Sie sind auf Seite 1von 35

An Introduction to Seismic Interpretation

Chapter 3: Acquisition and Processing

About this Disc

Disc Contents (PDF format)

Show/Hide Bookmarks

Previous Page | Next Page

Search

Print

Close

Navigation
Show Hide

3.1 Introduction
Overview

This chapter reviews aspects of seismic data acquisition and processing that are relevant to a seismic
interpreter. These topics are covered extensively by other authors and, because the focus of this text
is on interpretation, the interested reader is referred elsewhere (see below) for in-depth treatment.
The primary purposes of this chapter are to:
1. Briefly explain how seismic data are acquired and processed, including a short
discussion of aspects of 2-D and 3-D seismic survey design.
2. Ensure the reader understands that seismic data acquisition and processing parameters
will have a significant impact on the interpretability of the data (i.e. the appearance of
structural and stratigraphic features).
The chapter ends with a discussion of coherency (also referred to as semblance and other names)
processing, a processing step that is, at least in the petroleum industry, almost routinely applied to
3-D seismic data in order to enhance faults and stratigraphic features.
There is a consensus that a seismic interpretation begins at the survey design phase. As described
below, the target depth and dimensions, structural dip, rate at which high frequencies are attenuated
with depth, and other factors will influence survey design. If these parameters are improperly known
or planned for, the data acquisition effort is unlikely to yield useful results. Similarly, although
data processing may appear to have a cookbook character (follow the recipe for predictable results
every time), it is in fact full of interpretive choices made by the data processors. It is possible
to process seismic data such that they are optimized for some types of analyses but not others.
Although there may be cases for which processing can minimize problems caused by survey design
flaws, no amount of processing can make data interpretable if the survey design is too flawed, or the
acquisition is poorly executed.
Formerly it was common for seismic acquisition, processing and interpretation to be undertaken
Page 1

About this Disc

Disc Contents (PDF format)

Show/Hide Bookmarks

Previous Page | Next Page

Search

Print

Close

Navigation

An Introduction to Seismic Interpretation Chapter Three Acquisition and Processing

Show Hide

by separate groups, with little communication between those groups1. For example, the people
processing the seismic data may not have known what the data were to be used for, and the
interpreters may have had little knowledge of how the data were acquired or processed. This is not
a recipe for success. Seismic acquisition, processing and interpretation groups need to work closely
together in order to ensure that: 1) the data quality is optimized for the interpretation task at hand,
and 2) the interpreters understand the limitations of the data being analyzed.
As discussed below, seismic data collected in the field need to be processed to generate images that
come close to representing the subsurface stratigraphy and structure. Seismic processing has several
objectives that include:

Enhancing signal-to-noise ratios

Deriving subsurface velocity information

Deriving geometrically accurate images of the subsurface

Enhancing resolution

The seismic processing canon is Yilmaz (2001), a two-volume publication that presents an extensive,
mathematically focused treatment of the topic but also shows many real-data examples of the
effects of different processing steps on the appearance of seismic data. Duncans (1992) summary
is much shorter and somewhat dated, but still a useful starting point. General aspects of seismic
data acquisition were covered by Evans (1997) and, for marine surveying, Dessler (1992). Further
discussions of acquisition and processing were presented by Sheriff and Geldart (1995), Henry
(1997), Gadallah and Fisher (2005), Veeken (2007) and others. Aspects of data acquisition and
processing particular to 3-D seismic data were presented by Hardage (1997), Cordsen et al. (2000),
Galbraith (2001), Vermeer (2002) and Liner (2004). Sheriff (2002) presented definitions and short
explanations of many terms used in seismic data acquisition and processing. Mosher and Simpkin
(1999) reviewed sources and techniques for the acquisition of high-resolution marine seismic
profiles. Steeples and Miller (1998), Burger et al. (2006) and Schuck and Lange (2008) discussed
aspects of shallow land-based reflection profiling.

Stacking
We generally need or want to increase the signal-to-noise ratio of our seismic data in order to better
define structural and stratigraphic features, and rock properties. Recall from CHAPTER 2 how the
seismic method works. Very simplistically, we make a bang at the surface, the sound travels down
through the subsurface, reflects off various horizons, and we record the energy that was reflected
to the surface (FIGURE 2.1). Those reflections can be weak, depending on the strength of the
seismic source and the distance traveled by the sound. Furthermore, in addition to recording the
reflections, we also record noise. For example, at sea we will record acoustic energy generated by
breaking waves and the engines of the survey ship, in addition to the reflected energy. On land, our
receivers might record acoustic energy generated by vehicles driving by our survey site or electrical
noise generated by power lines passing overhead. Whatever its source, noise will be present in
our data and the noise can be quite strong compared to the reflections that come from horizons
several kilometers below the surface. The ratio of the amplitude (strength) of the reflections to the
__________________________________________________________________
1

This is, unfortunately, still common practice in some organizations.

Page 2

About this Disc

Disc Contents (PDF format)

Show/Hide Bookmarks

Previous Page | Next Page

Search

Print

Close

Navigation
Show Hide

An Introduction to Seismic Interpretation Chapter Three Acquisition and Processing


amplitude of the noise is known as the signal-to-noise ratio.
One way of increasing the signal-to-noise ratio is to make
more than one recording, and then to combine those
recordings together through a procedure called stacking.
FIGURE 3.1 illustrates how stacking works. At the far
left we have a signal, perhaps a reflection from the top of a
sandstone bed. There is no noise in this trace. The middle
of that figure shows the same reflection on several traces,
each of which has some random noise added. The noise
consists of a random series of positive and negative values
on each trace. For this example, the strength of the noise
is approximately equal to the strength of the signal, and the
location of the top of the sandstone is not readily apparent.
However, good things happen if we stack (add) several
of these traces together. As shown at right, the positive
amplitudes associated with the reflection combine together
to form a stronger signal that stands out above the noise.
If the noise above and below the level of that reflection is
truly random on each pre-stack recording, it should tend to
cancel out as more and more traces are stacked together.
The traces shown at right illustrate what our result might
be if we stack 5, 10 or 20 traces together. The number
of traces we stack together is known as the stacking fold,
also known simply as the fold or (less simply) as the
multiplicity. Our reflection in FIGURE 3.1 is difficult to
distinguish from noise when the stacking fold is 5, but it
begins to be visible above the noise when the fold is 10,
and the reflection is readily apparent when we stack 20
traces together. The stacking process clearly increases the
signal-to-noise ratio, and theoretically the signal-to-noise
ratio increase is given by N f where Nf is the stacking
fold (FIGURE 3.2).
Although stacking increases the signal-to-noise ratio, note
that the relationship is not linear. The greatest increase in
data quality for an increase in stacking fold (i.e. the steepest
part of the curve in FIGURE 3.2) is at relatively low fold.
The slope of the curve becomes lower as the fold increases.
Furthermore, the relationship between signal-to-noise ratio
and data interpretability (evaluated subjectively) is similar
to the relationship between stacking fold and signal-tonoise ratio. As a consequence, there are diminishing returns
as we increase the stacking fold. This relationship is
important because, from a practical perspective, increasing

Page 3

Back to Chapter

Figure 3.1: Stacking to improve signal-to-noise ratio

The trace at left is a noise-free signal, perhaps the reflection from the top of a bed.
In the middle, the signal is contaminated by random noise on each of the traces,
with the strength of the noise being approximately equal to the strength of the
signal. It is difficult to distinguish that there is indeed a signal. The traces at right
show the effect of stacking (adding) different numbers of noisy traces together. The
signal is still difficult to see when the stacking fold is five, but it becomes clearer for
stacking folds of 10 or higher. The more traces that are stacked together, the more
the random noise will tend to cancel itself out, making the signal easier to detect.

FIGURE 3.1:
Stacking to improve signalto-noise ratio
Back to Chapter

Figure 3.2: Theoretical increases in signal-to-noise


ratio as the stacking fold is increased

FIGURE 3.2:
Theoretical increases in
signal-to-noise ratio as the
stacking fold is increased
Back to Chapter

Figure 3.3: Reflections from a marine seismic shot

We are interested in primary reflections (black line)


that travel directly from the source, to an impedance
boundary, and back up to a receiver. Other raypaths
are possible and will be recorded by receivers at the
surface, including short-period multiples (red and
blue dotted lines) and long-period multiples (e.g.,
dashed green line). Eliminating multiples will be one
of the goals of seismic processing.

FIGURE 3.3:
Reflections from a marine
seismic shot

About this Disc

Disc Contents (PDF format)

Show/Hide Bookmarks

Previous Page | Next Page

Search

Print

Close

Navigation

An Introduction to Seismic Interpretation Chapter Three Acquisition and Processing

Show Hide

the stacking fold will entail increasing survey cost.


Another problem with our simplistic presentation of the seismic method from CHAPTER 2 is that
the energy does not always follow simple paths from the source to the top of a reflecting horizon and
back to the receiver, as might be implied by inspection of the figures in that chapter. FIGURE 3.3
shows various raypaths that energy might take as it travels from a source (at left) to receivers (at
right). We are interested in the shortest (black) raypath that travels from the surface down to the
top of our horizon and is reflected back up to the surface. This is known as a primary reflection.
Other raypaths have the sound taking more complicated paths, bouncing back up and down more
than once. These reflections are known as multiples. In some cases multiples can be quite strong
and they can obscure the primary reflections that we are interested in. Removing multiples will be
another objective of seismic processing.

3.2 Sources and Receivers


Sources
The type of source used to collect seismic data will depend first on whether the data are being
collected offshore or on land. Additional considerations (Evans, 1997) include:

Required penetration. All else being equal, the deeper the imaging target, the more
energy will be needed to image it.

Required bandwidth. Calculations such as those presented in SECTION 2.2 can


be used to determine what range of frequencies will be needed in order to detect or
resolve targets of a particular thickness, or to define the Fresnel Zone. Commonly
there is a trade off between penetration and frequency higher energy sources (that
would provide better penetration) tend to generate lower frequencies than lower
energy (i.e., smaller) sources.

Signal-to-noise characteristics. Surface and noise characteristics vary from place to


place, and a particular type of source may be needed to account for problems. For
example, a vibroseis source (see below) may produce a weak signal in an area of
loose sand because the vibrations are poorly transferred through the loose sand into
the underlying rock. Dynamite may be preferred in these cases, especially if the
dynamite charges can be placed down a shot hole at a level below the loose sand.

Environmental conditions. Environmental or safety requirements may prevent a


particular type of source from being used. For example, dynamite commonly cannot
be used in populated areas. Conversely, vibroseis trucks might do damage to fragile
vegetation and so a dynamite source could be preferred in that area. In marine areas
environmental assessment reports will need to be undertaken prior to surveying in
order to minimize potential damage to marine fauna.

Availability and cost. Seismic crews or particular types of equipment may not always
be available. This can be particularly true of areas far removed from traditional
petroleum exploration and development activities. Some seasons are preferred
for seismic data acquisition. For example, seismic data in northern Canada is
traditionally collected in the winter, when frozen ground makes these muskeg areas
accessible to vehicular traffic. Crews may wish to avoid marine seismic acquisition

Page 4

About this Disc

Disc Contents (PDF format)

Show/Hide Bookmarks

Previous Page | Next Page

Search

Print

Close

Navigation

An Introduction to Seismic Interpretation Chapter Three Acquisition and Processing


during traditionally stormy seasons offshore.
An ideal seismic source should be able to generate a repeatable pulse of known frequency, phase,
and other characteristics.
Most land sources can be included in one of three categories:

Weight drops. These include sledgehammers striking a metal plate on the ground,
and weights (e.g., leather bags filled with birdshot) that are dropped from a height
of 2 m or more. These types of sources are generally used in shallow seismic work.
Burger et al. (2006) suggested that under optimal conditions sledgehammer sources
can be used to detect the contact between overburden and bedrock at depths up to 50
m. Gendzwill et al. (1994) used a sledgehammer source in a mine dug into a granitic
batholith and were able to image fractures and other features over 200 m below the
mine floor.

