag on amplitude spectrum. Figure 2.5-5a
shows a common-shot gather with its autocorrelogram
and average amplitude spectrum. The eld record is
prepared for deconvolution by rst muting the guided
waves (Figure 2.5-5b) and applying t
2
-scaling (Figure
2.5-5c). Figure 2.5-5d shows the same record after spik-
ing deconvolution. Note the attening of the spectrum
FIG. 2.5-6. The shot record in Figure 2.5-5c after predic-
tive deconvolution using an operator length of 320 ms and
a prediction lag of: (a) unit-prediction, (b) 8 ms, and (c)
24 ms. The amplitude spectra (top) averaged over the shot
record, and the autocorrelograms (bottom) are used to to
choose deconvolution parameters and evaluate the data af-
ter the application of deconvolution.
within the passband of the recorded data and attenu-
ation of the energy at nonzero lags of the autocorrel-
ogram. With prediction lag greater than unity (Figure
2.5-6), for the same operator length, we note insucient
attening at the high-frequency end of the spectrum. At
larger prediction lags, note the insucient attening at
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
Deconvolution 217
the low-frequency end of the spectrum. A very large
prediction lag causes the amplitude spectrum of the de-
convolved data remain similar to that of the input data
(compare Figures 2.5-6c with 2.5-5c).
The data sometimes must be preconditioned for de-
convolution. If the data were too noisy, then a wide
band-pass lter could be necessary before deconvolu-
tion. If there is signicant coherent noise in the data,
dip ltering (Sections 6.2 and 6.3) can be applied be-
fore deconvolution so that coherent noise is not in-
cluded in the autocorrelation estimate. Alternatively,
time-variant spectral whitening (Section 2.6) can be ap-
plied to balance the spectrum before deconvolution.
Signature Deconvolution
In marine seismic exploration, the far-eld signature of
the source array can be recorded. The idea is to ap-
ply a deterministic deconvolution to remove the source
signature, then to apply predictive deconvolution. The
convolutional model is given by
x(t) = s(t) w(t) e(t), (2 41)
where s(t) is the source signature recorded in the far-
eld just before it travels down into the earth, which
has an impulse response e(t). Since s(t) is recorded,
an inverse lter can be deterministically designed, as
discussed in Section 2.2, then applied to the recorded
seismogram to remove it from equation (2-41). The un-
known wavelet w(t) includes the propagating eects in
the earth and the response of the recording system. This
remaining wavelet then is removed by the statistical
method of spiking deconvolution as discussed in Sec-
tion 2.3. Compare equation (2-41) with equation (2-3a)
and note that the old w(t) of equation (2-3a) is split
into two parts the source signature s(t), which is
the known component, and the new w(t), which is the
unknown component.
There are two ways to handle s(t). One way is to
convert it to its minimum-phase equivalent followed by
predictive deconvolution (Figure 2.5-7). Another way
is to convert s(t) into a spike followed by predictive
deconvolution (Figure 2.5-8). The process involves the
following steps:
(a) Estimate the minimum-phase equivalent of the
recorded source signature by computing the spiking
deconvolution operator (equation 2-39) and taking
its inverse.
(b) Design a shaping lter to convert the source sig-
nature to its minimum-phase equivalent or a zero-
delay spike (equation 2-30).
FIG. 2.5-7. Signature processing. A shaping lter is de-
signed to convert the recorded signature [s(t) in equation
(2-41)] to its minimum-phase equivalent and applied to the
input record (a). The output (b) has the same bandwidth
as the input (a). The output (b) then has been processed by
predictive deconvolution using operator length of 160 ms,
and prediction lags of (c) 4 ms (spiking deconvolution), (d)
12 ms, (e) 32 ms. Examine the autocorrelograms (bottom)
and note that those of (a) and (b) should be identical.
(c) Apply the shaping lter to each trace in each
recorded shot record.
(d) Apply predictive deconvolution to output data
from step (c).
The results shown in Figures 2.5-7 and 2.5-8 (pan-
els (c)) should be compared with single-step statistical
deconvolution (Figure 2.5-9). Since the source was not
minimum-phase in this case, Figure 2.5-9b should be
better than Figure 2.5-9d. Is it?
Actually, results of signature processing depend
on the accuracy of the recorded signature. One should
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
218 Seismic Data Analysis
FIG. 2.5-8. Signature processing. A shaping lter is de-
signed to convert the recorded signature (s(t) in equation
2-41) to a spike and applied to the input record (a). The
output (b) then has been processed by predictive deconvo-
lution using an operator length of 160 ms and prediction
lags of (c) 4 ms (spiking deconvolution), (d) 12 ms, and (e)
32 ms. Note from the autocorrelograms (bottom) that the
output from signature processing (b) still contains a wavelet
component the w(t) component in equation (2-41), that
still needs to be removed.
avoid signature processing of old marine data unless
there exists a concurrently recorded source signature.
Contemporary marine data almost always include the
recorded source signature at each shot location. Fig-
ure 2.5-10a shows a recorded water-gun signature with
its amplitude and phase spectra. The minimum-phase
equivalent is shown in Figure 2.5-10b, and the result of
signature deconvolution to convert the recorded wave-
form to its minimum-phase equivalent is shown in Fig-
ure 2.5-10c. While the amplitude spectrum is unaltered,
the phase spectrum is minimum phase.
Application of signature processing described in
Figure 2.5-10 to a recorded shot gather is shown in
Figure 2.5-11. Again, note that signature processing
FIG. 2.5-9. Signature processing compared with statistical
deconvolution. (a) Input shot record. (b) A shaping lter is
designed to convert the recorded signature (s(t) in equation
2-41) to its minimum-phase equivalent and applied to the
input record followed by spiking deconvolution. (This panel
is the same as Figure 2.5-7c.) (c) A shaping lter is designed
to convert the recorded signature (s(t) in equation 2-41) to
a spike and applied to the input record followed by spiking
deconvolution. (This panel is the same as Figure 2.5-8c.) (d)
Spiking deconvolution of the input record (a). The autocor-
relograms (bottom) suggest that the wavelet compression is
achieved in all three cases, (b), (c) and (d).
aimed at converting the recorded source signature to its
minimum-phase equivalent leaves the amplitude spec-
trum and autocorrelogram of the shot record unaltered
(Figures 2.5-11a,b). We can also observe from the cor-
responding CMP-stacked sections that the process has
not made any impact on the degree of vertical resolu-
tion; only a change in phase has taken place (Figure
2.5-12). Following the deterministic step of Figure 2.5-
11b, statistical deconvolution is applied to atten the
spectrum (Figure 2.5-11c).
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
Deconvolution 219
FIG. 2.5-10. (a) A water-gun source signature and its am-
plitude and phase spectra (top and middle graphs), (b) its
minimum-phase equivalent, and (c) the ouput of signature
processing to convert (a) to (b).
Vibroseis Deconvolution
The vibroseis source is a long-duration sweep signal in
the form of a frequency-modulated sinusoid that is ta-
pered on both ends. Just as a convolutional model was
proposed for the marine seismogram given by equation
(2-41), a similar convolutional model can be proposed
for the vibroseis seismogram
x(t) = s(t) w(t) e(t), (2 42)
where x(t) is the recorded seismogram, s(t) is the sweep
signal, w(t) is the seismic wavelet with the same mean-
ing as in equation (2-41), and e(t) is the earths impulse
response. Convolutions in equation (2-42) become mul-
tiplications in the frequency domain:
X() = S() W() E(). (2 43)
In terms of amplitude A() and phase () spec-
tra, equation (2-43) yields
A
x
() = A
s
() A
w
() A
e
() (2 44a)
and

x
() =
s
() +
w
() +
e
(). (2 44b)
Crosscorrelation of the recorded seismogram x(t)
with the sweep signal s(t) is equivalent to multiplying
equation (2-44a) by A
s
() and subtracting
s
() from
equation (2-44b). The correlated vibroseis seismogram
x

(t) therefore would have the following amplitude and


phase spectra:
A

() = A
2
s
() A
w
() A
e
() (2 45a)
and

() =
w
() +
e
(). (2 45b)
The inverse Fourier transform of A
2
s
() yields the
autocorrelation of the sweep signal, which is called the
Klauder wavelet k(t). Returning to the time domain,
equations (2-45a,b) yield
x

(t) = k(t) w(t) e(t). (2 46)


Figure 2.5-13 outlines the process of vibroseis cor-
relation where the seismic wavelet w(t) has been omit-
ted for convenience. Note that, following vibroseis cor-
relation, the sweep s(t) contained in the recorded seis-
mogram x(t) is replaced with its autocorrelogram
the Klauder wavelet k(t).
Since it is an autocorrelation, the Klauder wavelet
is zero-phase. Convolution of k(t) with the assum-
ingly minimum-phase wavelet w(t) yields a mixed-phase
wavelet. Because spiking deconvolution is based on
the minimum-phase assumption, it cannot recover e(t)
properly from vibroseis data.
One approach to deconvolution of vibroseis data
is to apply a zero-phase inverse lter to remove k(t),
followed by a minimum-phase deconvolution to remove
w(t). The amplitude spectrum of the inverse lter is
dened as 1/ A
2
s
(). In practice, problems arise because
of zeroes in the spectrum that are caused by the band-
limited nature of the Klauder wavelet. Inversion of an
amplitude spectrum, which has zeroes, yields an un-
stable operator (Section 2.3). To circumvent this prob-
lem, a small percent of white noise, say 0.1%, usually is
added before inverting the Klauder wavelet spectrum.
Another approach is to design a lter that converts
the Klauder wavelet to its minimum-phase equivalent
(Ristow and Jurczyk, 1975). A technique to compute
the minimum-phase spectrum from a given amplitude
spectrum is described in Section B.4 and is included
in the discussion on frequency-domain deconvolution.
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
220 Seismic Data Analysis
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
Deconvolution 221
FIG. 2.5-12. (a) A portion of a CMP-stacked section that corresponds to the data in Figure 2.5-11a, and (b) after signature
processing of shot records as in Figure 2.5-11b to convert the recorded water-gun signature to its minimum-phase equivalent
as described in Figure 2.5-10. Note that signature processing was not aimed at wavelet compression; instead, it was done to
convert the source signature to its minimum-phase equivalent.
FIG. 2.5-13. An outline of vibroseis correlation.
If the Klauder wavelet were converted to its minimum-
phase equivalent, then equation (2-45b) would take the
form:

() =
k
() +
w
() +
e
(). (2 47)
If we assume that w(t) is minimum-phase and if we
make k(t) minimum-phase, then the result of their con-
volution also is minimum-phase. Spiking deconvolution
now is applicable since the minimum-phase assumption
is satised.
There is a 90-degree phase dierence in some vi-
brator systems between the control sweep signal and
the baseplate response. As an option, we may want to
subtract out this phase dierence. Figure 2.5-14 shows
the recommended sequence of operations for vibroseis
processing.
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
222 Seismic Data Analysis
2.5-14. A owchart for vibroseis deconvolution.
Figure 2.5-15 shows how the owchart in Figure
2.5-14 is used with a synthetic reectivity series. By
including the step to convert the Klauder wavelet into
its minimum-phase equivalent before spiking deconvo-
lution, a closer representation of the impulse response
is produced as seen by comparing steps (k) and (l) with
(m).
Although sound in theory, the above scheme may
have problems in practice. Fundamental issues, such as
whether the convolutional model given in equation (2-
42) really represents what goes on in the earth, are not
resolved.
Vibroseis data often are deconvolved as dynamite
data, without converting the Klauder wavelet to its
minimum-phase equivalent. An example of deconvolu-
tion of a correlated vibroseis record is shown in Figure
2.5-16.
Despite the fact that the basic minimum-phase as-
sumption is violated for vibroseis data as compared to
explosive data, spiking deconvolution without conver-
sion of the Klauder wavelet to its minimum-phase equiv-
alent seems to work for most eld data. Figure 2.5-17
shows a set of correlated vibroseis records before and
after spiking deconvolution. Prominent reections after
deconvolution are enhanced and reverberations are at-
tenuated signicantly. Nonetheless, the problem of ty-
ing vibrator lines to lines recorded with other sources,
say dynamite, is more dicult if the vibrator data have
not been phase-corrected. Field systems now exist to do
minimum-phase vibroseis correlation in the eld.
Poststack Deconvolution
Poststack deconvolution often is considered for sev-
eral reasons. First, a residual wavelet almost always is
present on the stacked section. This is because none of
the underlying assumptions for deconvolution is com-
pletely met in real data; therefore, deconvolution never
can completely compress the basic wavelet contained in
prestack data to a spike. Second, since a CMP stack
is an approximation to the zero-oset section, predic-
tive deconvolution aimed at removing multiples may be
a viable process after stack. Figure 2.5-18 is an exam-
ple of poststack deconvolution applied to marine data.
After deconvolution, the spectrum is attened, albeit
incompletely, the wavelet is compressed further and the
marker horizons are better characterized. Again, as the
prediction lag is increased, the atness character of the
spectrum and thus vertical resolution is increasingly
compromised (Figure 2.5-19).
Figure 2.5-20 shows poststack deconvolution ap-
plied to land data. Note the signicant improvement
in vertical resolution as it can be veried by the au-
tocorrelogram and average amplitude spectrum of the
data.
2.6 THE PROBLEM OF
NONSTATIONARITY
Figure 2.6-1 shows a CMP gather and its ltered ver-
sions before deconvolution. The lter scans show that
there is signal between 10 to 40 Hz. Note that higher fre-
quencies are conned to the shallower part of the gather.
The same eld record after spiking deconvolution is
shown in Figure 2.6-2. Filter scans of the deconvolved
record also are shown in this gure. A comparison of
the records before and after deconvolution (Figures 2.6-
1 and 2.6-2) demonstrates the eects of the process; in
particular, compression of the wavelet and broadening
of the spectrum. The input signal level above 40 Hz is
relatively weaker than that below 40 Hz. Deconvolution
has attempted to reduce the dierences between the sig-
nal levels within dierent frequency bands by attening
the spectrum. The attening, however, was more eec-
tive in the shallow part of the record than in the deeper
part.
As discussed in Sections 1.4 and 2.1, the source
wavelet is not stationary its shape and bandwidth
change with traveltime (Figure 2.1-2). Specically, at-
tenuation of high frequencies in the wavelet increases as
waves travel deeper in the subsurface. Although multi-
window deconvolution was used in this case (Figure 2.6-
2), spectral attening was not achieved over the entire
length of the record because of the large degree of non-
stationarity of the data. Nonstationarity is primarily
the result of the eects of wavefront divergence and fre-
quency attenuation.
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
Deconvolution 223
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
224 Seismic Data Analysis
2.5-16. Deconvolution of a vibroseis record. Three windows were used. (a) Correlated vibroseis record and its autocorrelograms
(b); (d) spiking deconvolution output and its autocorrelograms (c).
The phenomenon of nonstationarity on stacked
data is exemplied in Figure 2.6-3. Following a single-
gate poststack spiking deconvolution, note that the av-
erage amplitude spectrum of the data does not indicate
a at spectrum (Figure 2.6-3b). Instead, the spectral be-
havior is similar to data with predictive deconvolution
with a large prediction lag (Figure 2.6-3c). The strong
reector in the neighborhood of 2.5 s separates the zone
with two dierent bandwidths a broad-band signal
zone above, and a relatively narrow-band zone below.
The usual approach to reduce nonstationarity is to
apply processes designed to compensate for the eects
wavefront divergence and frequency attenuation before
deconvolution. Wavefront divergence is corrected for by
applying a geometric spreading function (Section 1.4).
As yet, a method to compensate for attenuation has not
been discussed. Attenuation is measured by a quantity
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
Deconvolution 225
2.5-17. (a) Common-shot gathers recorded with vibroseis source. Geometric spreading correction and trace balancing have
been applied. (b) After spiking deconvolution.
called the quality factor Q. An innite Q means that
there is no attenuation. This factor can change in depth
and in the lateral direction. If we had an analytic form
for an attenuation function, then it would be easy to
compensate for its eect. Several models for Q have
been proposed. The constant-Q model is quite plausi-
ble and the easiest to deal with (Kjartannson, 1979).
However, the big problem of estimating Q from seismic
data still remains. If a reliable Q value is available, say
from borehole measurements, then, as will be discussed
later in this section, inverse Q ltering can be applied
to data to compensate for the frequency attenuation.
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
226 Seismic Data Analysis
2.5-18. (a) A portion of a CMP-stacked section, and after spiking deconvolution using an operator length of: (b) 120 ms,
(b) 160 ms, (c) 220 ms, and (d) 320 ms. Note from the autocorrelograms (bottom) that much of the reverberating energy
is attenuated using a 320-ms operator length. The amplitude spectra averaged over the CMP stack (top) indicate that,
irrespective of how long the operator length is, spiking deconvolution in this case has failed to atten the spectrum completely
within the passband. This is because of nonstationarity of the signal.
2.5-19. A portion of a CMP-stacked section as in Figure 2.5-7a after predictive deconvolution using an operator length of 320
ms and a prediction lag of: (a) 8 ms, (b) 12 ms, (c) 24 ms, (d) 32 ms, and (e) 48 ms. The amplitude spectra (top) averaged
over the CMP stack, and the autocorrelograms (bottom) are used to to choose deconvolution parameters and evaluate the
data after the application of deconvolution.
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
Deconvolution 227
2.5-20. (a) A portion of a CMP-stacked section, and (b)
after spiking deconvolution using an operator length of 240
ms. The amplitude spectra (top) averaged over the CMP
stack, and the autocorrelograms (bottom) are used to choose
deconvolution parameters and evaluate the data after the
application of deconvolution.
Time-Variant Deconvolution
Nonstationarity was discussed in Sections 1.4 and 2.1.
The time-variant character of the seismic wavelet (Fig-
ure 2.1-2) often requires a multiwindow deconvolution.
Figure 2.6-4 is a eld record that was deconvolved
2.6-1. A eld record (far left panel) with its band-pass l-
tered versions.
2.6-2. Spiking deconvolution applied to the eld record in
Figure 2.6-1 (far left panel), followed by application of a
series of band-pass lters.
by using three time gates. The autocorrelograms from
gates 1, 2, and 3 are shown in Figure 2.6-5. Note the
dierence in character of the reverberatory energy from
one gate to another. The shallow gate (1) has more high-
frequency signal than the middle gate (2); while the
middle gate has more high-frequency signal than the
deeper gate (3). For best results, we must design dier-
ent deconvolution operators from dierent parts of the
record and apply them to the corresponding time gates.
Up to three windows usually are sucient to handle the
nonstationary character of the seismic signal.
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
228 Seismic Data Analysis
2.6-3. (a) A portion of a CMP-stacked section with its amplitude spectrum averaged over the CMP range (top) and autocor-
relogram (bottom); (b) after time-invariant spiking deconvolution; and (c) after time-invariant predictive deconvolution with
a prediction lag of 24 ms.
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
Deconvolution 229
2.6-4. Three-window deconvolution. The solid bars indicate window boundaries. With data from each window, a deconvolution
operator is designed and applied to the data in that window. The operators are blended across the window boundaries. (a)
Input gather. Deconvolution using operator length of 160 ms and prediction lags of (b) 4 ms, (c) 12 ms, (d) 32 ms.
Another example of single- and multiwindow de-
convolution is shown in Figures 2.6-6 and 2.6-7. Here,
autocorrelograms from dierent gates do not show sig-
nicant variations. Therefore, it probably does not make
any dierence whether a single or multigate deconvo-
lution is used. In Figures 2.6-6 and 2.6-7, the record
is shown after deconvolution followed by a wide band-
pass lter application. Since the amplitude spectrum of
the input data is attened as a result of spiking de-
convolution, both the high-frequency ambient noise as
much as the high-frequency components of the signal
are boosted. Therefore, the output of spiking deconvo-
lution often is ltered with a wide band-pass operator.
A practical problem with time-variant deconvolu-
tion is limiting design gates to small time windows. Con-
sider, for instance, a three-window deconvolution of a 5-
s data. This means that at best an average gate kength
of 1.5 s at near oset and less than 1 s at far oset can
be used to design a deconvolution operator. To attain
good statistics in an autocorrelation estimate, an oper-
ator length no more than one-eighth to one-tenth of the
design gate, say 150 ms, should be considered. Hence, if
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
230 Seismic Data Analysis
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
Deconvolution 231
2.6-6. Spiking deconvolution (c) on a shot record (a) followed by band-pass ltering (d). (b) Autocorrelograms before and
after deconvolution.
you need to use a longer operator, time-variant decon-
volution may have limited eectiveness in attenuating
reverberations and short-period multiples. A way to
account for nonstationarity while avoiding the short-
operator eect of multiwindow deconvolution follows.
Time-Variant Spectral Whitening
Frequency attenuation and a way to compensate for it
are illustrated in Figure 2.6-8. Let us assume that we
have an input seismogram with amplitudes decaying in
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
232 Seismic Data Analysis
FIG. 2.6-7. Three-window deconvolution on the same shot record as in Figure 2.6-6. In this case, there is no signicant dif-
ference between the characters of the autocorrelograms estimated from three windows. (a) Input gather; (b) autocorrelograms
before and after spiking deconvolution; (c) three-window spiking deconvolution on (a); (d) band-pass ltering on (c).
time, as depicted. Now apply a series of narrow band-
pass lters to this trace. Examine the eld record in Fig-
ure 2.6-1 and associate the lter panels with the traces
sketched in Figure 2.6-8. Note that the low-frequency
component of trace F
L
has a lower decay rate than
the moderate-frequency component F
M
. Likewise, the
moderate-frequency component F
M
has a lower decay
rate than the high-frequency component of the signal
F
H
. A series of gain functions, such as G
1
, G
2
, G
3
, can
be computed to describe the decay rates for each fre-
quency band. This is done by computing the envelope
of the band-pass ltered traces (Figure 2.6-8). The in-
verses of these gain functions then are applied to each
frequency band and the results are summed. The am-
plitude spectrum of the resulting trace has thus been
whitened in a time-variant manner. This time-variant
spectral whitening process is outlined in Figure 2.6-
9. The number of the lter bands, the width of each
band, and the overall bandwidth of application of time-
variant spectral whitening are parameters that can be
prescribed for a particular application.
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
Deconvolution 233
2.6-8. A schematic illustration of the rate of decay of the
frequencies in a seismic trace (Gibson and Larner, 1982).
Figures 2.6-10 and 2.6-11 show some eld records
before and after spiking deconvolution, respectively.
Note that this process not only has compressed the
wavelet, but also has tried to suppress any reverber-
ations in the data. Refer to Figure 2.6-12 and note that
time-variant spectral whitening mainly has compressed
the wavelet without changing much of the ringy char-
acter of the data. Also note that little was done explic-
itly to the phase. Therefore, the action of time-variant
spectral whitening may be close to a zero-phase decon-
volution, although there is no rigorous theoretical proof
of this.
In practice, one of the main dierences between
time-variant spectral whitening and conventional de-
convolution is that the former seems to be able to do
a better job of attening the amplitude spectrum. This
dierence can be signicant for broad-band data with
large dynamic range.
Time-variant spectral whitening sometimes helps
attenuate ground roll on land records by way of its
spectral balancing eect. Note that, in Figure 2.6-13,
spiking deconvolution with dierent operator lengths
has failed to atten the spectrum, suciently. On the
other hand, following spiking deconvolution, applica-
tion of time-variant spectral whitening has balanced the
spectrum and thus attenuated the ground-roll energy.
The ability of time-variant spectral whitening in
attening the spectrum within the passband of stacked
data is observed in Figure 2.6-14. Note that spiking
deconvolution is fairly eective, but not sucient, for
attaining a at spectrum (Figure 2.6-14b). Following
2.6-9. A owchart for time-variant spectral whitening.
time-variant spectral whitening, the spectrum is at-
tened within the passband of the data as seen in Figure
2.6-14c.
It is a requirement to prepare stacked data input
to amplitude inversion with the broadest possible band-
width and attest possible spectrum. Hence, a process-
ing sequence tailored for amplitude inversion almost al-
ways includes poststack deconvolution and time-variant
spectral attening steps.
Frequency-Domain Deconvolution
Spectral attening can be achieved by an alternate ap-
proach in the frequency domain. As discussed in Section
B.4, minimum-phase spiking deconvolution can be for-
mulated in the frequency domain. Alternatively, we can
atten the amplitude spectrum without modifying the
phase. This is called zero-phase frequency-domain de-
convolution. When performed over multiple time gates
down the trace, it is essentially equivalent to time-
variant spectral attening. If we only want to atten
the spectrum, then the approach shown in Figure 2.6-
15 can be taken.
Although the domains of operations may dier,
both minimum-phase frequency-domain and Wiener-
Levinson deconvolution techniques should yield equiv-
alent results. Dierences between the results shown in
Figures 2.6-11 and 2.6-16 mainly are due to their com-
putational aspects.
The zero-phase frequency-domain deconvolution
aimed at achieving time-variant spectral whitening re-
quires partitioning the input seismogram into small
time gates, as well as designing and applying the process
described in Figure 2.6-15 to each gate, individually.
Figure 2.6-17 shows the eld records after zero-phase
frequency-domain deconvolution. The output is com-
parable to the time-variant spectral attening output
shown in Figure 2.6-12.
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
234 Seismic Data Analysis
FIG. 2.6-10. CMP gathers with no deconvolution.
Inverse Q Filtering
Frequency attenuation caused by the intrinsic attenu-
ation in rocks was discussed in Sections 1.4 and 2.1.
Attenuation causes loss of high frequencies in the prop-
gating waveform with increasing traveltime. This gives
rise to a nonstationary behavior in the shape of the
wavelets associated with reection events at dierent
times.
Wave attenuation usually is described by a dimen-
sionless factor Q, which is dened by the ratio of the
mean stored energy to the energy loss over a period
of time that is equivalent to one cycle of a frequency
component of the waveform (Kjartansson, 1979). Time-
variant deconvolution and time-variant spectral whiten-
ing discussed in this section are processes that can cor-
rect for the time-varying eects of attenuation by spec-
tral attening. A deterministic alternative to compen-
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
Deconvolution 235
FIG. 2.6-11. Spiking deconvolution applied to the CMP gathers in Figure 2.6-10.
sate for frequency-dependent attenuation is provided by
inverse Q ltering.
The amplitude spectrum of the inverse Q lter is
given by (Section B.9)
A(, ) = exp(