Explosive sources include dynamite and firearms. Dynamite charges are commonly
used onshore in the petroleum industry. They need to be buried, both for safety
reasons and to ensure that the blast energy is effectively transmitted to the ground,
rather than dissipated as an air blast. Shotguns (FIGURE 3.4A) and other firearms
are sometimes used in shallow seismic work (e.g., engineering and hydrogeology
studies), and they provide more energy than sledgehammer sources. Seeber and
Steeples (1986) described the use of a .50-caliber machine gun as a seismic source
that produced frequencies in the range of 30 to 170 Hz. Even automobile spark plugs
have been tested as seismic sources for very shallow surveying; they generate high
frequencies but low energy levels (Don Steeples, personal communication, 2008).

Vibratory sources include Vibroseis technology (commonly for petroleum industry


applications) and a Mini-Sosie, a rammer type device similar to devices used on
building sites to compact the earth. Vibroseis technology puts a signal with a known
range of frequencies into the earth. The reflected energy is recorded at the surface
and the input signal is mathematically removed to leave a trace that ideally shows the
earths reflectivity (e.g., Sheriff and Geldart, 1995; Gadallah and Fisher, 2005). The
signal is generated by a truck-mounted mass that can weigh a few tens of tonnes. The
mass is made to vibrate up and down on a base plate, thereby transferring the energy
to the ground. The period over which the mass vibrates is known as the sweep. The
rate at which the mass vibrates up and down, i.e. the frequency of the source signal,
is carefully controlled and changes over the length of the sweep. A typical range of
frequencies might be 10 120 Hz, and the sweep length might be 10 20 seconds.
The length of the sweep, the range of frequencies generated, and other variables are
all optimized for the project at hand. As a general rule, the signal-to-noise ratio of the
stacked section can be improved by increasing the force applied (F; i.e. the weight
of the truck), the length of the sweep (L) and the bandwidth of the signal (W), and
the improvement is proportional to FLW (Evans, 1997). More than one vibroseis
truck might be used at the same time, as a source array (FIGURE 3.4B), in order
to increase the applied force. By vibrating up and down, most vibroseis trucks are
designed to generate P-waves. Shear waves can be generated by vibroseis trucks that
are designed to vibrate laterally but, for a variety of reasons, these sources are not
commonly used.
Page 5

Show Hide

About this Disc

Disc Contents (PDF format)

Show/Hide Bookmarks

Previous Page | Next Page

Search

Print

Close

Navigation
Show Hide

An Introduction to Seismic Interpretation Chapter Three Acquisition and Processing


Weight drops and explosive sources are known as
impulse sources, wherein the energy is generated
and dissipated quickly. The wavelet they generate is
approximately minimum phase. The wavelet generated by
vibroseis trucks over a 10 20 second sweep is clearly not
an impulse (i.e. minimum phase).

Back to Chapter

Figure 3.4: Selected seismic sources

A) Shotgun shell source (Betsy gun)


B) Vibroseis trucks in an array (photo courtesy
Francois Gauthier)

Airguns, and their relatives known as sleeve guns, are the


most commonly used seismic source in marine surveying
for the petroleum industry. An airgun consists of a metal
cylinder into which high-pressure air is injected
(FIGURE 3.4C). On a signal from the ship, a piston
moves in the cylinder thereby allowing the pressurized air
to explode into the water column through ports on the side
of the device. This outward expansion of the air generates
a pressure pulse. The amount of energy and the range of
frequencies generated depend on the size of the airgun
(generally described in cubic inches, with commonly used
sizes varying from 5 to 300 cubic inches) and the pressure
used. Sleeve guns work on similar principles but have a
somewhat different design (Evans, 1997). Usually, at least
in the petroleum industry, more than one airgun is deployed
and fired at the same time, forming an airgun array. The
number of guns used, the tow depth (usually a few meters
below the water surface), and other variables are selected in
order to provide the needed energy and frequency content
to provide the best image. The wavelet generated by an
airgun is approximately minimum phase.
Other types of marine sources include boomers (the
pressure pulse is generated mechanically) and sparkers (the
pressure pulse is generated by an electrical current that is
discharged into the water, generating a bubble), both of
which are impulsive sources. Chirp systems are vibratory
sources, the marine equivalent of vibroseis sources
although they are not as energetic as land-based vibrators
(i.e. not as much penetration). Boomers, sparkers and
chirp systems are most commonly used in shallow-marine
surveying, and generate frequencies in the 100s to 1000s
of Hertz range. Because of their high frequency content,
boomers, sparkers and chirp systems provide much better
resolution than petroleum industry airgun seismic sources
(sometimes at the decimeter scale) but their relatively
low energy levels usually restricts their use to imaging
unconsolidated deposits. Evans (1997) and Mosher and
Simpkin (1999) describe these and other marine sources.

Page 6

C) A small (10 cubic inch) airgun

FIGURE 3.4:
Selected seismic sources

Back to Chapter

Figure 3.5: Schematic illustrations of seismic receivers

A) A P-wave geophone. Vertical


ground motion caused by upgoing P-wave reflections cause a
coil to move through a magnetic
field, thereby generating an
electrical current that is sent to a
recording truck.
B) A hydrophone. P-wave
reflections traveling through the
water act as pressure pulses that
squeeze a piezoelectric crystal,
thereby generating an electrical
signal that is relayed to the ship.

FIGURE 3.5:
Schematic illustrations of
seismic receivers

Back to Chapter

Animation 5: Various types of equipment and activities associated


with both marine and land-based seismic acquisition

Courtesy Global Geophysical Services

ANIMATION 5:
Various types of equipment
and activities associated with
both marine and land-based
seismic acquisition

About this Disc

Disc Contents (PDF format)

Show/Hide Bookmarks

Previous Page | Next Page

Search

Print

Close

Navigation

An Introduction to Seismic Interpretation Chapter Three Acquisition and Processing

Show Hide

ANIMATION 5 (courtesy Global Geophysical Services) shows various types of equipment and
activities associated with both marine and land-based seismic acquisition. It shows seismic crews
engaged in various types of activities, including drilling shot holes (for dynamite), collecting
airgun seismic data with relatively small vessels, having safety meetings, surveying the locations of
geophones and shotpoints for land acquisition, and generating seismic waves with vibroseis trucks.

Receivers
Like the choice of sources, the type of receiver used depends on whether the seismic data is being
collected on land or at sea.
On land, the primary type of receiver is the geophone. A geophone is a motion detector, and a
schematic geophone design is shown in FIGURE 3.5A. The spike of a geophone is planted into
the ground to ensure good coupling between the ground motion and the device itself. Inside the
geophone is a hollow magnet that is attached to the body of the geophone. Reflected P-waves cause
the ground surface, and therefore the geophone and the magnet, to vibrate up and down. A wire
coil is suspended within the opening of the magnet. Because of inertia, the wire coil will tend not
to move at the same rate as the geophone housing. This differential movement causes the coil to
pass through the magnetic field generated by the magnet and therefore, in accordance with Faradays
Law, generating an electrical current in the coil. This electrical signal is the recorded seismic trace,
which is then transmitted to the recording truck.
At least in most petroleum applications, more than one geophone is planted at a receiver location,
forming what is known as a geophone array. Designing the geophone array (e.g., should they be
arranged in a line, in a cross or some other shape?) can be important. The location of each geophone
group needs to be accurately known through surveying. The recordings of each geophone array are
stacked prior to recording in an effort to enhance the signal-to-noise ratio. With newly designed
Q-Technology (by WesternGeco) all active geophones or hydrophones (perhaps more than 30,000
for a single shot on land) are recorded separately. This acquisition effort allows all sorts of to noise
be more effectively removed, preserves signal fidelity and high frequencies in the pre-stack data.
Shabrawi et al (2005) described the first application of this technology to improve seismic imaging
of a carbonate reservoir in Kuwait.
When the objective is to record shear waves, the geophone needs to be designed to detect horizontal
ground motions. FIGURE 3.6 compares the vertical motions recorded by P-wave geophones to the
horizontal motions (in two mutually orthogonal directions) measured by S-wave geophones. Threecomponent geophones have two S-wave detectors arranged at right angles and a P-wave detector.
New geophones have been designed that capture the full wavefield motion generated as the reflected
energy returns to the surface. Subsequent processing is used to decompose this energy into the
constituent P-wave and S-wave signals.
During normal land-based operations, the geophones are planted into the ground and left there for
most of the acquisition program. Only the source is moved, although the receivers can be selectively
turned on or off depending on whether they are needed to record the reflections coming from
below. Manually planting and removing the geophones can be a time-consuming, and therefore
costly, process. One innovative design, used to date in shallow geophysical work, involves a land
Page 7

About this Disc

Disc Contents (PDF format)

Show/Hide Bookmarks

Previous Page | Next Page

Search

Print

Close

Navigation
Show Hide

An Introduction to Seismic Interpretation Chapter Three Acquisition and Processing


streamer that is towed behind a vehicle (e.g., van der Veen
and Green, 1998; FIGURE 3.7).
The geophones in this case have flat bottoms and are
dragged along the ground as the truck moves, rather than
being planted in the ground via spikes. Gimbals within
the streamer allow the geophones to maintain a vertical
orientation. This setup allows several kilometers of data to
be acquired in a day, rather than a few hundred meters.
Hydrophones are traditionally used to record seismic
reflections at sea. A hydrophone is a pressure sensor. The
reflected energy generates a pressure pulse as it travels
through the water column. A piezoelectric crystal converts
that pressure into an electrical signal the voltage of which
is proportional to the pressure. FIGURE 3.5B shows a
schematic representation of a hydrophone.
Hydrophones are enclosed in a streamer, a neutrally
buoyant cable that is towed behind the ship. The depth at
which the streamer is towed (typically 10 m), the length of
the streamer, and the number of hydrophone groups are all
selected on a survey-by-survey basis to optimize the final
image. The location of the streamer needs to be monitored
during survey operations because waves and ocean currents
can push the streamer sideways, a problem known as
cable feathering. Monitoring the cable location can be a
technological challenge because some streamers are over
10 km in length. Gadallah and Fisher (2005) discussed
how the streamer location is monitored during marine
seismic profiling.
In some cases it might be more useful to have the receivers
directly on the seafloor. This might be the case when
collecting S-wave seismic (CHAPTER 8), in areas where
production platforms obstruct survey operations, in water
too shallow for ship traffic, or if rough seas will cause too
many problems for streamer deployment. Two options are
possible.

Ocean-bottom cables (OBC) are laid out


on the seafloor during survey operations.
Receivers, which may comprise P-wave
and two S-wave geophones as well as a
hydrophone (4-component surveying), are
located at fixed distances along the cable.
The cable is attached to a ship that records
the reflected energy. Like land-based

Page 8

Back to Chapter

Figure 3.6: Comparison of the vertical motions recorded by P-wave


geophones to the horizontal motions (in two mutually
orthogonal directions) measured by S-wave geophones

A) P-wave reflections cause vertical ground motions that are detected


by conventional geophones. Shear waves travel as horizontal
motions that are sensed by geophones that detect mutually
orthogonal horizontal ground motions (B and C).

FIGURE 3.6:
Comparison of the vertical
motions recorded by P-wave
geophones to the horizontal
motions (in two mutually
orthogonal directions)
measured by S-wave
geophones
Back to Chapter

Figure 3.7: A small vibroseis truck towing a land streamer, a series of


geophones in a cable that is dragged along the ground surface

From Cummings and Russell, 2007.

FIGURE 3.7:
A small vibroseis truck
towing a land streamer, a
series of geophones in a cable
that is dragged along the
ground surface

About this Disc

Disc Contents (PDF format)

Show/Hide Bookmarks

Previous Page | Next Page

Search

Print

Close

Navigation

An Introduction to Seismic Interpretation Chapter Three Acquisition and Processing

Show Hide

surveying, the location of each receiver group (x, y and z coordinates) needs to be
known precisely.

Ocean-bottom seismometers (OBS) are self-contained devices that are placed on


the seafloor to record reflected seismic energy. They may contain P-wave and
S-wave sensors and recordings are stored in solid-state memory within the device.
The recordings are downloaded from the device when it is retrieved. The cost of
deploying and retrieving OBS systems is high, and therefore they tend to be used to
solve special problems that warrant the cost.