2Q
), (2 48)
where is the angular frequency component associated
with the input trace and is the time variable asso-
ciated with the output trace from inverse Q ltering.
The phase spectrum of the inverse Q lter usually is as-
sumed to be minimum-phase, which can be computed
by taking the Hilbert transform of the amplitude spec-
trum (Section B.4).
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
236 Seismic Data Analysis
FIG. 2.6-12. Time-variant spectral whitening (TVSW) applied to the CMP gathers in Figure 2.6-10. Compare this with
Figures 2.6-11, 2.6-16, and 2.6-17.
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
Deconvolution 237
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
238 Seismic Data Analysis
FIG. 2.6-14. (a) A portion of a CMP stack and its amplitude spectrum averaged over the CMP range (top) and auto-
correlogram (bottom); (b) after spiking deconvolution, (c) followed by time-variant spectral whitening, and (d) time-variant
ltering.
FIG. 2.6-15. A owchart for a frequency-domain deconvo-
lution.
Application of the inverse Q lter requires knowl-
edge of the attenuation factor Q, which usually is as-
sumed to be constant. A compilation of laboratory mea-
surements of Q for some rock samples is given by Table
2-28.
Note from Table 2-28 that most measurements have
been made at extremely high frequencies compared to
the typical frequency band of seismic waves. Never-
theless, by assuming frequency-independent Q factor
(Kjartansson, 1979), these measurements can still be
considered useful. Also note that the Q factor can vary
signicantly for limestone, sandstone, and shale rock
samples of dierent origin.
The inverse of the amplitude spectrum dened by
equation (2-48) can be used to obtain a quantitative
measure of attenuation. In terms of frequency f, wave
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
Deconvolution 239
FIG. 2.6-16. Minimum-phase frequency-domain deconvolution applied to the CMP gathers in Figure 2.6-10. Compare this
with Figures 2.6-11, 2.6-12, and 2.6-17.
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
240 Seismic Data Analysis
FIG. 2.6-17. Zero-phase frequency-domain deconvolution applied to the CMP gathers in Figure 2.6-10. Compare this with
Figures 2.6-11, 2.6-12, and 2.6-16.
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
Deconvolution 241
Table 2-28. Intrinsic attenuation measurements in
rocks (adapted from Waters, 1981).
Rock Type Attenuation Frequency
Constant, Q Range (Hz)
Basalt 550 3,000-4,000
Granite 300 20,000-200,000
Quartzite 400 3,000-4,000
Limestone I 200 10,000-15,000
Limestone II 50 2-120
Limestone III 650 4-18,000
Chalk 2 150
Sandstone I 25 550-4,000
Sandstone II 125 20,000
Sandstone III 75 2,500-5,000
Sandstone IV 100 2-40
Shale I 15 75-550
Shale II 75 3,300-12,800
velocity v and depth z = v, the inverse is
A
1
(f, v, z, Q) = exp(
fz
Qv
). (2 49)
To determine how far in depth the wave has to travel
before its amplitude reduces to, say, one-tenth of its
amplitude at the surface z = 0, rewrite equation (2-49)
as follows:
z =
2.3Qv
f
. (2 50)
Note that the smaller the Q factor, the lower the
velocity and the higher the frequency, the shallower the
depth at which the wave amplitude decays to a fraction
of the wave amplitude at z = 0. Table 2-29 lists the z
values for a frequency of 30 Hz and a velocity of 3000
m/s for a range of Q values.
Table 2-29. Depth at which wave amplitude drops to
one-tenth of its original at the surface for a range of Q
values (equation 2-50).
v = 3000 m/s
f = 30 Hz
Q Factor Depth in m
25 1,830
50 3,660
100 7,325
250 18,312
500 36,625
Note that the smaller the Q factor the shallower
the depth at which the amplitude drops to the specied
value of one-tenth of the original value at the surface
z = 0. For very large Q values, that is, for small atten-
uation, the amplitude reduction to the specied value
does not take place until the wave reaches very large
depths beyond the exploration objectives.
Unfortunately, there is no reliable way to estimate
the attenuation factor Q directly from seismic data.
At best, inverse Q ltering can be applied to post- or
prestack data (Section B.9) using a range of constant Q
factors to create a Q panel, much like a lter panel (Sec-
tion 1.1). The factor that yields the attest spectrum
in combination with other signal processing applications
deconvolution and time-variant spectral whitening,
is chosen as the optimum Q value.
Deconvolution Strategies
Throughout the development of deconvolution theory,
several alternatives have been proposed to better solve
the deconvolution problem. Still, predictive deconvolu-
tion is used more than the other methods, although
the minimum-phase and white reectivity sequence as-
sumptions have been key issues of concern.
Follow the common sequence for deconvolution of
marine data in Figures 2.6-18 through 2.6-22. Note the
presence of reverberations and short-period multiples
in the CMP-stacked data with no deconvolution ap-
plied (Figure 2.6-18). Signature processing, in this case,
was done to convert the recorded source signature to
its minimum-phase equivalent (Figure 2.6-19). There-
fore, aside from phase, there is no dierence between
the sections in Figures 2.6-18 and 2.6-19. Deconvolu-
tion before stack has helped attenuate reverberations
and short-period multiples and, to some extent, com-
pressed the basic wavelet (2.6-20). The additonal step of
poststack deconvolution has restored much of the high
frequencies attenuated during stacking (Figure 2.6-21).
Finally, time-variant spectral whitening has attened
the spectrum within the passband of the data (Figure
2.6-22) and yielded a crisp section with high resolution.
The same sequence can be followed in Figures 2.6-
23 through 2.6-27. The CMP-stacked section includes
reections associated with a shallow, low-relief sedimen-
tary strata (Figure 2.6-23). Following the signature pro-
cessing (Figure 2.6-24), observe the gradual increase in
the vertical resolution by prestack deconvolution (Fig-
ure 2.6-25), poststack deconvolution (Figure 2.6-26) and
time-variant spectral whitening (Figure 2.6-27).
The following formal processing sequence for de-
convolution theoretically should yield optimum results:
(text continues on p. 247)
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
242 Seismic Data Analysis
2.6-18. A portion of a CMP-stacked section with no deconvolution.
2.6-19. The section in Figure 2.6-18 with signature processing to convert the recorded source signature to its minimum-phase
equivalent.
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
Deconvolution 243
2.6-20. The section in Figure 2.6-19 with prestack spiking deconvolution.
2.6-21. The section in Figure 2.6-20 with poststack spiking deconvolution.
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
244 Seismic Data Analysis
2.6-22. The section in Figure 2.6-21 with time-variant spectral whitening.
2.6-23. A portion of a CMP-stacked section with no deconvolution.
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
Deconvolution 245
2.6-24. The section in Figure 2.6-23 with signature processing to convert the recorded source signature to its minimum-phase
equivalent.
2.6-25. The section in Figure 2.6-24 with prestack spiking deconvolution.
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
246 Seismic Data Analysis
2.6-26. The section in Figure 2.6-25 with poststack spiking deconvolution.
2.6-27. The section in Figure 2.6-26 with time-variant spectral whitening.
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
Deconvolution 247
(a) Apply a geometric spreading compensation func-
tion to remove the amplitude loss due to wavefront
divergence.
(b) Apply an exponential gain or minimum-phase in-
verse Q lter (Hale, 1982; Hargreaves and Calvert,
1991). This compensates for frequency attenuation.
(c) Optionally apply signature processing to marine
data. For vibroseis data, apply the lter that con-
verts the Klauder wavelet to its minimum-phase
equivalent.
(d) Apply predictive deconvolution to compress the
basic wavelet and attenuate reverberations and
short-period multiples. If required, apply surface-
consistent deconvolution (Section B.8). This ac-
counts for the near-surface variations eect on the
wavelet because of inhomogeneities in the vicinity
of sources and receivers. In step (b), the vertical
variations eect on the wavelet is handled.
(e) Apply predictive deconvolution after stack to
broaden the spectrum and attenuate short-period
multiples.
(f) Apply time-variant spectral whitening. This pro-
vides further attening of the spectrum within the
signal bandwidth without aecting phase.
The idea is to do as much deterministic deconvolu-
tion as possible. Inverse Q ltering, signature deconvo-
lution, and the lter that converts the Klauder wavelet
to its minimum-phase equivalent are deterministic op-
erators. Any remaining issues then are handled by sta-
tistical means. However, for most data cases, just doing
the geometric spreading correction followed by predic-
tive prestack and poststack deconvolution is adequate.
EXERCISES
Exercise 2-1. Write the z-transform of wavelet
(1, 0,
1
4
). Design a three-term inverse lter and apply
it to the original. Hint: The z-transform of the wavelet
can be written as a product of two doublets, (1,
1
2
)
and (1,
1
2
).
Exercise 2-2. Consider the following set of
wavelets:
Wavelet A: (3,-2,1)
Wavelet B: (1,-2,3)
(a) Plot the percent of cumulative energy as a func-
tion of time for each wavelet. Use Robinsons en-
ergy delay theorem to determine the minimum- and
maximum-phase wavelet.
(b) Set up matrix equation (2-31) for each wavelet,
compute the spiking deconvolution operators, then
apply them.
(c) Let the desired output be (0, 0, 1, 0). Set up ma-
trix equation (2-30) for each wavelet, compute the
shaping lters, and apply them. Find that the error
for wavelet B with the delayed spike is smaller.
Exercise 2-3. Consider wavelet A in Exercise 2-
2. Set up matrix equation (2-32), where = 0.01, 0.1.
Note that = 0 already is assigned in Exercise 2-2. As
the percent prewhitening increases, the spikiness of the
deconvolution output decreases.
Exercise 2-4. Consider a multiple series associ-
ated with a water bottom with a reection coecient
c
w
and two-way time t
w
. Design an inverse lter to
suppress the multiples. [This is called the Backus l-
ter (Backus, 1959)].
Exercise 2-5. Consider an earth model that com-
prises a water-bottom reector and a deep reector at
two-way times of 500 and 750 ms, respectively. What
prediction lag and operator length should you choose
to suppress (a) water-bottom multiples, and (b) peg-
leg multiples?
Exercise 2-6. Refer to Figure 2-6.9. Consider the
following three bandwidths low (F
L
), medium (F
M
)
and high (F
H
), for TVSW application:
F
L
: 10 to 30 Hz
F
M
: 30 to 50 Hz
F
H
: 50 to 70 Hz
What kind of slopes should you assign to each band-
width so that the output trace has an amplitude spec-
trum that is unity over the 10-to-70-Hz bandwidth?
Exercise 2-7. If the signal character down the
trace changes rapidly (strong nonstationarity), should
you consider narrow or broad bandwidths for the lters
used in TVSW?
Exercise 2-8. Consider a minimum-phase wavelet
and the following two processes applied to it:
(a) Spiking deconvolution followed by 10-to-50-Hz
zero-phase band-pass ltering.
(b) Shaping lter to convert the minimum-phase
wavelet to a 10-to-50-Hz zero-phase wavelet.
What is the dierence between the two outputs?
Exercise 2-9. How would you design a minimum-
phase band-pass lter operator?
Exercise 2-10. Consider (a) convolving a
minimum-phase wavelet with a zero-phase wavelet, (b)
convolving a minimum-phase wavelet with a minimum-
phase wavelet, and (c) adding two minimum-phase
wavelets. Are the resulting wavelets minimum-phase?
Exercise 2-11. Consider the sinusoid shown in
Figure 1-1 (frame 1) as input to spiking deconvolution.
What is the output?
Exercise 2-12. Order the panels in Figure 2.E-1
with increasing prediction lag.
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
248 Seismic Data Analysis
2.E-1. The shot record shown in Figure 2.4-36c after predictive deconvolution using an operator length of 480 ms, and four
dierent prediction lags (Exercise 2-12). The amplitude spectra averaged over the shot record are shown at the top and the
autocorrelograms are shown at the bottom.
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
Appendix B
MATHEMATICAL FOUNDATION OF DECONVOLUTION
B.1 Synthetic Seismogram
Consider an earth model that consists of homogeneous horizontal layers with thicknesses cor-
responding to the sampling interval. The seismic impedance associated with a layer is dened
as I = v, where is density and v is the compressional-wave velocity within the layer. The
instantaneous value of seismic impedance for the kth layer is given by
I
k
=
k
v
k
. (B 1)
For a vertically incident plane wave, the pressure amplitude reection coecient associated
with an interface is given by
c
k
=
I
k+1
I
k
I
k+1
+ I
k
. (B 2)
Assume that the change of density with depth is negligible compared to the change of velocity
with depth. Equation (B-2) then takes the form
c
k
=
v
k+1
v
k
v
k+1
+ v
k
. (B 3)
If the amplitude of the incident wave is unity, then the magnitude of the reection coecient
corresponds to the fraction of amplitude reected from the interface.
With knowledge of the reection coecients, we can compute the impulse response of a
horizontally layered earth model using the Kunetz method (Claerbout, 1976). The impulse
response contains not only the primary reections, but also all the possible multiples. Finally,
convolution of the impulse response with a source wavelet yields the synthetic seismogram.
Although random noise can be added to the seismogram, the convolutional model used here
to establish the deconvolution lters does not include random noise. We also assume that the
source waveform does not change as it propagates down into the earth; hence, the convolutional
model does not include intrinsic attenuation. The convolution of a seismic wavelet w(t) with the
impulse response e(t) yields the seismogram x(t)
x(t) = w(t) e(t). (B 4)
By Fourier transforming both sides, we get
X() = W() E(), (B 5)
where X(), W(), and E() represent the complex Fourier transforms of the seismogram,
the source waveform, and the impulse response, respectively. In terms of amplitude and phase
spectra, the Fourier transforms are expressed as
X() = A
x
() exp[i
x
()], (B 6a)
W() = A
w
() exp[i
w
()], (B 6b)
and
E() = A
e
() exp[i
e
()]. (B 6c)
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
250 Seismic Data Analysis
By substituting equations (B-6) into equation (B-5), we have
A
x
() = A
w
() A
e
() (B 7a)
and