3.3 2-D Acquisition The Common Midpoint Method


Data Acquisition, Normal Moveout and Stacking
Recall from SECTION 3.1 that we typically want to record reflections from a horizon more than
once, and then stack those recordings, in order to improve the signal-to-noise ratio. In principle
we could have a single source and a single receiver, install them at a particular location and make
multiple recordings at that location. We could then move our source and receiver and repeat the
exercise at many locations in order to generate a 2-D seismic line or a 3-D seismic profile. This
would be an ineffective (and therefore expensive) way of collecting seismic data.
A simplified 2-D seismic acquisition geometry (on land) is shown in FIGURE 3.8. The receivers
(blue spikes) are strung out in a line away from the source location (red dot). After making the bang,
the acoustic energy spreads out away from the source location in all directions. Raypaths are shown
for energy reflecting from a horizon at the base of the image, and traveling back up to four receiver
locations. The horizontal distance between the source and each of the receiver locations is known as
the offset. The nearest offset in the highly simplified geometry shown in the figure is 60 m, and the
farthest is 240 m. The distance between the nearest and farthest offset is known as the spread length.
Only four receiver groups are shown in this simple example, but in a typical petroleum-industry
seismic survey many more receiver locations (commonly over 100) would be used.
In FIGURE 3.8, both the ground surface and the horizon generating the reflection are planar and
horizontal. In this case, the raypaths give reflections from a location which, if projected up to
the surface, would be mid-way between the source and the associated receiver, i.e. a midpoint.
The spacing between receivers is constant (60 m). The distance between midpoints (30 m in this
example) will be one half the distance between our receivers.
The farthest offset should be at least equal to the depth of the primary target in order for subsequent
processing steps to perform adequately. For example, if the primary target is at 5 km depth, we
need to ensure that the farthest offset is at least 5 km (i.e. we need a spread length of at least that
magnitude). There are limits to the longest offset that can be used. For example, if the offset is too
long, the angle of incidence for rays to travel from source to receiver can exceed the critical angle
and the rays are refracted rather than reflected (CHAPTER 2). Note that the area of midpoint
coverage is one half the length of the spread (i.e. if the farthest offset is 2 km, the farthest midpoint
will be 1 km from the source).
The simple geometry shown in FIGURE 3.8 is known as an end-on geometry, where reflected
energy is recorded by a line of receivers located to one side of the source location. This geometry
Page 9

About this Disc

Disc Contents (PDF format)

Show/Hide Bookmarks

Previous Page | Next Page

Search

Print

Close

Navigation
Show Hide

An Introduction to Seismic Interpretation Chapter Three Acquisition and Processing


can be used on land, and nearly all marine 2-D seismic data
is acquired this way, with a ship pulling an airgun array
that is followed by a streamer. Another geometry used on
land is known as a split spread. In this case receivers are
placed on both sides of the source. FIGURE 3.9 compares
the end-on and split-spread geometries. Whereas the
locations of geophones and shotpoints are fixed during
land acquisition, at sea the source and receivers (at least for
streamers) are attached to a ship that is moving constantly.
Good navigation, i.e. accurately knowing the locations of
sources and receivers, is vital.
Returning to the example of FIGURE 3.8, we now move
the source and receiver locations along the line, and shoot
again2. FIGURE 3.10 shows four such shots, with the
source and receiver locations moved by a constant distance
for each shot. Four shots, each with four receivers, will
generate 16 recordings. Inspection of FIGURE 3.10 shows
that four of the source-receiver combinations, each from a
different shot, generated a reflection from the same spot,
i.e. they have a common midpoint. We now seek to isolate
those recordings from the others and stack them, as shown
in FIGURE 3.1, to produce one trace that shows what the
geology looks like at that midpoint location. However,
there is a problem. Although they provide reflections from
the same subsurface location, the offset for each of the
four recordings is different. As such, the time it takes for
the sound to travel from the source to the reflection point
and up to the receiver will vary. The longer the offset,
the longer it will take for the sound to make the journey.
Placing those recordings side by side (FIGURE 3.11A)
generates a common midpoint gather (CMP gather). It
can be seen that the reflections do not line up, and so they
cannot yet be stacked.
The increase in travel time with increase in offset is known
as normal moveout (NMO). The word normal indicates
that this is the anticipated response for a horizon that does
not dip, and moveout refers to the increase in travel time.
A plot of arrival times versus offset can be approximated
by a hyperbola (Figures 3.11A, 3.12). The shape of the
hyperbola is a function of the velocity between the surface
and the horizon that generated the reflection, and is given

Back to Chapter

Figure 3.8: Schematic illustration of raypaths generated


by a shot from a source point into a line of
receivers during 2-D seismic acquisition

All the energy is


A
aassumed to travel
down and up in a
d
plane (light blue)
p
beneath the receivers.
b
SSee text for further
description.
d
R
Redrawn with
modifications from
m
Ray (1995).
R

FIGURE 3.8:
Schematic illustration of
raypaths generated by a shot
from a source point into a
line of receivers during 2-D
seismic acquisition
Back to Chapter

Figure 3.9: Comparison of an end-on 2-D seismic


acquisition geometry (top) and a split-spread
2-D seismic acquisition geometry (below)

FIGURE 3.9:
Comparison of an end-on 2-D
seismic acquisition geometry
and a split-spread 2-D
seismic acquisition geometry
Back to Chapter

Figure 3.10: The common midpoint method

Four shots are shown at top,


each with a geometry similar
to that shown in FIGURE 3.8
but the source and receiver
locations are moved for each
shot. This geometry will
produce 16 field recordings (4
shots x 4 recordings per shot).
Four of the 16 source-receiver
combinations have a common
midpoint. Those four pre-stack
traces are collected to form a
common midpoint gather, as
shown in FIGURE 3.11, and
will be stacked to improve
the signal-to-noise ratio, as
illustrated in FIGURE 3.1.

FIGURE 3.10:
The common midpoint
method

__________________________________________________________________
2

Recall however from SECTION 3.2 that the receiver groups are usually not physically moved between shots on land.
Instead all the receivers may be laid out prior to surveying and different combinations of receiver groups are activated
for each shot.

Page 10

About this Disc

Disc Contents (PDF format)

Show/Hide Bookmarks

Previous Page | Next Page

Search

Print

Close

Navigation
Show Hide

An Introduction to Seismic Interpretation Chapter Three Acquisition and Processing


by the following equation (Dix, 1955):
2

t t0

h2
V 2 NMO

Back to Chapter

Figure 3.11: Schematic representation of the common midpoint


gather derived in Figure 3.10 before (left) and after
(right) normal moveout (NMO) corrections

[3.1]

where t is the traveltime at offset h, t0 is the zero-offset


or normal incidence traveltime (i.e. for sound going
straight down and coming straight back up), and VNMO
is the normal moveout velocity. One of the key seismic
processing steps in the common midpoint method attempts
to find the velocity that can be used to level out the
hyperbola generated by a reflection, and adjust the traces
in the CMP gather accordingly. This velocity is known
as a stacking velocity, and the corrections are known as
NMO corrections. Following the NMO corrections, the
reflections are aligned horizontally and the traces can be
stacked (FIGURE 3.11B).
EQUATION 3.1 is, very simply, designed to find the
velocity that gives the best stack (i.e. aligns reflections the
best). It is an approximation that works best for relatively
small offsets. For various types of advanced analyses and
imaging applications (e.g., amplitude-variation with offset;
CHAPTER 8) non-hyperbolic moveout (perhaps caused
by velocity anisotropy, e.g. different velocities parallel and
perpendicular to bedding) may need to be accounted for.
An obvious simplification in our discussion so far is
that there is only one horizon at depth that is producing
a reflection. In reality there will potentially be many
horizons that generate reflections and, because velocity
typically changes with depth, we will need to derive
different velocities to correct for NMO at various depths
(FIGURES 3.12, 3.13). In this way, we gain information
about how velocity changes with depth at each point in
the seismic survey where a velocity analysis is completed.
This process is known as velocity picking. The distance
between velocity control points along our 2-D line will
depend on factors such as our perception of how laterally
variable the velocity field is, and the time and money
available for processors to undertake many velocity
analyses. FIGURE 3.14 shows a 1980s vintage 2-D
seismic line that was plotted on paper. At top are a number
of velocity panels that show where stacking velocities
were derived. In this case velocities were calculated
approximately every 7 or 8 km, even though the line is

Page 11

Although the reflections all come from the same midpoint, the distances traveled by
the raypaths are different because of the differences in distance (offset) between the
sources and receivers. NMO corrections level out the reflection, making it possible to
stack the traces to produce one trace that represents the geology at the CMP location.

FIGURE 3.11:
Schematic representation of
the common midpoint gather
derived in Figure 3.10 before
and after normal moveout
(NMO) corrections
Back to Chapter

Figure 3.12: Hyperbolae approximating arrival times versus offset


Normal moveout in synthetic seismic
data. A) Geologic model showing
raypaths from source locations to
receiver locations for a hypothetical
common midpoint (CMP) gather.
Reflections are generated at three
interfaces in this four-layer model.
Note bending of raypaths at
interfaces. B) CMP gather showing
normal moveout (NMO) for the three
reflections. C) Hyperbolae have
been fit to the three reflections by
substituting the values at right into
Equation 3.1. The optimum values
have been estimated visually for
this example. Note that the VNMO
corresponds to a type of average
velocity between the surface and the
reflection-generating interface.

FIGURE 3.12:
Hyperbolae approximating
arrival times versus offset
Back to Chapter

Figure 3.13: A common midpoint gather A) before, and B) after


velocity analyses and NMO corrections

Primary reflections
form hyperbolae (red
lines) in part A that
are leveled out after
NMO corrections.
All traces in part B
will subsequently be
stacked to generate
a single trace in a
2-D seismic line or
3-D seismic volume.
From Duncan (1992).

FIGURE 3.13:
A common midpoint
gather before, and after
velocity analyses and NMO
corrections

About this Disc

Disc Contents (PDF format)

Show/Hide Bookmarks

Previous Page | Next Page

Search

Print

Close

Navigation
Show Hide

An Introduction to Seismic Interpretation Chapter Three Acquisition and Processing


nearly 150 km long and has nearly 2500 CMP locations.
Picking velocities at each CMP location in this type of
survey would not be economically feasible because of the
time (computer processing and human) required to do so.
The stacking velocities derived using EQUATION 3.1
represent the average velocity from the surface down to the
level that generated the reflection. Stacking velocities may
be derived for many levels at a common midpoint location.
Dix (1955) presented a way of calculating the interval
velocities between layers defined this way:
2

Vint

t 2Vrms 2 t1Vrms1
t 2 t1

[3.2]

Back to Chapter

Figure 3.14: Sample 1980s vintage 2-D seismic profile showing


locations of velocity analyses used for stacking

Velocity analyses are not performed at each CMP location.


Seismic data courtesy of Canada Nova Scotia Offshore
Petroleum Board.

FIGURE 3.14:
Sample 1980s vintage 2-D
seismic profile showing
locations of velocity analyses
used for stacking
Back to Chapter

where Vint is the interval velocity between layers 1 and 2, t1


and t2 are the zero-offset times calculated for layers 1 and
2, and Vrms1 and Vrms2 are the root-mean-square velocities
(approximated by the stacking velocities from
EQUATION 3.1) calculated for layers 1 and 2.
FIGURE 3.15 shows an expanded view of one of the
velocity panels from FIGURE 3.14. The three columns
represent, from left to right, the zero-offset time, the
stacking velocity at that time, and the interval velocities
between two layers. The interested reader is encouraged to
use EQUATION 3.2 and the data provided in the table to
recalculate the interval velocities shown in the table. Note
that EQUATION 3.2 was derived for situations where Vrms
increases with depth. As such, it is inappropriate for areas
which have velocity reversals.
Recall from CHAPTER 2 that raypaths bend at interfaces
where there is a change in velocity. In the simplest case,
where all horizons are planar and horizontal, and there are
no lateral velocity variations within layers, reflections from
all depths still line up at the midpoint (FIGURE 3.16).
Originally the seismic data acquisition technique we are
describing was termed the Common Depth Point (CDP)
method, and geophysicists referred to CDPs rather than
CMPs. This was because seismic data were commonly
collected to identify structure at one stratigraphic or
structural level (depth). Now, multiple targets at multiple
depths are the norm, and common usage generally refers
to common midpoints (CMP) rather than common depth
points (CDP).
Page 12

Figure 3.15: Application of the dix equation (Equation 3.2)


to calculate interval velocities
A) Enlarged view of one of the
velocity panels from FIGURE 3.14.
Upper row indicates the common
depth point number (CDP) and
corresponding shot point number
(SPN) for the velocity analysis.
The left-hand column shows the
zero-offset TWT (in milliseconds
labeled MSEC) and the middle
panel shows the stacking velocity
(an approximation of the rootmean-squared velocity VRMS
- from the surface down to the
level generating the reflection;
see EQUATION 3.1) in meters
per second (labeled MT/SEC).
The right-hand column shows
the interval velocity (meters
per second labeled MT/SEC)
between successive reflections
calculated using EQUATION 3.2.
B) Schematic representation of
root-mean-squared velocities
down to two levels indicated in
red box of part A.
C) Application of EQUATION 3.2 to
derive interval velocity.