x
() =
w
() +
e
(). (B 7b)
Hence, in convolving the seismic wavelet with the impulse response, their phase spectra are
added, while the amplitude spectra are multiplied.
We assume that the earths impulse response can be represented by a white reectivity
series; hence, its amplitude spectrum is at
A
e
() = A
0
= constant. (B 8)
By substituting into equation (B-7a), we obtain
A
x
() = A
0
A
w
(). (B 9)
This implies that the amplitude spectrum of the seismogram is a scaled version of the amplitude
spectrum of the source wavelet.
We now examine the autocorrelation functions r() of x(t), w(t), and e(t). Autocorrelation
is a measure of similarity between the events on a time series at dierent time positions. It is a
running sum given by the expression
r
e
() =
1
N
N1

t=0
e
t
e
t+
, = 0, 1, 2, . . . , (N 1), (B 10)
where is time lag. A random time series is an uncorrelated series. (Strictly, it is uncorrelated
when it is continuous and innitely long.) Therefore,
r
e
() = 0, = 0 (B 11a)
and
r
e
(0) = r
0
= constant. (B 11b)
Equation (B-11) states that the autocorrelation of a perfect random series is zero at all lags
except at zero lag. The zero-lag value actually is the cumulative energy contained in the time
series:
r
0
= e
2
0
+ e
2
1
+ + e
2
N1
. (B 12)
Consider the z-transform of the convolutional model in equation (B-4):
X(z) = W(z) E(z). (B 13)
By putting 1/z in place of z and taking the complex conjugate, we get
X(1/z) = W(1/z) E(1/z), (B 14)
where the bar denotes the complex conjugate. By multiplying both sides of equations (B-13)
and (B-14), we get
X(z) X(1/z) =
_
W(z) E(z)
_
W(1/z) E(1/z)

. (B 15)
By rearranging the right side,
X(z) X(1/z) =
_
W(z) W(1/z)
_
E(z) E(1/z)