FIGURE 3.15:
Application of the Dix
equation (Equation 3.2) to
calculate interval velocities
Back to Chapter

Figure 3.16: Reflections from all depths where all horizons are planar
and horizontal, and there are no lateral velocity
variations within layers, still line up at the midpoint

Previous images (e.g., FIGURE 3.8) showed only one reflecting horizon,
but midpoints should line up vertically for different levels if the
stratigraphy does not dip and there are no lateral velocity variations.

FIGURE 3.16:
Reflections from all depths
where all horizons are planar
and horizontal, and there are
no lateral velocity variations
within layers, still line up at
the midpoint

About this Disc

Disc Contents (PDF format)

Show/Hide Bookmarks

Previous Page | Next Page

Search

Print

Close

Navigation
Show Hide

An Introduction to Seismic Interpretation Chapter Three Acquisition and Processing


FIGURE 3.13 shows a common midpoint gather before
and after NMO corrections. Each trace in this gather
corresponds to the recording made by a single sourcereceiver combination. These traces will be stacked
together to produce a single trace in the 2-D seismic line
that, ideally, illustrates what the geology would look like
seismically if the source and receiver were both located
at the CMP location (vertical incidence). Note the shape
of the hyperbolae (marked with red lines) associated with
specific reflections in the original CMP gather and how
those reflections are leveled out after NMO corrections. If
incorrect velocities are picked, the reflections will not line
up and data quality will be degraded. FIGURE 3.17 shows
the importance of picking the correct velocities when
stacking the data. Obviously, picking the correct stacking
velocities greatly improves the image quality.
If the stratigraphy is dipping significantly, then NMO
corrections will not be able to line up the reflections from
a horizon. Dip moveout (DMO) corrections, described by
Yilmaz (2001), Liner (2004) and others, may need to be
applied in order to improve the quality of the stacked image
in these cases.
As might be expected from inspection of FIGURE 3.1,
the data quality of a stacked seismic image will depend
on the number of traces added together in the common
midpoint gather. FIGURE 3.18 shows the results of a
decimation experiment to test the impact of stacking fold
on data quality. These 2-D data were collected with a
target stacking fold of 60, and a portion of the stacked data
generated by stacking all 60 source-receiver combinations
for every CMP is shown at right. The middle image
shows the same data, but only using every other sourcereceiver combination to artificially produce a stacking fold
of 30. The image at left was generated using only every
fourth source-receiver combination in the stack to produce
an image with a stacking fold of 15. Everything else in
the three images is identical only the stacking fold is
different. Notice the improvement in reflection continuity
in the 60-fold version of the data. Clearly this version of
the data is better for defining stratigraphic features, and
it would be better for defining structural features (if they
exist) or for quantitative prediction of rock properties.
These images demonstrate that, all else being equal, higher
fold data will produce better quality images, as predicted by
FIGURE 3.2.
Page 13

Back to Chapter

Figure 3.17: Importance of correctly picking stacking velocities


on the quality of the stacked seismic image

Correct stacking velocities are used in part A, but velocities in part B have been
deliberately modified by approximately 10% to show how the image would be
degraded. From Henry (1997a). Reproduced with permission from Editions Technip.

FIGURE 3.17:
Importance of correctly
picking stacking velocities
on the quality of the stacked
seismic image
Back to Chapter

Figure 3.18: Decimation test to illustrate the importance


of stacking fold on image quality

The 2-D seismic image at right was


generated using a stacking fold
of 60. The image in the middle
was generated from the same
input data, but was decimated
before stacking so that only every
2nd source-receiver combination
was used, thereby generating a
stacking fold of 30. The image
at left was generated using
only every 4th source-receiver
combination, thereby generating
an image with a stacking fold of
15. Everything else has remained
constant. Clearly, the higher the
stacking fold the better the image
quality. From Leetaru (1995).

FIGURE 3.18:
Decimation test to illustrate
the importance of stacking
fold on image quality

About this Disc

Disc Contents (PDF format)

Show/Hide Bookmarks

Previous Page | Next Page

Search

Print

Close

Navigation

An Introduction to Seismic Interpretation Chapter Three Acquisition and Processing

Show Hide

The stacking fold needed to generate a useable or good seismic image varies from basin to
basin, and also varies according to parameters such as source energy, etc. A stacking fold of 30
might generate an acceptable image in one basin whereas a stacking fold of 120 might be needed
somewhere else to generate an equivalent image. Experience, gained from examination of existing
data, can help guide acquisition efforts. The stacking fold for a 3-D seismic survey does not have to
be as high as the fold in a 2-D survey to generate a comparable image. Krey (1987) suggested that
to obtain comparable data quality, the stacking fold of the 3-D data Nf3 can be calculated using:

Nf3 Nf2

f
100

[3.3]

where Nf2 is the 2-D stacking fold and f is the frequency of interest. Lansley (2004) suggested
however that for 3-D seismic data the important quantity is not the stacking fold but rather the trace
density, defined as the traces per square kilometer.
In addition to improvements to signal to noise ratio, another benefit of the NMO correction and
stacking combination is that it tends to remove long-period multiples. Multiples take different
paths through the subsurface than primary reflections (FIGURE 3.3). As such, NMO corrections
that cause primaries to line up horizontally do not cause multiples to line up. The energy from
multiples does not stack constructively and therefore the multiples tend to be attenuated in the
stacked 2-D data.
One disadvantage of the stacking process is that it reduces the frequency content of the seismic data
somewhat. Reflections that are imperfectly lined up will stack together to generate a broader, i.e.
lower frequency, stacked trace (FIGURE 3.19).
In some cases the sources and receivers cannot be laid out in a straight line, perhaps because the
seismic acquisition crew needs to follow a bending road. These types of 2-D acquisition geometries
are known as crooked line geometries and a variety of problems can be encountered when processing
and interpreting these data. Wu (1992) illustrated these problems using data acquired from a mining
camp, but indicated that similar issues would be encountered when collecting seismic data in any
type of environment. Similar problems can be encountered during 2-D marine surveying if cross
currents cause receiver cable drift, such that the cable does not truly form a straight line behind the
ship (a problem known as cable feathering). Nedimovi et al. (2003) studied this latter problem and
proposed a way for imaging 3-D structure in these cases.

Multi-Channel and Single-Channel Acquisition


Seismic data collected using multiple receiver locations to record reflections from the subsurface
are called multi-channel seismic (MCS) data. In some applications, especially in shallow marine
profiling, a single receiver group is used and the data are referred to as single-channel seismic (SCS)
data. This type of data is not stacked, and commonly a printout of the data is made in real time.
This instant gratification can be satisfying. Furthermore, because they are not stacked, SCS data
can have higher frequencies than MCS data. A disadvantage of SCS is that no subsurface velocity
information is gained.

Page 14

About this Disc

Disc Contents (PDF format)

Show/Hide Bookmarks

Previous Page | Next Page

Search

Print

Close

Navigation
Show Hide

An Introduction to Seismic Interpretation Chapter Three Acquisition and Processing

3.4 3-D Acquisition

Back to Chapter

Figure 3.19: The high-frequency content of the seismic data is reduced


if reflections are imperfectly aligned prior to stacking

When collecting 2-D seismic data, we arrange the source


and receivers in a line and anticipate that the sound will
travel in a plane that extends beneath that line
(FIGURE 3.8). However, in reality sound expands out in
three dimensions beneath the ground surface. As such, if
receivers are placed around the shot point we can generate
and record reflections from an area around the source.
FIGURE 3.20 shows a simple land acquisition geometry
for 3-D seismic data. In this case the source lines might
be arranged in an east-west direction and the receiver lines
are arranged at right angles to that (i.e. north-south in this
example). For each shot, receivers in a rectangular pattern
around the shotpoint are turned on, forming a recording
patch. The sound travels from the source and generates
a grid of midpoints, rather than a line of midpoints such
as that generated in 2-D work. Some simple geometric
rules apply. The distance between midpoints in the source
direction will be the distance between source locations
at the surface, and the distance between midpoints in the
receiver direction will be the distance between receiver
groups.
Following the first shot, the source is then moved to a
different shotpoint location and a different recording
patch of receivers is activated. This shot will generate a
new set of midpoints, some of which will be the same as
those generated by the previous shot. However, the offset
(distance between source and receiver) and the azimuth (the
compass direction traveled by the sound on its way from
source to receiver) will vary from shot to shot. One goal of
modern 3-D survey design is to ensure that a broad range of
azimuths and offsets contributes to each midpoint.

FIGURE 3.19:
The high-frequency content
of the seismic data is reduced
if reflections are imperfectly
aligned prior to stacking
Back to Chapter

Figure 3.20: Schematic illustration of a simple 3-D acquisition geometry on land

Source lines (red dots) are


arranged at right angles to
receiver lines. Receivers will be
active all around the source for
each shot, thereby generating
a grid of midpoints for each
shot, rather than a single line
(compare with FIGURE 3.8).
Redrawn with modifications
from Ray (1995).

FIGURE 3.20:
Schematic illustration of
a simple 3-D acquisition
geometry on land
Back to Chapter

Figure 3.21: Plan-view image of midpoints (blue dots) from a 3-D


seismic survey such as the one shown in Figure 3.20

The seismic trace at each midpoint location will be used to represent an area known as a bin,
the x-y dimensions of which are equal to the trace spacing. Binning the data in this way gives
the appearance of spatial continuity when viewing 3-D seismic data (CHAPTER 4).

In principle, and like 2-D seismic data, the CMP gathers


are stacked to produce a single trace that (ideally) portrays
the geology at that midpoint location. However, with
3-D seismic data we seek to portray the data as being
laterally continuous, i.e. no gaps between adjacent traces,
especially when viewing horizontal slices through the
data (time slices, horizon slices, etc.; see CHAPTER 4).
Accordingly, each CMP gather is treated as if it represents
an area known as a bin (FIGURE 3.21) rather than a single
point. The bin dimensions are the same as the spacing
between CMP locations. For example if the CMP spacing
Page 15

FIGURE 3.21:
Plan-view image of midpoints
(blue dots) from a 3-D
seismic survey

About this Disc

Disc Contents (PDF format)

Show/Hide Bookmarks

Previous Page | Next Page

Search

Print

Close

Navigation
Show Hide

An Introduction to Seismic Interpretation Chapter Three Acquisition and Processing


is 30 x 30 m (as per FIGURE 3.20), the bin size will be
30 x 30 m. Square bins are desirable from a software/
visualization perspective, but rectangular bins (e.g. 30 x 45
m) are also possible.
Through the years, different types of acquisition geometries
have been proposed (e.g., Galbraith, 2001; Gadallah
and Fisher, 2005) in response to different geological and
geophysical concerns. Liner (2004) discussed some of the
important aspects of 3-D seismic survey design that need
to be considered. Vermeer (2002) also discussed different
3-D seismic survey designs, and presented and compared
images from two different 3-D volumes that were generated
using two different designs (FIGURE 3.22). Comparing
the advantages and disadvantages of different acquisition
geometries is beyond the scope of this volume, but it
should be clear to the reader that survey design will have an
impact on the seismic image.

Back to Chapter

Figure 3.22: Horizontal slices through two different 3-D seismic cubes collected
over the same area but with different acquisition geometries

High amplitudes in orange/red and low amplitudes in blue. Curvilinear blue areas/lines show the locations of faults.
The acquisition geometry used for the survey on the right allowed the faults to be more sharply defined (compare
arrowed faults between the two images) and is therefore the preferred acquisition geometry in this setting.
Modified from Vermeer (2002) and reproduced with permission.