. (B 16)
Finally, by denition, equation (B-16) yields
r
x
= r
w
r
e
, (B 17)
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
Deconvolution 251
where r
x
, r
w
, and r
e
are the autocorrelations of the seismogram, seismic wavelet, and impulse
response, respectively. Based on the white reectivity series assumption (equation B-11), we
have
r
x
= r
0
r
w
. (B 18)
Equation (B-18) states that the autocorrelation of the seismogram is a scaled version of that
of the seismic wavelet. We will see that conversion of the seismic wavelet into a zero-lag spike
requires knowledge of the wavelets autocorrelation. Equation (B-18) suggests that the autocor-
relation of the seismogram can be used in lieu of that of the seismic wavelet, since the latter
often is not known.
B.2 The Inverse of the Source Wavelet
A basic purpose of deconvolution is to compress the source waveform into a zero-lag spike so
that closely spaced reections can be resolved. Assume that a lter operator f(t) exists such
that
w(t) f(t) = (t), (B 19)
where (t) is the Kronecker delta function. The lter f(t) is called the inverse lter for w(t).
Symbolically, f(t) is expressed in terms of the seismic wavelet w(t) as
f(t) = (t)
1
w(t)
. (B 20)
The z-transform of the seismic wavelet with a nite length m + 1 is (Appendix A)
W(z) = w
0
+ w
1
z + w
2
z
2
+ + w
m
z
m
. (B 21)
The z-transform of the inverse lter can be obtained by polynomial division:
F(z) =
1
W(z)
. (B 22)
The result is another polynomial whose coecients are the terms of the inverse lter
F(z) = f
0
+ f
1
z + f
2
z
2
+ + f
n
z
n
+ . (B 23)
Note that the polynomial F(z) in equation (B-23) has only positive powers of z; this means
f(t) is causal. If the coecients of F(z) asymptotically approach zero as time goes to innity,
so that the lter has nite energy, we say that the lter f(t) is realizable. If the coecients
increase without bound, we say that the lter is not realizable. In practice, we prefer to work
with a causal and realizable lter. Such a lter, by denition, also is minimum-phase. If f(t) is
minimum-phase, then the seismic wavelet w(t) also must be minimum-phase. Finally, to apply
the lter with a nite length n + 1, the polynomial F(z) must be truncated. Truncation of the
lter operator induces some error in spiking the seismic wavelet.
The inverse of the seismic wavelet also can be computed in the frequency domain. By Fourier
transforming equation (B-19), we get
W() F() = 1. (B 24)
By substituting equation (B-6b), we obtain
F() =
1
A
w
() exp [i
w
()]
. (B 25)
We express the Fourier transform of the inverse lter F() as
F() = A
f
() exp[i
f
()], (B 26)
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
252 Seismic Data Analysis
and compare it with equation (B-25) to get
A
f
() =
1
A
w
()
(B 27a)
and

f
() =
w
(). (B 27b)
Equations (B-27) show that the amplitude spectrum of the inverse lter is the inverse of that
of the seismic wavelet, and the phase spectrum of the inverse lter is negative of that of the
seismic wavelet.
B.3 The Inverse Filter
Instead of the polynomial division procedure described by equation (B-22), consider a dierent
approach to derive the inverse lter. Start with the z-transform of the autocorrelation of the
seismic wavelet:
R
w
(z) = W(z) W(1/z), (B 28)
and the z-transform of equation (B-19):
W(z) F(z) = 1, (B 29)
from which we get
W(z) =
1
F(z)
. (B 30)
By substituting into equation (B-28), we obtain
R
w
(z) F(z) = W(1/z). (B 31)
Since the wavelet is a real-time function,
r
w
() = r
w
(). (B 32)
Consider the special case of a three-point inverse lter (f
0
, f
1
, f
2
). We will assume that the
seismic wavelet w(t) is minimum-phase (causal and realizable); hence, its inverse f(t) also is
minimum-phase. The z-transform of f(t) is
F(z) = f
0
+ f
1
z + f
2
z
2
. (B 33a)
The z-transform of its autocorrelogram r
w
() is, by way of equation (B-32),
R
w
(z) = + r
2
z
2
+ r
1
z
1
+ r
0
+ r
1
z + r
2
z
2
+
(B 33b)
The z-transform of w(t) is
W(z) = w
0
+ w
1
z + w
2
z
2
+ + w
m
z
m
,
therefore,
W(1/z) = w
0
+ w
1
z + w
2
z
2
+ + w
m
z
m
.
(B 33c)
By substituting equations (B-33a), (B-33b), and (B-33c) into equation (B-31), we obtain
_
r
2
z
2
+ r
1
z
1
r
0
+ r
1
z + r
2
z
2
_ _
f
0
+ f
1
z + f
2
z
2
_
= w
0
+ w
1
z
1
+ w
2
z
2
. (B 34)
To solve for (f
0
, f
1
, f
2
), we identify the coecients of powers of z. The coecient of z
0
yields
r
0
f
0
+ r
1
f
1
+ r
2
f
2
= w
0
,
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
Deconvolution 253
the coecient of z
1
yields
r
1
f
0
+ r
0
f
1
+ r
1
f
2
= 0,
while the coecient of z
2
yields
r
2
f
0
+ r
1
f
1
+ r
0
f
2
= 0.
When put into matrix form, these equations for the coecients of powers of z yield
_
_
r
0
r
1
r
2
r
1
r
0
r
1
r
2
r
1
r
0
_
_
_
_
f
0
f
1
f
2
_
_
=
_
_
w
0
0
0
_
_
. (B 35)
Note that w
0
, which equals w
0
for the usual case of a real source wavelet, is the amplitude of
the wavelet at t = 0. There are four unknowns and three equations. By normalizing with respect
to f
0
, we get
_
_
r
0
r
1
r
2
r
1
r
0
r
1
r
2
r
1
r
0
_
_
_
_
1
a
1
a
2
_
_
=
_
_
L
0
0
_
_
, (B 36a)
where a
1
= f
1
/ f
0
, a
2
= f
2
/ f
0
, and L = w
0
/ f
0
. We now have three unknowns, a
1
, a
2
, and L,
and three equations. The square matrix elements on the left side of the equation represent the
autocorrelation lags of the seismic wavelet, which we do not know. However, the autocorrelation
lags from equation (B-18) of the seismogram that we do know can be substituted.
For the general case of an n-point inverse lter (f
0
, f
1
, f
2
, . . . , f
n
), equation (B-36a) takes
the form
_
_
_
_
_
_
r
0
r
1
r
2
r
n
r
1
r
0
r
1
r
n1
r
2
r
1
r
0
r
n2
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
r
n
r
n1
r
n2
r
0
_
_
_
_
_
_
_
_
_
_
_
_
1
a
1
a
2
.
.
.
a
n1
_
_
_
_
_
_
=
_
_
_
_
_
_
L
0
0
.
.
.
0
_
_
_
_
_
_
. (B 36b)
The autocorrelation matrix in equation (B-36b) is of a special type. First, it is a symmetric
matrix; second, its diagonal elements are identical. This type of matrix is called a Toeplitz
matrix. For n number of normal equations, the standard algorithms require a memory space
that is proportional to n
2
and CPU time that is proportional to n
3
. Because of the special
properties of the Toeplitz matrix, Levinson devised a recursive scheme (Claerbout, 1976) that
requires a memory space and CPU time proportional to n and n
2
, respectively (Section B.6).
B.4 Frequency-Domain Deconvolution
We want to estimate a minimum-phase wavelet from the recorded seismogram. The inverse of the
wavelet is the spiking deconvolution operator. We start with the autocorrelation of the seismic
wavelet w(t) in the frequency domain:
R
w
() = W() W(), (B 37)
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
254 Seismic Data Analysis
where W() is the complex conjugate of the Fourier transform of w(t). Since w(t) normally is
not known, R
w
() is not known. However, based on an assumption of white reectivity series
(equation B-18), we can substitute the autocorrelation function of the seismogram in equation
(B-37).
Dene a new function U():
U() = ln
_
R
w
()

. (B 38)
When both sides of equation (B-38) are exponentiated:
R
w
() = exp
_
U()

. (B 39)
Suppose that another function, (), is dened and that equation (B-39) is rewritten as
(Claerbout, 1976)
R
w
() = exp
_
1
2
_
U() + i()

_
exp
_
1
2
_
U() i()

_
(B 40)
When equation (B-40) is compared with equation (B-37), we see that
W() = exp
_
1
2
_
U() + i()

_
. (B 41)
We know R
w
() from equation (B-18) and therefore we know U() from equation (B-38).
To estimate W() from equation (B-41), we also need to know (). If we make the minimum-
phase assumption, () turns out to be the Hilbert transform of U() (Claerbout, 1976). To
perform the Hilbert transform, rst inverse Fourier transform U() back to the time domain.
Then, double the positive time values, leave the zero-lag alone, and set the negative time values
to zero. This operation yields a time function u
+
(t), which vanishes before t = 0. Then, return
to the transform domain to get
U
+
() =
1
2
_
U() + i()

, (B 42)
where U
+
() is the Fourier transform of u
+
(t). Finally, exponentiating U
+
() yields the Fourier
transform W() of the minimum-phase wavelet w(t) (equation B-41).
Once the minimum-phase wavelet w(t) is estimated, its Fourier transform W() is rewritten
in terms of its amplitude and phase spectra,
W() = A() exp
_
i()

. (B 43)
The deconvolution lter in the Fourier transform domain is
F() =
1
W()
. (B 44)
By substituting equation (B-43), we obtain the amplitude and phase spectra of this lter:
A
f
() =
1
A()
(B 45a)
and

f
() = (). (B 45b)
Since the estimated wavelet w(t) is minimum phase, it follows that the deconvolution lter
[whose Fourier transform is given by equation (B-44)] also is minimum phase. By inverse Fourier
transforming, equation (B-44) yields the deconvolution operator. The frequency-domain method
of deconvolution is outlined in Figure B-1.
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
Deconvolution 255
FIG. B-1. Flowchart for frequency-domain deconvolution.
Figure B-2 illustrates the frequency-domain deconvolution that is described in Figure B-1.
The panels in Figure B-2 should be compared with the corresponding results of the Wiener-
Levinson (time-domain) method in Figure 2-20. As expected, there is virtually no dierence
between the two results.
From the discussion so far, we see that the spiking deconvolution operator is the inverse of
the minimum-phase equivalent of the seismic wavelet. The process of estimating the minimum-
phase equivalent of a seismic wavelet is called spectral decomposition. The minimum-phase
wavelet, once computed, can be inverted easily to get the deconvolution operator. To avoid
dividing by zero in equation (B-45a) and to ensure that the lter is stable, a small number
often is added to the amplitude spectrum before division. This is called prewhitening. Equation
(B-45a) then takes the form
A
f
() =
1
A() +
. (B 46)
The amplitude spectrum also can be smoothed to get a more stable operator. Smoothing the
spectrum is analogous to shortening the equivalent time-domain deconvolution operator.
Zero-phase deconvolution can be implemented conveniently in the frequency domain. To do
this, the phase spectrum given by equation (B-45b) is set to zero and thus yields a deconvolution
operator that attens the amplitude spectrum of the input seismogram, but does not alter the
phase.
B.5 Optimum Wiener Filters
The following concise discussion of optimum Wiener lters is based on Robinson and Treitel
(1980). Consider the general lter model in Figure B-3. Wiener ltering involves designing the
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
256 Seismic Data Analysis
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
Deconvolution 257
FIG. B-3. Wiener lter model.
lter f(t) so that the least-squares error between the actual and desired outputs is minimum.
Error L is dened as
L =

t
(d
t
y
t
)
2
. (B 47)
The actual output is the convolution of the lter with the input
y
t
= f
t
x
t
. (B 48)
When equation (B-48) is substituted into equation (B-47), we get
L =

t
_
d
t

x
t
_
2
. (B 49)
The goal is to compute the lter coecients (f
0
, f
1
, . . . , f
n1
) so that the error is minimum.
Filter length n must be predened. The minimum error is attained by setting the variation of
L with respect to f
i
to zero:
L
f
i
= 0, i = 0, 1, 2, . . . , (n 1). (B 50)
By expanding the square term in equation (B-49), we have
L =

t
d
2
t
2

t
d
t

x
t
+

t
_

x
t
_
2
. (B 51)
By taking the partial derivatives and setting them to zero, we get
L
f
i
= 2

t
d
t
x
ti
+ 2

t
_

x
t
_
x
ti
= 0, (B 52)
or

t
x
t
x
ti
=

t
d
t
x
ti
,
i = 0, 1, 2, . . . , (n 1).
(B 53)
By using

t
x
t
x
ti
= r
i
(B 54a)
and

t
d
t
x
ti
= g
i
(B 54b)
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
258 Seismic Data Analysis
for each ith term, we get

r
i
= g
i
, i = 0, 1, 2, . . . , (n 1). (B 55)
When put into matrix form, equation (B-55) becomes
_
_
_
_
r
0
r
1
r
2
r
n1
r
1
r
0
r
1
r
n2
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
r
n1
r
n2
r
n3
r
0
_
_
_
_
_
_
_
_
f
0
f
1
.
.
.
f
n1
_
_
_
_
=
_
_
_
_
g
0
g
1
.
.
.
g
n1
_
_
_
_
. (B 56)
Here, r
i
are the autocorrelation lags of the input and g
i
are the lags of the crosscorrelation
between the desired output and input. Since the autocorrelation matrix is Toeplitz, the optimum
Wiener lter coecients f
i
can be computed by using Levinson recursion (Claerbout, 1976).
The least-squares error involved in this process now is computed. Expressed again, equation
(B-51) becomes
L
min
=

t
d
2
t
2
_

d
t
f

x
t
_
+

t
_

x
t
_
2
. (B 57)
By substituting the relationships

t
d
t
x
t
= g

(B 58a)
and

t
x
t
x
t
= r

(B 58b)
into equation (B-57), we get
L
min
=

t
d
2
t
2

i
f
i

t
x
t
x
ti
(B 59)
or
L
min
=

t
d
2
t
2

i
f
i
r
i
. (B 60)
Finally, by using equation (B-55), we get
L
min
=

t
d
2
t

. (B 61)
B.6 Spiking Deconvolution
Consider the desired output to be the zero-delay spike d
t
: (1, 0, 0, . . .). Given the input series
x
t
: (x
0
, x
1
, x
2
, . . .), equation (B-56) takes the form
_
_
_
_
r
0
r
1
r
2
r
n1
r
1
r
0
r
1
r
n2
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
r
n1
r
n2
r
n3
r
0
_
_
_
_
_
_
_
_
f
0
f
1
.
.
.
f
n1
_
_
_
_
=
_
_
_
_
x
0
0
.
.
.
0
_
_
_
_
. (B 62)
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
Deconvolution 259
Divide both sides by f
0
to obtain
_
_
_
_
r
0
r
1
r
2
r
n1
r
1
r
0
r
1
r
n2
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
r
n1
r
n2
r
n3
r
0
_
_
_
_
_
_
_
_
1
a
1
.
.
.
a
n1
_
_
_
_
=
_
_
_
_
v
0
.
.
.
0
_
_
_
_
, (B 63)
where a
i
= f
i
/f
0
, i = 1, 2, . . . , n 1, and v = x
0
/f
0
. Equation (B-63) is solved for the unknown
quantity v and the lter coecients (a
1
, a
2
, . . . , a
n1
). Since the desired output is a zero-delay
spike, the lter (1, a
1
, a
2
, . . . , a
n1
) describes a spiking deconvolution process.
The solution to equation (B-63) can be obtained eciently using the Levinson recursion
(Claerbout, 1976). Start with the solution of equation (B-63) for the two-term lter (1, a
1
), then
solve for the lter (1, a
1
, a
2
) and so on. Equation (B-63) for n = 2 takes the form
_
r
0
r
1
r
1
r
0
__
1
a
1
_
=
_
v
0
_
. (B 64)
Write out the simultaneous equations:
r
0
+ r
1
a
1
= v
and
r
1
+ r
0
a
1
= 0,
which yield the lter coecient a
1
and the unknown variable v from the rst iteration:
a
1
= r
1
/r
0
(B 65a)
and
v = r
0
+ r
1
a
1
. (B 65b)
Now write equation (B-63) for the three-term lter (1, a

1
, a

2
):
_
_
r
0
r
1
r
2
r
1
r
0
r
1
r
2
r
1
r
0
_
_
_
_
1
a

1
a

2
_
_
=
_
_
v

0
0
_
_
. (B 66)
We want to solve for the lter coecients (a

1
, a

2
) by using the results of the previous step
(equations B-65a,b). First, rewrite equation (B-64) by adding a row in the following manner:
_
_
r
0
r
1
r
2
r
1
r
0
r
1
r
2
r
1
r
0
_
_
_
_
1
a
1
0
_
_
=
_
_
v
0
e
_
_
, (B 67)
where
e = r
2
+ r
1
a
1
. (B 68)
Rewrite equation (B-67) by changing the order of the rows:
_
_
r
0
r
1
r
2
r
1
r
0
r
1
r
2
r
1
r
0
_
_
_
_
0
a
1
1
_
_
=
_
_
e
0
v
_
_
, (B 69)
Multiply both sides of equation (B-69) with variable c, yet to be determined, and subtract the
result from equation (B-67):
_
_
r
0
r
1
r
2
r
1
r
0
r
1
r
2
r
1
r
0
_
_
_
_
1
a
1
ca
1
c
_
_
=
_
_
v ce
0
e cv
_
_
, (B 70)
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
260 Seismic Data Analysis
Finally, compare equation (B-70) with equation (B-66), and note that
_
_
1
a

1
a

2
_
_
=
_
_
1
a
1
ca
1
c
_
_
(B 71a)
and
_
_
v

0
0
_
_
=
_
_
v ce
0
e cv
_
_
. (B 71b)
Solve for c and v

:
c =
e
v
, (B 72a)
and
v

= v
_
1
_
e
v
_
2
_
. (B 72b)
Hence, the new lter coecients (a

1
, a

2
) are (equations B-71a)
a

1
= a
1

e
v
a
1
(B 73a)
and
a

2
=
e
v
. (B 73b)
This recursive scheme is repeated to determine the next set of lter coecients:
(a) Compute v and e using the autocorrelation lags of the input series and the present lter
coecients (equations B-67).
(b) Compute the next set of lter coecients (equations B-71a).
(c) Compute a new value for v (equation B-72b).
This recursive process yields the Wiener lter (1, a
1
, a
2
, . . . , a
n1
) of the desired length n.
B.7 Predictive Deconvolution
Suppose that the desired output in the lter model of Figure B-3 is a time-advanced version of
the input, d(t) = x(t + ). We want to design a Wiener lter f(t) that predicts x(t + ) from
the past values of the input x(t). In this special case, the crosscorrelation function g becomes
g

t
d
t
x
t
=

t
x
t+
x
t
=

t
x
t
x
t(+)
. (B 74)
By denition, we have
r

t
x
t
x
t
. (B 75)
For the + lag, equation (B-75) becomes
r
+
=

t
x
t
x
t(+)
. (B 76)
Combine equations (B-74) and (B-76) and note that r
+
= g

. By substituting this result into


equation (B-56), we get the set of normal equations that must be solved to nd the prediction
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
Deconvolution 261
lter (f
0
, f
1
, . . . , f
n1
):
_
_
_
_
r
0
r
1
r
2
r
n1
r
1
r
0
r
1
r
n2
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
r
n1
r
n2
r
n3
r
0
_
_
_
_
_
_
_
_
f
0
f
1
.
.
.
f
n1
_
_
_
_
=
_
_
_
_
r

r
+1
.
.
.
r
+n1
_
_
_
_
. (B 77)
For a unit prediction lag, = 1, equation (B-77) takes the form:
_
_
_
_
r
0
r
1
r
2
r
n1
r
1
r
0
r
1
r
n2
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
r
n1
r
n2
r
n3
r
0
_
_
_
_
_
_
_
_
f
0
f
1
.
.
.
f
n1
_
_
_
_
=
_
_
_
_
r
1
r
2
.
.
.
r
n
_
_
_
_
. (B 78)
By augmenting the right side to the square matrix on the left side, we have
_
_
_
_
r
1
r
0
r
1
r
2
r
n1
r
2
r
1
r
0
r
1
r
n2
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
r
n
r
n1
r
n2
r
n3
r
0
_
_
_
_
_
_
_
_
_
_
1
f
0
f
1
.
.
.
f
n1
_
_
_
_
_
_
=
_
_
_
_
_
_
0
0
0
.
.
.
0
_
_
_
_
_
_
(B 79)
By adding one row, then putting the negative sign to the lter column, we obtain
_
_
_
_
_
_
r
0
r
1
r
2
r
n
r
1
r
0
r
1
r
n1
r
2
r
1
r
0
r
n2
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
r
n
r
n1
r
n2
r
0
_
_
_
_
_
_
_
_
_
_
_
_
1
f
0
f
1
.
.
.
f
n1
_
_
_
_
_
_
=
_
_
_
_
_
_
L
0
0
.
.
.
0
_
_
_
_
_
_
. (B 80)
We now have n+1 equations and n+1 unknowns (f
0
, f
1
, . . . , f
n1
, L). From equation (B-80)
L = r
0
r
1
f
0
r
2
f
1
r
n
f
n1
. (B 81)
By using equation (B-61), the minimum error associated with the unit prediction-lag lter
can be computed as follows. Start with
d
t
= x
t+1
(B 82)
so that

t
d
2
t
= r
0
(B 83a)
and
g
t
= r
t+1
. (B 83b)
Substituting into equation (B-61), we get
L
min
= r
0
(r
1
f
0
+ r
2
f
1
+ + r
n
f
n1
), (B 84)
which is identical to quantity L given by equation (B-81). Therefore, when we solve equation
(B-80), we compute both the minimum error and the lter coecients.
Equation (B-77) gives us the prediction lter with prediction lag . The desired output is
x
t+
. The actual output is x
t+
, which is an estimate of the desired output. This desired output
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
262 Seismic Data Analysis
is the predictable component of the input series; i.e., periodic events such as multiples. The error
series contains the unpredictable component of the series and is dened as
e
t+
= x
t+
x
t+
= x
t+

x
t
. (B 85)
We assume that the unpredictable component e
t
is the uncorrelated reectivity we want to
extract from the seismogram. By taking the z-transform of equation (B-85), we have
z