FIGURE 3.22:
Horizontal slices through two
different 3-D seismic cubes
collected over the same area
but with different acquisition
geometries
Back to Chapter

Figure 3.23: Highly simplified traditional seismic processing flow showing


the order in which selected processing steps are completed

The inset quo


quote reminds us of
the difficulty aassociated with
simplifying complex
co
workflows.

Marine 3-D seismic data acquisition differs from landbased acquisition. In early programs, a ship towed an
airgun array and two streamers. Each shot generated two
lines of midpoints, one for each streamer. As the ship
sailed to the new shotpoint location, new shots would
generate new midpoints, some of which would be the
same as for previous shots. Unlike land-based acquisition
where geophones commonly need to be planted manually
and dynamite or vibroseis trucks moved from shotpoint
to shotpoint, vessels designed for 3-D seismic acquisition
continuously sail from shotpoint to shotpoint and move the
source and receivers as they sail. As such marine seismic
acquisition costs are lower than land-based surveying
and, as a consequence, trace spacing in marine surveys is
typically much closer than in land-based 3-D surveys.
Currently, two airgun arrays and two or more streamers
(perhaps as many as 12) are usually deployed behind a
ship during 3-D acquisition at sea. Liner (2004) discussed
the advantages of these recording geometries. Marinewide azimuth seismic data are proving to be particularly
beneficial for problems such as sub-salt imaging surveys.
A parallel geometry is the most commonly used marine
3-D acquisition design. The survey ship sails a series of
parallel lines. In some cases it may be desirable to use two
ships, one pulling the airgun array(s) and the other towing

Page 16

FIGURE 3.23:
Highly simplified traditional
seismic processing flow
showing the order in which
selected processing steps are
completed

About this Disc

Disc Contents (PDF format)

Show/Hide Bookmarks

Previous Page | Next Page

Search

Print

Close

Navigation

An Introduction to Seismic Interpretation Chapter Three Acquisition and Processing


the streamers. In this way it can be possible to generate midpoints below surface obstacles such as
production platforms3. Other recording geometries are possible (e.g., Vermeer, 2002; Gadallah and
Fisher, 2005).

3.5 Other Processing Steps


Normal moveout corrections and stacking are important parts of a seismic processing workflow,
but other processing steps are needed to generate seismic images that come close to representing
the subsurface geology. A highly simplified traditional processing flow is shown in FIGURE 3.23.
In this section we examine some additional processing steps that have an impact on seismic data
interpretability.

Amplitude Recovery
The acoustic pulse loses energy as it spreads away from the source (i.e. the farther one is from a
source of sound, the quieter the sound will seem). As a result, horizons in the shallow part of the
section will appear to have strong amplitudes and reflections from deeper horizons will be weaker.
Spherical divergence is a term that refers to the apparent loss of energy due to geometric spreading
of a wavefront. The energy at a point on an expanding wavefront decreases with the square of
the distance from the energy source. We need to compensate for this loss of signal amplitude as
it travels through the earth because we want the strength of a reflection to be proportional to the
acoustic impedance change, and not be affected by the depth of the reflector.
Automatic gain control (AGC) is one type of process, used in many electronics applications, that
compensates for loss of signal strength with distance. AGC is an adaptive system that attempts to
maintain a constant signal level. In the seismic processing realm it will produce a seismic image
in which reflection amplitudes will remain relatively constant with depth. Although this type of
processing can help enhance the appearance of structural and stratigraphic features deep in the
section, information about seismic amplitudes that is associated with acoustic impedance changes
is lost. As such, AGC is no longer in favor in seismic processing situations where amplitude
preservation is critical.
Other approaches exist to recover amplitudes, including application of a correction for spherical
divergence. FIGURE 3.24 shows a shot gather before and after amplitude recovery. Note how the
amplitude recovery enhances the appearance of deep reflections. These events will be visible in the
stacked data following the amplitude recovery, but would not be visible otherwise.
True-amplitude recovery is a processing approach that attempts to compensate for attenuation,
spherical divergence and other effects. The objective is to produce seismic images in which the
reflection amplitudes are directly proportional to changes in acoustic impedance (i.e. reflectivity) at
all levels of the seismic data (both before and after stacking). Compensating for energy losses due
to spherical divergence is relatively straightforward, but other factors are more problematic. Recall

__________________________________________________________________
3

This technique is known as undershooting. Undershooting is also used in land-based acquisition to collect data
beneath surface obstructions such as small lakes or land held by uncooperative landowners.Instead all the receivers
may be laid out prior to surveying and different combinations of receiver groups are activated for each shot.

Page 17

Show Hide

About this Disc

Disc Contents (PDF format)

Show/Hide Bookmarks

Previous Page | Next Page

Search

Print

Close

Navigation
Show Hide

An Introduction to Seismic Interpretation Chapter Three Acquisition and Processing


from CHAPTER 2 that energy loss (attenuation) because
of absorption is related to a physical property called Q (the
quality factor). If that property can be measured (perhaps
via analysis of vertical seismic profiles see CHAPTER 5)
or estimated (sometimes estimated as 3% of the velocity
expressed in m/s; Sheriff, 2002) then attenuation
compensation (sometimes referred to as an inverse Q filter)
can be helpful for amplitude recovery.

Static Corrections
Seismic data record the two-way traveltime from the
surface to a reflecting horizon and back up to the surface.
By collecting a 2-D profile or a 3-D volume we seek to
image the structure at depth, taking into account velocity
problems like those illustrated in FIGURE 2.1. However
we will have an imaging problem if the surface over which
we move our sources and receivers is not horizontal.
FIGURE 3.25 illustrates this problem in a simple
conceptual example. In the illustration, the surface over
which the seismic data are to be collected has a hill, and
the target horizon at depth is planar and horizontal. If all
we know is the two-way traveltime, our seismic image will
show a false structure, because it takes longer for the sound
to reach the horizon beneath the hill. Clearly we have to
account for changes in surface elevation along the length
(or area) of our seismic survey. In order to do so, we need
to specify an elevation, known as the seismic reference
datum, which corresponds to 0 s two-way traveltime in our
seismic image. At sea, or in low-lying coastal areas, mean
sea level is the obvious choice. In areas on land which are
considerably above sea level, it is necessary to arbitrarily
pick a seismic reference datum.
Changes in ground elevation are but one type of problem
(FIGURE 3.26). In some areas unconsolidated materials
in the shallow subsurface (e.g., marsh deposits, glacial
till, sand dunes) can have velocities that are much lower
than the underlying rocks. In some cases, high-porosity,
uncemented sediments in the vadose zone (not all pore
space filled with water) can have velocities of nearly 100
m/s, less than the velocity of sound in air (Schuck and
Lange, 2008). This surface layer is sometimes referred to
as the low-velocity layer or the weathered layer. Changes
in thickness of this layer, if not properly accounted for, can
cause imaging problems for the underlying strata.
Page 18

Back to Chapter

Figure 3.24: Comparison of a source gather before (left)


and after (right) amplitude recovery

Notice the improved imaging


of reflections lower in the
section. From Duncan (1992).

FIGURE 3.24:
Comparison of a source
gather before and after
amplitude recovery
Back to Chapter

Figure 3.25: Illustration of imaging problem that occurs when the surface over
which we move our sources and receivers is not horizontal

Seismic raypaths below


a hill will take different
times to reflect back
up to the surface from
a horizontal bed (top).
If changes in ground
surface elevation are
not accounted for, the
result will be a false
structure in the seismic
data (below).

FIGURE 3.25:
Illustration of imaging
problem that occurs when the
surface over which we move
our sources and receivers is
not horizontal
Back to Chapter

Figure 3.26: Statics corrections

Changes in groundsurface elevation and


changes in thickness of
the low-velocity layer
need to be accounted for
in order to adequately
image the subsurface
seismically. An arbitrary
elevation, the seismic
reference datum (SRD),
is chosen to represent 0
seconds TWT. All of these
corrections are known
as statics corrections.
Short-period statics
occur within the length
of the recording spread,
whereas long-period
statics represent changes
over distances longer
than the spread length.

FIGURE 3.26:
Statics Corrections

About this Disc

Disc Contents (PDF format)

Show/Hide Bookmarks

Previous Page | Next Page

Search

Print

Close

Navigation
Show Hide

An Introduction to Seismic Interpretation Chapter Three Acquisition and Processing


The corrections necessary to solve both of these issues,
changes in elevation and changes in thickness of the low
velocity layer (if present), are known as statics corrections.
Several different approaches can be used to deal with these
problems (refraction statics, surface-consistent statics,
residual statics, etc.). If not properly accounted for, statics
problems can leave false structures in the data and degrade
the seismic image quality (Figures 3.27, 3.28).
Although it might seem that no statics corrections are
necessary when processing marine data, this is not
necessarily the case. Changes in sea-surface elevation due
to changes in tide levels may need to be accounted for.
The depth of the streamer and the depth of the airguns also
need to be considered. Finally, seafloor relief needs to be
accounted for when working with ocean-bottom cables or
seismometers.

Back to Chapter

Figure 3.27: Common midpoint gather A) before, and B) after


statics corrections, and prior to any NMO corrections

In the upper image,


reflections from the leftcentral part of the gather
appear to be pushed down,
because those geophones
were located above a lowvelocity surface feature (e.g.,
swamp). NMO corrections
will not adequately align
these reflections, and the
quality of the stacked seismic
image will be reduced.
After the statics corrections
(below) the reflections more
closely fall along a hyperbola
that can be adjusted via NMO
corrections. They will stack
better, producing a betterquality seismic image. From
Duncan (1992).

FIGURE 3.27:
Common midpoint gather
before and after statics
corrections, and prior to any
NMO corrections
Back to Chapter

Figure 3.28: Stacked seismic image A) without and B) with statics corrections

Deconvolution
Deconvolution is a processing operation, generally applied
before stacking, that attempts to:

Shorten the wavelet that is embedded in the


seismic data

Attenuate reverberations and short-period


multiples

Produce a wavelet of known phase (usually


zero phase)

Different approaches to deconvolution are possible and


a full discussion of these methods is beyond the scope of
this text. Interested readers are referred to Yilmaz (2001),
Gadallah and Fisher (2005) and other sources. What is
important here is that deconvolution is the processing
step that attempts to produce a seismic image containing
a broad-bandwidth, zero-phase wavelet when the source
wavelet was not zero phase.
FIGURE 3.29 compares two versions of a seismic
image. The image in part A has been processed without
deconvolution whereas the image in part B shows the same
data but with deconvolution applied prior to stacking.
There are somewhat subtle, but important, differences
between the two versions of the data. In particular, look
at the middle of the image at about 2 seconds down into
the data. Three prominent peaks are visible at this level in
Page 19

Note how statics corrections


improve data quality and
remove false structures.
From Yilmaz (2001).

FIGURE 3.28:
Stacked seismic image
without and with statics
corrections
Back to Chapter

Figure 3.29: Comparison of stacked seismic section generated


A) without and B) with deconvolution prior to stacking

Stratigraphic features are


more sharply defined in
the lower image. From
Yilmaz (2001).