E(z) = z

X(z) F(z) X(z) (B 86)


or
E(z) =
_
1 z

F(z)

X(z). (B 87)
Now dene a new lter a(t) whose z-transform is
A(z) = 1 z

F(z) (B 88)
so that, when substituted into equation (B-87), we have
E(z) = A(z) X(z). (B 89)
The corresponding time-domain relationship is
e(t) = a(t) x(t). (B 90)
After dening e(t) as reectivity, equation (B-90) states that by applying lter a(t) to the
input seismogram x(t), we obtain the reectivity series. Since computing the reectivity series
is a goal of deconvolution, prediction ltering can be used for deconvolution. The time-domain
form a(t) of the lter dened by equation (B-88) is
a
t
= (1,
1
..
0, 0, , 0, f
0
, f
1
, , f
n1
). (B 91)
The lter a(t) is obtained from the prediction lter f(t), which is the solution to equation
(B-77). We call a(t) the prediction error lter. For unit-prediction lag, the lter coecients given
by equation (B-91) are of the form
a
t
= (1, f
0
, f
1
, , f
n1
). (B 92)
This is the same as the solution of equation (B-80). Moreover, equation (B-80) is equivalent
to equation (B-36) for n = 2. Thus, we can conclude that a prediction error lter with unit-
prediction lag and with n + 1 length is equivalent to an inverse lter of the same length except
for a scale factor.
B.8 Surface-Consistent Deconvolution
Deconvolution can be formulated as a surface-consistent spectral decomposition (Taner and
Coburn, 1981). In such a formulation, the seismic trace is decomposed into the convolutional
eects of source, receiver, oset, and the earths impulse response, thus explicitly accounting
for variations in wavelet shape aected by near-source and near-receiver conditions and source-
receiver separation. Decomposition is followed by inverse ltering to recover the earths impulse
response. The assumption of surface-consistency implies that the basic wavelet shape depends
only on the source and receiver locations, and not on the details of the raypath from source to
reector to receiver.
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
Deconvolution 263
The convolutional model discussed in Section 2.1 is described by
x(t) = w(t) e(t) + n(t), (B 93)
where x(t) is the recorded seismogram, w(t) is the source waveform, e(t) is the earths impulse
response that we want to estimate, and n(t) is the noise component.
A postulated surface-consistent convolutional model is given by
x

ij
(t) = s
j
(t) h
l
(t) e
k
(t) g
i
(t) + n(t), (B 94)
where x

ij
(t) is a model of the recorded seismogram, s
j
(t) is the waveform component associated
with source location j, g
i
(t) is the component associated with receiver location i, and h
l
(t) is
the component associated with oset dependency of the waveform dened for each oset index
l = |i j|. As in equation (B-93), e
k
(t) represents the earths impulse response at the source-
receiver midpoint location, k = (i + j)/2. By comparing equations (B-93) and (B-94), we infer
that w(t) represents the combined eects of s(t), h(t), and g(t). Cambois and Stoa (1992) oer
an alternative to equation (B-94) in which the oset term is ignored.
To illustrate a method of computing s
j
(t), h
l
(t), e
k
(t), and g
i
(t), assume n(t) = 0 and
Fourier transform equation (B-94):
X

ij
() = S
j
() H
l
() E
k
() G
i
(). (B 95)
This equation can be separated into the following amplitude spectral components:
X

ij
() = S
j
() H
l
() E
k
() G
i
(), (B 96a)
and phase spectral components (Section A.1):

ij
() =
sj
() +
hl
() +
ek
() +
ri
(). (B 96b)
If the minimum-phase assumption is made, only the amplitude spectra (equation B-96a) need
to be considered.
Equation (B-96a) now can be linearized by taking the logarithm of both sides:

ij
() =

S
j
() +

H
l
() +

E
k
() +

G
i
(). (B 97)
The left side is the logarithm of the amplitude spectrum X

ij
() of the modeled input trace as
in the left-hand side of equation (B-96a), and the right-hand terms are the logarithms of the
amplitude spectra of the individual components as in the right-hand side of equation (B-96a).
An alternative model equation is given by Cary and Lorentz (1993) based on the work by
Morley and Claerbout (1983):

ij
() =

S
j
() +

H
l
() +

E
k
() +

G
i
(), (B 98a)
where

ij
() =

X

ij
()

X

avg
(). (B 98b)
The spectral component

X

avg
() is associated with the average amplitude spectrum for the
entire data set. The terms on the right-hand side of equation (B-98) now correspond to residual
spectral components for each source, oset, midpoint, and receiver location.
The spectral components on the right-hand side of equation (B-98a) can be computed by
least-squares error minimization. For each frequency , equation (B-98a) is written for each trace
of each CMP gather in the data set. Consider a data set with n
s
shot locations and n
c
channels,
so that the total number of traces is n
s
n
c
. For n
s
n
c
values of the actual spectral components

X
ij
at frequency , and n
s
shot locations, n
r
receiver locations, n
e
midpoint locations, and n
h
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
264 Seismic Data Analysis
osets, we have the following set of model equations:
_
_
_
_
_
_
_
_
_
_
_
_
_
_
_
.
.
.

ij
.
.
.
_
_
_
_
_
_
_
_
_
_
_
_
_
_
_
=
_
_
_
_
_
_
_
_
_
_
_
_
_
1 1 1 1
_
_
_
_
_
_
_
_
_
_
_
_
_
_
_
_
_
_
_
_
_
_
_
_
_
_
_
_
_
_
_
_
_
_
_
_
_
_
_
.
.
.

S
j
.
.
.
.
.
.

H
l
.
.
.
.
.
.

E
k
.
.
.
.
.
.

G
i
.
.
.
_
_
_
_
_
_
_
_
_
_
_
_
_
_
_
_
_
_
_
_
_
_
_
_
_
_
. (B 99a)
Write equation (B-99a) in matrix notation:

= Lp, (B 99b)
where

X

is the column vector of m-length on the left-hand side in equation (B-99), L is the
sparse matrix with dimensions (n
s
n
c
) (n
s
+ n
h
+ n
e
+ n
r
) and p is the column vector of
(n
s
+n
h
+n
e
+n
r
)-length on the right-hand side of the same equation. Except the four elements
in each row, the L matrix contains zeros.
We want to estimate for each frequency the model parameters p such that the dierence
between the actual spectral component

X and the modeled spectral component

X

is minimum
in the least-squares sense.
The error vector v is dened as the dierence between the modeled and the actual spectral
component for each frequency
v =

X

X

. (B 100a)
Substitute equation (B-99b) into equation (B-100a) to obtain
v =

XLp. (B 100b)
Following Lines and Treitel (1984), the least-squares solution for equation (B-100b) can be
determined. First, the cumulative squared error C is expressed as
C = v
T
v. (B 101a)
where T is for transpose and is for complex conjugate. By substituting for v from equation
(B-100b), we get
C = (

XLp)
T
(

XLp). (B 101b)
Minimization of C with respect to p requires that
C


S
j
=
C


H
l
=
C


E
k
=
C


G
i
= 0.
This requirement yields the desired least-squares solution:
p = (L
T
L)
1
L
T

X. (B 102)
Application of the least-squares minimization to surface-consistent prediction-error ltering
is given by Levin (1989). A practical scheme for solving equation (B-98a) is based on the Gauss-
Seidel method. In this scheme, each term on the right-hand side of equation (B-98a) is computed
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
Deconvolution 265
by the following set of recursive equations:

S
m
j
=
1
n
r
n
r

i
{

X
ij


H
m1
l


E
m1
k


G
m1
i
}, (B 103a)

G
m
i
=
1
n
s
n
s

j
{

X
ij


H
m1
l


E
m1
k


S
m1
j
}, (B 103b)

H
m
l
=
1
n
e
n
e

k
{

X
ij


S
m1
j


E
m1
k


G
m1
i
}, (B 103c)
and

E
m
k
=
1
n
h
n
h

l
{

X
ij


S
m1
j


H
m1
l


G
m1
i
}, (B 103d)
where m is the iteration index. The solutions in equations (B-103) are based on the orthogonality
of the shot and receiver axes, and the orthogonality of the midpoint and oset axes. Equations
(B-103) can be modied as follows:

S
m
j
=
1
n
r
n
r

i
{

X
ij
}
1
n
r
n
r

i
{

H
m1
l


E
m1
k


G
m1
i
}, (B 104a)

G
m
i
=
1
n
s
n
s

j
{

X
ij
}
1
n
s
n
s

j
{

H
m1
l


E
m1
k


S
m1
j
}, (B 104b)

H
m
l
=
1
n
e
n
e

k
{

X
ij
}
1
n
e
n
e

k
{

S
m1
j


E
m1
k


G
m1
i
}, (B 104c)
and

E
m
k
=
1
n
h
n
h

l
{

X
ij
}
1
n
h
n
h

l
{

S
m1
j


H
m1
l


G
m1
i
}. (B 104d)
This modication enables us to compute and store the sum of the spectral components of input
data

X
ij
, thus circumventing the need for storing the individual spectral components X
ij
(Cary and Lorentz, 1993). The process is iterated until an index m that attains the least-squares
minimization.
The parameter vector p that contains the spectral components