FIGURE 3.29:
Comparison of stacked
seismic section generated
without and with
deconvolution prior to
stacking

About this Disc

Disc Contents (PDF format)

Show/Hide Bookmarks

Previous Page | Next Page

Search

Print

Close

Navigation

An Introduction to Seismic Interpretation Chapter Three Acquisition and Processing

Show Hide

the version of the data without deconvolution, whereas only two are visible when deconvolution has
been applied. One of the peaks was probably either a reverberation or a short-period multiple and
deconvolution has removed it. The resulting image more truly represents changes in amplitude that
are due to changes in acoustic impedance (i.e. the stratigraphy) in the image. Similar changes are
visible in other parts of the image.
Henry (2001) divided most deconvolution methods into two categories: deterministic deconvolution
and statistical deconvolution. Deterministic deconvolution can be applied when the source wavelet
is known, either because it is measured or it can be modeled. In this case, the deconvolution can
produce a seismic section that truly is zero phase. Generally, the source wavelet is an unknown
and so a statistical approach is taken. Two key assumptions that need to be made when working
with statistical deconvolution are that: a) the reflectivity series is random (sometimes referred to as
having a white spectrum), and b) that the input wavelet was minimum phase. Neither of these
assumptions is usually true, and so statistical deconvolution will not produce a truly zero-phase
wavelet. Although in CHAPTER 2 we noted that the wavelets generated by airgun and dynamite
sources are approximately minimum phase, the differences between the actual character and an ideal
minimum phase wavelet are significant enough to affect the deconvolution result.
FIGURE 3.30 compares the results of two different deconvolution methods applied to the same
dataset. In part A, a statistical deconvolution has been applied that resulted in a mixed-phase wavelet
being embedded in the seismic data. In part B, deterministic deconvolution was applied to the same
dataset for purposes of comparison (the source wavelet was known). The wavelet embedded in the
statistical deconvolution result is not zero phase, whereas the wavelet embedded in the deterministic
deconvolution result is truly zero phase. The deterministic deconvolution (i.e. zero phase) image has
better definition of structural and stratigraphic features.
A consequence of the shortcomings of statistical deconvolution methods is that we are never sure
of the seismic data phase even if, in principle, the data were processed to zero phase. We need to
determine the phase of the seismic data independently, and methods for doing so are described in
CHAPTER 5 when discussing synthetic seismograms.
Deconvolution is commonly not applied in shallow seismic work (Don Steeples, Personal
Communication, 2008). Reasons include:

The reflectivity profile is not random. Commonly only 3 or 4 reflections may be real,
and the statistical basis of some deconvolution methods is not valid in these cases.

Low signal-to-noise ratios. Deconvolution does not work properly in these cases.

Non-stationary wavelets. High attenuation in unconsolidated deposits causes the


wavelet to change shape rapidly with depth.

Migration/Seismic Imaging
In the simple cases examined so far, horizons in the subsurface have been planar and horizontal. In
this case, and assuming that there are no statics-related problems at the surface, reflections are from
the midpoint between a source and a receiver. We now consider a scenario where a seismic transect
is collected above a dipping horizon.

Page 20

About this Disc

Disc Contents (PDF format)

Show/Hide Bookmarks

Previous Page | Next Page

Search

Print

Close

Navigation
Show Hide

An Introduction to Seismic Interpretation Chapter Three Acquisition and Processing


Recall from CHAPTER 2 (FIGURE 2.24B) that the angle
of incidence will be equal to the angle of reflection (1 =
2). A source and receiver combination located at Points A
and C (respectively) in FIGURE 3.31 will therefore record
a reflection from Point B on the dipping surface. Note that
Point B, if projected to the surface, does not correspond
to the midpoint between Points A and C. However, and
because we do not know the true subsurface structure, we
need to make the assumption that the reflection is coming
from a position below the midpoint. The reflection will
be displayed in the seismic image as if it was generated at
point B. In other words, the reflection will be improperly
located in space.
Let us now consider a surface that has a slightly greater
degree of complexity. Perhaps the concave-up feature on
the surface shown in FIGURE 3.32A represents a buried
syncline or a buried channel. That image shows a source
location and a series of receiver locations. Because the
surface is not planar, and because the acoustic energy
expands out in all directions from the source, there are at
least three points along the surface that are located and
oriented in such a way that they generate a reflection that
will be recorded by each of the receivers at the surface. In
other words, each source-receiver combination will record
three reflections from the one surface, and potentially none
of those reflections is generated at a midpoint. However,
the common midpoint method assumes that the reflections
are coming from a location mid-way between the source
and receiver pair. A 2-D seismic transect collected above
the surface, using multiple source and receiver locations,
would generate (after stacking) the image shown in
FIGURE 3.32B. The crossing series of reflections seen
in that image are known as a bow-tie. Clearly this seismic
image does not represent the true subsurface geometry.
Another type of seismic imaging problem that needs to be
accounted for is related to discontinuities in the subsurface.
Consider a point reflector (perhaps a buried boulder) such
as the feature shown in FIGURE 3.33A. Downgoing
energy that hits the feature will be scattered and return
to the surface along many different raypaths. If we have
receivers at many surface locations, the seismic expression
of the point reflector will be a hyperbolic series of
reflections known as a diffraction. Clearly the diffraction
does not show the true shape of the buried object. In

Page 21

Back to Chapter

Figure 3.30: Comparison of two different deconvolution


results on the same data

Seismic image on left was generated using a statistical deconvolution approach that left
a mixed-phase wavelet in the data. Seismic image on the right was generated using a
deterministic deconvolution that generated a truly zero-phase version of the data. Arrows point
to improvements in image quality in the zero-phase version of the data. From Henry (2001).

FIGURE 3.30:
Comparison of two different
deconvolution results on the
same data
Back to Chapter

Figure 3.31: A source and receiver combination located at Points A and C


recording a reflection from Point B on the dipping surface

When beds dip the reflection points are not located at midpoints because the angle of incidence
(1) is equal to the angle of reflection (2) for the dipping bed. However, and in the absence of
any other knowledge of subsurface structure, the reflection is still assumed to come from the
midpoint location. In other words, the reflection will be misplaced spatially from its true location.

FIGURE 3.31:
A source and receiver
combination located at Points
A and C recording a reflection
from Point B on the dipping
surface
Back to Chapter

Figure 3.32: Surface with a concave-up feature showing a


source location and a series of receiver locations

A) Raypaths from a source


location (red dot) to
receivers (blue) above
a concave-up feature
(perhaps a buried
channel or a syncline).
Each receiver will record
three reflections from the
surface, and generally the
reflections do not come
from midpoint locations.
B) Acquisition of a 2-D
seismic profile will
generate a seismic feature,
known as a bow tie, that
clearly does not portray
the true subsurface
geometry. The reflections
are misplaced spatially,
and need to be moved to
their true positions using a
processing step known as
migration.

FIGURE 3.32:
Surface with a concave-up
feature showing a source
location and a series of
receiver locations

About this Disc

Disc Contents (PDF format)

Show/Hide Bookmarks

Previous Page | Next Page

Search

Print

Close

Navigation
Show Hide

An Introduction to Seismic Interpretation Chapter Three Acquisition and Processing


addition to point reflectors, diffractions will be generated
at any abrupt change in physical properties such as a bed
termination at a fault or beneath an unconformity. The
apex of the diffraction shows the true location of the feature
generating the reflection.
Because of these and related problems, after stacking a
seismic image has various distortions:

Dips of reflections are decreased and their


spatial extent is increased compared to
reality

Concave-up features (synclines, etc.) appear


as bow-ties

Anticlines appear to be broader

Diffractions are present at abrupt


terminations

Back to Chapter

Figure 3.33: A single point reflector generates a


diffraction in a stacked seismic image

In this schematic
sch
example, the source
location is directly
above the buried body
energy is
and diffracted
ra
by receivers on
received b
sides.
both sides

FIGURE 3.33:
A single point reflector
generates a diffraction in a
stacked seismic image
Back to Chapter

Figure 3.34: Comparison of stacked, unmigrated seismic image


(A) with migrated version of the same data (B)

A seismic processing step known as migration is used to


correct for the imaging problems just discussed. Properly
executed, migration achieves three goals:

Repositioning reflected energy to its true


subsurface location in the two- or threedimensional space (x,time or x,y,time
respectively) of the seismic data

Shrinking the Fresnel zone. After migration,


and if the migration is properly done, the

Note the improvement


in imaging of the
synclines (note bow
ties in unmigrated
version) and anticlines
due to migration. From
Duncan (1992).

FIGURE 3.34:
Comparison of stacked,
unmigrated seismic image
with migrated version of the
same data

Fresnel zone diameter will be 4 (but see


below for a comparison of 2-D and 3-D
migration.)

Back to Chapter

Figure 3.35: Comparison of A) an unmigrated and


B) migrated seismic image of a fault

Collapsing diffractions.

The result should be a much more geologically realistic


image of the subsurface. Compared to a seismic image
produced without migration, the migrated image is an
improvement in that it allows the true geometries of
anticlines and synclines to be observed, dips are steeper,
and diffractions are eliminated. Figures 3.34 and 3.35
show the difference between migrated and unmigrated
seismic images.
A variety of mathematical methods exist for migrating
seismic data (e.g., Kirchhoff migration, finite-difference
migration, wave-equation migration, reverse time

Page 22

Note how the migration


has moved the fault
back to its true location
and improved the
imaging of adjacent
reflections. Reproduced
with permission from
Henry (1997a).

FIGURE 3.35:
Comparison of an unmigrated
and migrated seismic image
of a fault

About this Disc

Disc Contents (PDF format)

Show/Hide Bookmarks

Previous Page | Next Page

Search

Print

Close

Navigation

An Introduction to Seismic Interpretation Chapter Three Acquisition and Processing

Show Hide

migration), and a discussion of these methods is beyond the scope of this book. Interested readers
are referred to the texts listed in the Introduction to this chapter for more details on migration. Here,
and following Sheriff (2002), Liner (2004) and others, we distinguish between migration approaches
according to two principal factors:

Is the migration performed before or after stacking the data (pre- or post-stack
migration respectively)?

Does the migration properly handle lateral velocity variations and account for ray
bending (depth migration) or does it not (time migration)?

All migration methods require some sort of velocity model in order to undertake the mathematical
operation of migration. Velocity information needed to build that model can be obtained in a
variety of ways. The derivation of stacking velocities was discussed in SECTION 3.3. Checkshot
surveys, vertical seismic profiling and sonic logs are described in CHAPTER 5. Seismic traveltime
tomography uses seismic recordings to construct a velocity model that minimizes the error between
the measured and theoretical traveltimes. Different approaches are possible to tomography, and
Rawlinson et al. (2003), Accaino et al. (2005), and Lehmann (2007) discuss various applications
ranging from engineering to lithospheric imaging.
Bednar (2005) discussed the history of seismic migration. Originally, i.e. before digital recording,
migration was performed manually (on paper). Even once digital recording became commonplace
(e.g., 1960s and 1970s) much seismic data was processed and displayed without being migrated.
Subsequently, migration was applied after stacking, and stacking velocities, usually derived from
NMO corrections, were used as a first guess. These velocities were varied by perhaps 5 or 10%
to examine the effects of changing the migration velocity on the seismic image. If the migration
velocities are too low, the data are said to be under-migrated and concave-up patterns known as
migration smiles are apparent. Migrated noise in the deep part of seismic images can generate
prominent migration smiles. If the migration velocities are too high, i.e. the data are over-migrated,
concave-down patterns known as frowns may be visible. The velocity field that provided the
sharpest image and best removed the diffractions and other problems was selected as the appropriate
velocity model for migration. This approach is known as post-stack time migration (Post-STM)
because the data are migrated after stacking with the vertical axis in time.
Post-stack time migration is an acceptable, and relatively cheap, way of migrating seismic data
when velocities vary vertically in a more-or-less uniform manner, and lateral velocity variations
are minimal or non-existent (e.g., areas with a structurally undisturbed layer-cake stratigraphy).
However this type of migration is inadequate in the presence of abrupt lateral and vertical velocity
variations such as those that might be present adjacent to salt diapirs or in thrusted areas (e.g.,
Paleozoic carbonates thrust over Tertiary clastics). In these cases, raypaths (or, more correctly,
wavefronts) can follow complicated paths as they are refracted (according to Snells law, which is
not honored in time imaging) at dipping interfaces having high velocity contrasts (e.g., Figures 2.24,
2.26). Reflections do not necessarily come from common midpoints in these areas (Figures 3.36,
3.37), and stacking then migrating these reflections will not produce a clear or geometrically
accurate image.
In cases of steep dips and significant vertical and lateral velocity variations the data need to be

Page 23

About this Disc

Disc Contents (PDF format)

Show/Hide Bookmarks

Previous Page | Next Page

Search

Print

Close

Navigation
Show Hide

An Introduction to Seismic Interpretation Chapter Three Acquisition and Processing


migrated before stacking in order to define common
reflection points (CRP; also referred to as common image
points CIP). Common reflection point gathers represent
combinations of sources and receivers that all image the
same subsurface point. To identify common image points
it is necessary to track rays or wavefronts as they bend
through the subsurface according to Snells Law. Pre-stack
depth migration (Pre-SDM) is used in these cases. Two
classic settings where PSDM is critical are:

Imaging adjacent to, or beneath, salt domes,


where the Vp of the salt is significantly
different from that of the adjacent
sedimentary deposits.

Back to Chapter

Figure 3.36: Source-receiver combinations with a common midpoint

Although these two source-receiver combinations share the same midpoint, the reflections do not
come from the same subsurface locations because of the dipping bed. (Raypaths bend according
to Snells Law and the angle of incidence is equal to the angle of reflection). Stacking these two
traces will not help to generate a good-quality, or geometrically accurate, seismic image.

FIGURE 3.36:
Source-receiver combinations
with a common midpoint
Back to Chapter

Figure 3.37: Reflections below a salt body do not come from a


common midpoint

Imaging in overthrust areas, especially


where high-velocity strata are thrust over
rocks with lower velocities.

Depth migration bends rays through the subsurface, thereby


repositioning reflections to their correct locations in space.
Although it can be applied after stacking, the best results
are obtained when the data are migrated before stacking.
An accurate velocity model is needed in order to properly
characterize how rays bend. The velocity models are
constructed in a variety of ways, such as tomography. In all
cases there is a need for integration and interaction between
structural geologists (who are hopefully in a position
to make intelligent guesses about the likely subsurface
structure) and data processors (who build the velocity
models and migrate the data). Albertin et al. (2001) and
Abriel et al. (2004) illustrate depth imaging of 3-D seismic
data from the Gulf of Mexico. FIGURE 3.38 compares a
time-migrated image (part A) with a depth-migrated version
of the same seismic data (part B). Note the considerable
improvement in imaging the base and flanks of the salt
body and the sub-salt area that is visible in the depthmigrated seismic image.
Although it is beyond the scope of this text to provide
a detailed discussion or comparison of the algorithms
employed to migrate seismic data, it is worth highlighting
that different migration results can be useful for different
purposes. For example, FIGURE 3.39 compares three
different pre-stack depth migration results. The controlledbeam migration result (FIGURE 3.39C) provides better

Page 24

Raypaths for reflections from a single horizon for a common midpoint gather in the presence of a
salt body and associated structural deformation. Note that reflection points do not coincide. Inset at
lower right shows CMP gather. Reflections from this one interface do not fall along a simple hyperbola
(e.g., FIGURE 3.12) and NMO corrections followed by migration will not be adequate to generate an
acceptable image. Pre-stack depth migration is needed in this case. From Abriel et al. (2004).

FIGURE 3.37:
Reflections below a salt body
do not come from a common
midpoint

Back to Chapter

Figure 3.38: Comparison of A) post-stack time-migrated seismic image and B) pre-stack


depth migrated seismic image of a salt body in the Gulf of Mexico

Note how the pre-stack depth migration has


improved the seismic imaging adjacent to and
below the salt. Images courtesy of TGS-Nopec.

FIGURE 3.38:
Comparison of post-stack
time-migrated seismic image
and pre-stack depth migrated
seismic image of a salt body
in the Gulf of Mexico

About this Disc

Disc Contents (PDF format)

Show/Hide Bookmarks

Previous Page | Next Page

Search

Print

Close

Navigation
Show Hide

An Introduction to Seismic Interpretation Chapter Three Acquisition and Processing


structural imaging of the sub-salt area, but information
about lateral amplitude variations is suppressed with
respect to the other two images. Potentially, either
Figure 3.39A or 3.39B might be more useful for mapping
of stratigraphic features (CHAPTER 7) or detecting
hydrocarbons using amplitudes (CHAPTER 8).
Because post-stack migration methods work with seismic
data after stacking, they involve working with less
data than pre-stack methods and are therefore cheaper
(cost is proportional to the computer time used, and the
computer time used is proportional to the amount of data
manipulated). Consider a hypothetical 60-fold 2-D seismic
line that is 100 traces long. Post-stack migration would
involve migrating 100 traces whereas pre-stack migration
of the same dataset would include migrating 60 x 100 =
6000 traces. Depth migration methods more accurately
capture the physics than time migration, meaning more
calculations and further costs. It should be noted that
small discrepancies in a velocity model used for post-stack
migration will not significantly change the migrated seismic
image, however depth migration is very sensitive to errors
in interval velocities because the method performs exact
calculations of ray paths or wavefronts.

Back to Chapter

Figure 3.39: Comparison of three different pre-stack migration


algorithms: A) Kirchoff migration B) wave-equation
migration, C) controlled-beam migration

The wave-equation result in part B provides better


sub-salt imaging than the Kirchoff result in part
A. The result in part C provides better structural
imaging (more continuous reflections, sharper
defintion of faults) than either of the other results,
but lateral changes in amplitude that might be
related to hydrocarbons or other factors are
suppressed in this controlled-beam migration image.
Images courtesy CGGVeritas.

FIGURE 3.39:
Comparison of three
different pre-stack migration
algorithms: Kirchoff
migration, wave-equation
migration, and controlledbeam migration
Back to Chapter

Figure 3.40: Schematic illustration showing how sideswipe is generated

Although acoustic energy is assumed


to travel down and back up to the
surface along a raypath (red) in a
vertical plane that underlies the
source (red oval) and receiver (blue
spike), features off to the side of
that plane can generate a reflection
(sideswipe) that is also recorded by
the receiver.

Pre-stack depth migration of large petroleum-industry 3-D


seismic volumes is usually undertaken on supercomputers
or Linux clusters. Post-stack migration can be handled
by workstations or, for small datasets, even personal
computers. Additional costs include the time spent
constructing and validating velocity models. Because of
these and other considerations, pre-stack depth migration is
not yet used ubiquitously. Although it is common for timemigrated seismic data to be plotted with the vertical axis
in time and depth-migrated seismic data to be plotted with
the vertical axis in depth, in principle both methods can be
used to plot seismic data with either time or depth as the
vertical axis.

FIGURE 3.40:
Schematic illustration
showing how sideswipe is
generated

In cases where velocity anisotropy (CHAPTER 2)


is pronounced, even Pre-SDM can be inadequate.
Conventional PSDM corrects for lateral velocity
heterogeneity, but anisotropic PSDM also corrects for
velocity changes with direction (i.e., perpendicular or
parallel to bedding in shale) in the anisotropic strata.
Where pronounced, and the difference in velocity parallel

FIGURE 3.41:
Modeling experiment
undertaken by Fred Hilterman
in the early 1970s to illustrate
the appearance of sideswipe
from a carbonate buildup on
2-D seismic profiles

Page 25

Back to Chapter

Figure 3.41: Modeling experiment undertaken by Fred Hilterman in


the early 1970s to illustrate the appearance of sideswipe
from a carbonate buildup on 2-D seismic profiles
Image at top shows the outline of an
oval-shaped carbonate buildup in map
view. Immediately below is a crosssection view through the buildup
showing the simple stratigraphy used
in the modeling. At bottom are the
seismic model images corresponding to
transects collected at various distances
from the center of the buildup. The
carbonate buildup appears to be as
wide in the transect collected 500 from
the center of the buildup as it is in the
center, even though it is narrower at
that location. The buildup appears
in the seismic data at 1000 and 1500
from the center of the buildup, even
though neither of those lines crosses
it. These problems are related to
sideswipe. No migration has been
applied to these data (migration of
any kind was not routine in the early
1970s). 2-D migration could improve
the imaging of the flanks of the buildup
but could not remove the sideswipe.
(Jackson and Hilterman, 1979; cited by
Crawley Stewart, 1995. Reproduced with
permission from Hilterman).

About this Disc

Disc Contents (PDF format)

Show/Hide Bookmarks

Previous Page | Next Page

Search

Print

Close

Navigation

An Introduction to Seismic Interpretation Chapter Three Acquisition and Processing

Show Hide

to the bedding and perpendicular to the bedding can be as high as 30% (Lawton et al., 2001) the
assumption of an isotropic velocity field can lead to misplaced and poorly imaged reflections.

3-D Migration
One of the biggest pitfalls of 2-D seismic methods is that acoustic energy does not simply follow
2-D raypaths such as those shown previously. Sound expands outwards from the source as a 3-D
wavefront in the subsurface. As such, geological bodies located off to the side of the line can
generate reflections that are recorded by receivers on the line (FIGURE 3.40). These out-of-plane
reflections are known as sideswipe. Sideswipe, or seeing off to the side on a 2-D profile, needs to
be acknowledged as a potential problem when working with that type of data. Conventional 2-D
migration cannot account for this problem. Instead, and as recognized at least in the 1970s (e.g.,
French, 1974), 3-D migration can be used to eliminate sideswipe.
The image shown in FIGURE 3.41 illustrates this point. The upper part of the figure shows a map
view (top) and transect (below) through a geological model of a carbonate buildup. The locations
of six modeled 2-D seismic lines are also shown. The lower images show these modeled seismic
profiles. Note that the carbonate buildup is imaged even on lines that do not cross the feature (e.g.,
1500). That is because the sound expands out in three dimensions away from the seismic source.
Some of that sound may be reflected from points located below the line of receivers, but some of the
sound may travel out away from the line, then bounce off a surface back into the line of geophones.
There is commonly no way of knowing whether a feature on a 2-D profile lies directly under the
surface survey line or off to the side. Note that no migration was applied to the modeled seismic
profiles shown in that figure because it was not common to apply even 2-D migration in the early
1970s. As such, the flanks of the buildup are poorly imaged in the seismic profiles. Although it
would improve the seismic definition of the flanks of the buildups, 2-D seismic migration would not
solve the sideswipe problem illustrated here.
The modeling shown in FIGURE 3.41 was designed to illustrate the effects of sideswipe on the
exploration for carbonate buildups in the Michigan Basin, but has proven to be useful elsewhere.
For example, Crawley Stewart (1995) described problems encountered by companies exploring for
Pennsylvanian algal mounds in southeast Utah using 2-D seismic-based mapping. There would be
times that the seismic profiles indicated the presence of a buildup and drilling would confirm the
seismic prognosis. However, there would be times that a buildup would be clearly observed on
a 2-D seismic profile, but drilling did not find one. Only 3-D seismic, and the application of 3-D
migration, could reliably indicate the presence or absence of productive mounds. This particular
exploration problem is illustrated in FIGURE 3.42.
FIGURE 3.43 illustrates how 3-D migration eliminates sideswipe. Part A of that figure shows the
same image as shown in FIGURE 3.39, but this time with the Fresnel zone illustrated. The diameter
of the Fresnel zone is given by EQUATION 2.19. FIGURE 3.43B shows how 2-D migration

shrinks the Fresnel zone. The diameter is shrunk to 4 but only in the direction of the 2-D line. The
Fresnel zone diameter is unchanged perpendicular to the line, and so sideswipe problems remain.
Because of the way that sources and receivers image midpoints from a variety of azimuths with a

3-D acquisition geometry, the Fresnel zone is shrunk to 4 in all directions following 3-D migration
Page 26

About this Disc

Disc Contents (PDF format)

Show/Hide Bookmarks

Previous Page | Next Page

Search

Print

Close

Navigation
Show Hide

An Introduction to Seismic Interpretation Chapter Three Acquisition and Processing


and so sideswipe is eliminated (FIGURE 3.43C).
FIGURE 3.44 clearly illustrates the sideswipe problem
with another real data example. The images here are both
from the same 3-D seismic survey. However, the version of
the data in the image on the left (part A) was processed as
a 2-D seismic line (i.e. 2-D migration) whereas the version
of the data in the image on the right (part B) was processed
with 3-D migration. Note the prominent bump seen
near the center of the image in the 2-D migration version.
Perhaps this bump represents a carbonate buildup or some
other type of potential drilling target. The bump disappears
when the 3-D migration is applied. These images
demonstrate that the feature is an out-of-plane reflection
(sideswipe) similar to the one shown schematically
in Figure 3.40 or 3.43B. It is impossible to tell from
inspection of FIGURE 3.44A that the bump is actually due
to sideswipe.
All types of migration discussed previously, including prestack depth migration and post-stack time migration, can be
applied to migrate 3-D seismic data.