S
j
,

G
i
,

H
l
, and

E
k
, which
are associated with the source and receiver locations, oset dependency, and earths impulse
response, is solved for each frequency component using equations (B-104). Results from all
frequency components are then combined to obtain the terms in equation (B-98a). The surface-
consistent spiking deconvolution operator to be applied to each trace in the data set is then
the minimum-phase inverse of s
j
(t) g
i
(t) h
l
(t). In the case of predictive deconvolution with a
desired prediction lag, for each source, receiver and, midpoint location, a deconvolution operator
is computed by using the autocorrelograms of the terms s
j
(t), g
i
(t), and h
l
(t). To each trace in
the data set, these operators are then applied in a cascaded manner.
As a by-product of the derivation of the surface-consistent spectral decomposition equation
(B-97), trace amplitudes themselves can be corrected for in a surface-consistent manner (Taner
and Koehler, 1981). Sum the individual terms in equation (B-97) over the frequencies:

ij
() =

S
j
() +

H
l
() +

E
k
() +

G
i
(). (B 105a)
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
266 Seismic Data Analysis
Each term yields a scalar that is related to the source, receiver, oset and midpoint locations:
x

ij
= s
j
+

h
l
+ e
k
+ g
i
. (B 105b)
Once computed in the same manner as the terms in equation (B-98a), these scalars are then
applied to individual traces in the data set for surface-consistent amplitude corrections.
In practice, application of surface-consistent deconvolution to eld data usually involves
two terms, only the source term s
j
(t) and the receiver term g
i
(t). In a transition zone,
surface conditions at source and receiver locations may vary signicantly from dry to wet surface
conditions. Hence, the most likely situation where surface-consistent amplitude corrections and
deconvolution may be required is transition-zone data. Figure B-4 shows a eld data example of
surface-consistent deconvolution. Note the variations in the autocorrelograms from one source
location to the next and from one receiver location to the next. Dierences in reection continuity
are observed within the rst 1 s on the stacked sections created by application of conventional
trace-by-trace deconvolution and surface-consistent deconvolution.
B.9 Inverse Q Filtering
Consider a 1-D seismogram that represents a compressional plane wave that propagates vertically
downward in a homogeneous medium with intrinsic attenuation. This plane wave is expressed
as the solution to the scalar wave equation:
1
v
2

2
P
t
2
=

2
P
z
2
, (B 106)
where P(t, z) is the plane wave represented by the 1-D seismogram a CMP-stacked trace, t
is the traveltime, z is the depth variable and v is the wave velocity. We shall assume that the
wave velocity is constant.
To solve equation (B-106), rst, Fourier transform in the time direction:

2
v
2
P =

2
P
z
2
, (B 107)
where P(, z) is the Fourier tramsform of the waveeld P(t, z), and is the angular frequency.
The upcoming wave solution is then given by
P(, z) = P(, z = 0) exp(i

v
z). (B 108)
To include amplitude decay in wave propagation in a medium with intrinsic attenuation,
the wave velocity is dened as a complex variable:
v = + i. (B 109)
Substitute equation (B-109) into equation (B-108) to get
P(, z) = P(, z = 0) exp(i

+ i
z). (B 110)
By simple algebra, rewrite equation (B-110) as follows:
P(, z)=P(, z = 0) exp(i

2
+
2
z) exp(

2
+
2
z). (B 111)
For most rocks, the assumption that is much smaller than can be made. As a result, equation
(B-111) can be simplied as follows:
P(, z) = P(, z = 0) exp(i

z) exp(

2
z). (B 112)
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
Deconvolution 267
FIG. B-4. Surface-consistent deconvolution applied to eld data: autocorrelograms of (a) the source
term, and (b) the receiver term as in equation (B-94), (c) conventional trace-by-trace prestack decon-
volution, and (d) surface-consistent deconvolution using the autocorrelation estimates as in (a) and (b)
(Analysis by Duane Dopkin).
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
268 Seismic Data Analysis
Now, dene a vertical time variable equivalent to the depth variable z via z = , and rewrite
equation (B-112):
P(, ) = P(, = 0) exp(i) exp(

). (B 113)
Assume an attenuation constant Q that is independent of frequency (Kjartansson, 1979):
1
2Q
=

, (B 114)
and substitute into equation (B-113) to obtain
P(, ) = P(, = 0) exp(i) exp(

2Q
). (B 115)
Note from equation (B-115) that the higher the frequency, the greater the attenuation.
For a nondissipative medium, = 0; hence, equation (B-114) states that Q is innite. As a
result, equation (B-115) takes the special form
P(, ) = P(, = 0) exp(i). (B 116)
The amplitude spectrum of the inverse Q lter is thus given by the exponential scaling function
A(, ) = exp(

2Q
). (B 117a)
The phase spectrum can either be set to zero, or more appropriately, assumed to be minimum-
phase. In the latter case, it can be computed by taking the Hilbert transform of the amplitude
spectrum given by equation (B-117a) (Section B.4):
(, ) = H{A(, )}, (B 117b)
where H represents the Hilbert transform.
By combining the amplitude and phase spectra given by equations (B-117a,b), we dene
the minimum-phase inverse Q lter as
W(, ) = A(, ) exp{i(, )}. (B 118)
The inverse Q ltering equation (B-115) now takes the form
P(, ) = P(, = 0) exp(i)W(, ). (B 119)
Note that the time variable t is associated with the input trace P(t, = 0) and the time variable
is associated with the output trace P(t = 0, ) after the application of the inverse Q lter.
To apply the lter W(, ) to a trace P(, = 0), dene a time step and write equation
(B-119) in its recursive form:
P(, k) = P[, (k 1)] exp(i) W(, ), (B 120)
where k = 1, 2, . . . , n, with n number of time steps (number of samples in the input trace).
Equation (B-120) can now be used to describe a procedure for inverse Q ltering:
(a) Fourier transform the input trace P(t, = 0) to obtain the complex transform function
P(, = 0).
(b) Dene a time step and apply the linear phase shift to P(, = 0) by multiplying with
the exponential exp(i).
(c) Specify a constant Q and apply the inverse Q lter given by equation (B-117a,b) that
represents the inverse Q lter.
(d) Repeat steps (a), (b), and (c) for all frequencies.
(e) Sum over all frequencies to obtain the inverse Q-ltered waveeld at time step given by
P(t = 0, ).
(f) Repeat step (d) for all time steps k, k = 1, 2, . . . , n, to obtain the inverse Q-ltered
waveeld P(t = 0, ) at all times .
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
Deconvolution 269
FIG. B-5. Inverse Q ltering applied to eld data (Saatcilar, 1996): Portion of a stacked section with
(a) no deconvolution, (b) inverse Q ltering, (c) inverse Q ltering followed by deconvolution, and (d)
deconvolution, only; (e) average amplitude spectrum of the data shown in (a); (f) average amplitude
spectrum of the data shown in (b); (g) average amplitude spectrum of the data shown in (c).
Inverse Q ltering often is applied to data using a constant Q factor. An ecient scheme
for a vertically varying Q factor is described by Hargreaves and Calvert (1991).
Figure B-5 shows an example of inverse Q ltering applied to eld data. Compare the
stacked sections and the average amplitude spectra with no deconvolution, inverse Q ltering,
inverse Q ltering followed by deconvolution and deconvolution, only. An inverse Q lter restores
the high-frequency components of signal subjected to intrinsic attenuation by the propagation
medium. The deconvolution that follows the inverse Q ltering then easily attens the spectrum
within the passband.
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/
270 Seismic Data Analysis
REFERENCES
Backus, M. M., 1959, Water reverberations: Their nature and elimination: Geophysics, 24, 233-261.
Cambois, G. and Stoa, P. L., 1992, Surface-consistent deconvolution in the log-Fourier domain:
Geophysics, 57, 823-840.
Cary, P. W. and Lorentz, G. A., 1993, Four-component surface-consistent deconvolution: Geo-
physics, 58, 383-392.
Claerbout, J. F., 1976, Fundamentals of geophysical data processing: McGraw-Hill Book Co.
Gibson, B. and Larner, K. L., 1982, Comparison of spectral attening techniques: unpublished
technical document, Western Geophysical Company.
Goupillaud, P., 1961, An approach to inverse ltering of near-surface layer eects from seismic
records: Geophysics, 26, 754-760.
Hale, I. D., 1982, Q-adaptive deconvolution: Stanford Expl. Proj., Rep. No. 30, 133-158.
Hargreaves, N. D. and Calvert, A. J., 1991, Inverse Q ltering by Fourier transform: Geophysics,
56, 519-527.
Kjartansson, E., 1979, Constant Q-wave propagation and attenuation: J. Geophys. Res., 84, 4737-
4748.
Levin, S. A., 1989, Surface-consistent deconvolution: Geophysics, 54, 1123-1133.
Lines, L. R. and Treitel, S., 1984, Tutorial: A review of least-squares inversion and its application
to geophysical problems: Geophys. Prosp., 32, 159-186.
Morley, L. and Claerbout, J. F., 1983, Predictive deconvolution in shot-receiver space: Geophysics,
48, 515-531.
Peacock, K. L. and Treitel, S., 1969, Predictive deconvolution theory and practice: Geophysics,
34, 155-169.
Ristow, D. and Jurczyk, D., 1975, Vibroseis deconvolution: Geophys. Prosp., 23, 363-379.
Robinson, E. A. and Treitel, S., 1980, Geophysical signal analysis: Prentice-Hall Book Co.
Saatcilar, R., 1996, An algorithm for Q-ltering: J. Seis. Expl., 5, 157-168.
Taner, M. T. and Coburn, K., 1981, Surface-consistent deconvolution: Presented at the 51st Ann.
Internat. Mtg., Soc. Expl. Geophys.
Taner, M. T. and Koehler, F., Surface-consistent corrections: Geophysics, 46, 17-22.
Treitel, S. and Robinson, E. A., 1966, The design of high-resolution lters: Inst. Electr. Electron.
Eng., GE-4, 1.
Walden, A. T. and Hosken, J. W. J., 1984, An investigation of the spectral properties of primary
reections coecients: Presented at the 46th Ann. Mtg. Eur. Assoc. Expl. Geophys.
Waters, K. H., 1981, Reection seismology: Second edition, John Wiley & Sons.
Yilmaz, O., 1974, The problems of resolution and reverberations in reection seismology: J. Geo-
phys. Soc. Turkey, 5, 2.
D
o
w
n
l
o
a
d
e
d

0
1
/
0
5
/
1
3

t
o

1
9
2
.
1
5
9
.
1
0
6
.
2
0
0
.

R
e
d
i
s
t
r
i
b
u
t
i
o
n

s
u
b
j
e
c
t

t
o

S
E
G

l
i
c
e
n
s
e

o
r

c
o
p
y
r
i
g
h
t
;

s
e
e

T
e
r
m
s

o
f

U
s
e

a
t

h
t
t
p
:
/
/
l
i
b
r
a
r
y
.
s
e
g
.
o
r
g
/

Você também pode gostar