Processing Summary
In an ideal world, the seismic data represent geometrically
accurate images of the subsurface after processing. The
seismic traces simulate data that would have been collected
using waves that traveled vertically up and down from
the surface to the reflecting horizons and back up at the
common-midpoint location. Stacking, migration and statics
corrections are all processing steps that are intended to
produce this type of image. Multiples and other sources of
noise have been removed through deconvolution, stacking
and potentially other processing steps for noise or multiple
suppression that were discussed by Yilmaz (2001) and
Gadallah and Fisher (2005). We have a true image of the
geology, within the resolution limits of the seismic data.
Unfortunately, in practice most seismic datasets have
varying degrees of imaging problems and various types of
noise, and an experienced seismic interpreter needs to be
able to be aware of these limitations.

3.6 Survey Design


A simple conceptual framework for acquiring seismic
data is shown in FIGURE 3.45. Geologic knowledge,

Page 27

Back to Chapter

Figure 3.42: 2-D seismic transect (top) and corresponding transect from a 3-D
seismic volume (below) showing the appearance of a sideswipe

A) The 2-D image appears to


show a carbonate buildup
that is represented by a
dim-out of the Upper Ismay
reflection (a strong peak).

B) This arbitrary line from a 3-D


survey corresponds to the 2-D
seismic transect shown in part
A. The Upper Ismay reflection
does not dim here, indicating
that no carbonate buildup
is present at this location.
Reproduced with permission
from Crawley-Stewart (1995).

FIGURE 3.42:
2-D seismic transect and
corresponding transect from a
3-D seismic volume showing
the appearance of a sideswipe
Back to Chapter

Figure 3.43: 3-D migration removes sideswipe

A) The original circular Fresnel zone (blue) for a 2-D


seismic line images the dipping surface to the
left, and sideswipe is recorded.
B) Migrating the 2-D seismic data shrinks the Fresnel
zone in the direction of the 2-D line, but cannot
reduce the diameter perpendicular to the line
and sideswipe remains a problem. The Fresnel
zone is elliptical.
C) The acquisition geometry of 3-D seismic methods
allows 3-D migration that shrinks the Fresnel
zone to a small circle, eliminating sideswipe.

FIGURE 3.43:
3-D migration removes
sideswipe
Back to Chapter

Figure 3.44: Comparison of seismic transect with 2-D migration (A)


with the same transect through a 3-D volume (B)

The 2-D line was in fact extracted from the 3-D acquisition program and processed
using 2-D migration that did not remove sideswipe. Notice the presence of the
bump in part A (red arrow) that is not seen in the image with the 3-D migration.
That feature is due to sideswipe on the 2-D image. Data courtesy Western Geco.

FIGURE 3.44:
Comparison of seismic
transect with 2-D migration
with the same transect
through a 3-D volume

About this Disc

Disc Contents (PDF format)

Show/Hide Bookmarks

Previous Page | Next Page

Search

Print

Close

Navigation
Show Hide

An Introduction to Seismic Interpretation Chapter Three Acquisition and Processing


geophysical knowledge and logistical issues are all
considered when designing a seismic program, either 2-D
or 3-D. Geologic knowledge includes factors such as
the depth to the primary target, the thickness of the units
that need to be detected and mapped, and structural dip.
Geophysical knowledge includes understanding how the
seismic energy will be attenuated with depth, how best to
design source or receiver arrays to increase the signal-tonoise ratio, the ideal stacking fold, data input requirements
for certain processing or interpretation steps, and other
factors. Logistical issues include defining the availability
of seismic crews or equipment, obtaining permission
to collect data in the desired area (permitting), and,
inevitably, cost considerations. Survey design almost
always involves making compromises between what
should be done and what can be done given restrictions on
available time, money and expertise.
Some of the parameters that need to be defined when
designing a 2-D or 3-D seismic survey include:

Trace spacing in 2-D seismic data. The trace


spacing helps to define the lateral resolution
and can be a concern if a problem called
spatial aliasing is to be avoided. Spatial
aliasing arises when the trace spacing is not
close enough to unambiguously define true
reflection dip (FIGURE 3.46). Liner (2004)
illustrated examples of spatial aliasing
on real seismic data. The minimum trace
spacing needed to avoid spatial aliasing
problems can be calculated using:

V
4 F sin D

[3.4]

where X is the trace spacing, V is the


velocity, F is the maximum unaliased
frequency (Hz) and D is the structural dip
(degrees). Consider the following example.
In an area of thrusted Paleozoic carbonates,
the velocity is 5000 m/s, the expected
maximum frequency is 50 Hz and the
structural dip is 30. After substituting these
values into EQUATION 3.4, we conclude
that the minimum trace spacing needed to
avoid spatial aliasing is 50 m, indicating that
our receiver groups should not be less than
100 m apart.
Page 28

Back to Chapter

Figure 3.45: Conceptual diagram for seismic acquisition

The inset quote reminds us of the difficulties associated with simplifying complex workflows.

FIGURE 3.45:
Conceptual diagram for
seismic acquisition
Back to Chapter

Figure 3.46: Spatial aliasing

A) When seismic traces are located closely enough together,


definition of reflection dip is unambiguous.
B) If traces are too widely separated, other dip interpretations
are possible, i.e. the data are spatially aliased.

FIGURE 3.46:
Spatial aliasing

About this Disc

Disc Contents (PDF format)

Show/Hide Bookmarks

Previous Page | Next Page

Search

Print

Close

Navigation

An Introduction to Seismic Interpretation Chapter Three Acquisition and Processing

Show Hide

Bin size in 3-D seismic data. As per the discussion of 2-D seismic trace spacing, the
bin size of a 3-D survey needs to be small enough to avoid spatial aliasing problems.
The bin size also helps to define the lateral resolution of a 3-D seismic survey. One
rule of thumb is that the bin size should be small enough so that three bins will fall
into the narrowest feature that needs to be resolved. For example, if an interpretation
team needs to be able to consistently map channels that are 50 m wide, then the
50
desired maximum bin size would be 3 18 m. In reality, the true lateral resolution
of a 3-D survey is defined by whatever is greatest, 3 times the bin size or the postmigration Fresnel zone. Consider this example: A 3-D survey has a bin size of 15 m,
a dominant frequency of 50 Hz, and an interval velocity of 2400 m/s at the level of a
particular horizon. The wavelength is determined to be 48 m (using

EQUATION 2.4) and therefore the post-migration Fresnel zone ( ) will be 12 m.


4
The width of three bins (45 m) is greater than the diameter of the post-migration
Fresnel zone (12 m), and therefore the bin size determines the lateral resolution.
In some cases, particularly when imaging deep targets and bin sizes are small (i.e.
marine surveys), the post-migration Fresnel zone can be larger than three bin widths.

Size of survey. The survey should be designed so that the entire area of the target is
covered by full-fold data (FIGURE 3.47). This area is sometimes referred to as the
sweet spot and, in the petroleum industry, it should be big enough for the structural
or stratigraphic limits of a hydrocarbon reservoir to be defined. At the end of a 2-D
seismic line or around the margins of a 3-D survey will be an area that has midpoints
but the fold and range of offsets will not be adequate to properly stack or migrate the
data. This poorly imaged area, commonly trimmed off by processors and sometimes
referred to as the migration fringe, combines with the sweet spot to constitute the
image area, i.e. the area for which seismic traces are recorded. Outside of the image
area is the acquisition fringe, an area at the surface where receivers (and possibly
sources) need to be located but for which no CMP locations (i.e. no traces) will be
generated4. The dimensions of the acquisition area are a function of the depth to the
target and the structural dip.

Once data collection begins, it is advisable to monitor the acquisition effort in order to ensure
that hardware and crews are properly located and performing properly. Field records need to be
monitored to ensure that the data quality will be adequate. If not, acquisition might need to be
terminated.

3.7 Reprocessing and Post-Stack Processing


Sometimes an interpreter is asked to work on data (2-D or 3-D) that was acquired and processed
several or more years previously, and the data quality is not good. The interpreter has three options:
1. Work with the data as is. This might be the only option if the data are paper copies
and the original tapes are not accessible.

__________________________________________________________________
4

Recall from FIGURE 3.8 that the area of midpoint coverage does not extend to include the entire length of the spread.

Page 29

About this Disc

Disc Contents (PDF format)

Show/Hide Bookmarks

Previous Page | Next Page

Search

Print

Close

Navigation
Show Hide

An Introduction to Seismic Interpretation Chapter Three Acquisition and Processing


2. Reprocess the original data. This approach
seeks to take advantage of improvements in
processing capabilities with time. Perhaps
processing algorithms are available now
that were not when the data were originally
processed. This option will incur a time
delay, while the data are being reprocessed,
and additional costs.
3. Attempt some post-stack processing on the
existing seismic line(s) or volume(s).
Having the data reprocessed is the preferred option.
Improvements in processing algorithms between the
original processing and today could lead to improved
imaging. Alternatively, the data might have had a generic
processing workflow applied and special attention needs
to be paid to a particular stratigraphic horizon or structural
feature. Figures 3.48 and 3.49 show examples of how
reprocessing was used to improve seismic image quality.
Seismic reprocessing is helping to open new hydrocarbon
exploration opportunities in many parts of the world (e.g.,
Roberts et al., 2008).
In some cases the time, or money, may be missing to
send data out for reprocessing. Some workstation-based
interpretation software packages provide an interpreter
with the ability to implement certain processing steps on
fully processed (stacked and migrated) seismic data
and then see the results in near-real time. For example,
an interpreter might judge that a certain data set has
significant high-frequency noise that obscures stratigraphic
or structural features. Filtering out those high frequencies
could make the low-frequency information more readily
interpretable. At other times, the interpreter might attempt
to attenuate coherent noise that is related to acquisition and
processing (Marfurt et al., 1998).
Post-stack processing allows the interpreter to undertake
various types of processing on the data, typically with the
intent of enhancing the interpretability of the data. Various
manipulations can be applied to the seismic data (e.g.,
bandpass filtering, deconvolution, amplitude balancing,
dip filtering, trace averaging). Ogiesoba and Hart (2009)
described how post-stack processing was used to improve
fault definition in a relatively poor quality 3-D seismic

Page 30

Back to Chapter

Figure 3.47: Illustration of the sweet spot for 3-D seismic imaging

The entire seismic target should be covered by a full-fold area known as the
sweet spot. The image area includes the sweet spot and the adjacent, poorly
imaged areas that cannot be properly stacked or migrated because not enough
traces are available. The acquisition area includes areas outside of the image
area where receivers will be located, but for which no data will be collected.

FIGURE 3.47:
Illustration of the sweet spot
for 3-D seismic imaging
Back to Chapter

Figure 3.48: Examples of how reprocessing was used


to improve seismic image quality

Reprocessing data can improve


seismic image quality.
A) Transect through a 3-D
seismic cube processed in
1991.
B) Corresponding transect
through a cube that was
reprocessed in 1998.
Imaging of structural and
stratigraphic features is
much better in the lower
image, at least in the
shallow part of the section.
The main differences were
the application of dip
moveout (DMO) corrections
and removal of some high
frequencies thought to be
associated with noise in
1998. From Hart (1999).

FIGURE 3.48:
Examples of how
reprocessing was used to
improve seismic image
quality
Back to Chapter

Figure 3.49: Timeslice through some marine seismic data showing


strong acquisition footprint (A) and corresponding
timeslice through a reprocessed version of the data (B)

Note the improvement in seismic


imaging because of reprocessing.
Images courtesy TGS-NOPEC.

FIGURE 3.49:
Timeslice through some
marine seismic data showing
strong acquisition footprint
and corresponding timeslice
through a reprocessed version
of the data

About this Disc

Disc Contents (PDF format)

Show/Hide Bookmarks

Previous Page | Next Page

Search

Print

Close

Navigation
Show Hide

An Introduction to Seismic Interpretation Chapter Three Acquisition and Processing


survey. Masaferro et al. (2004) illustrate the application
of structure-oriented filtering to improve image quality
while preserving faults. The interpreter needs to realize
that generally these operations are a double-edge sword
and need to be used with caution and appropriate judgment.
While they may enhance data interpretability, i.e. make
the data appear more continuous and smoother, these
methods can also remove important information (e.g.,
small-scale structures, lateral amplitude variations) or,
worse, introduce artifacts that can be mistaken for geology.

Back to Chapter

Figure 3.50: Factors affecting reflection coherency across a fault

Factors affecting seismic


coherency across a
vertical fault include:
A) changes in reflection
amplitude,
B) changes in reflection
phase,
C) changes in reflection
dip,
D) changes in noise
level,
E) reflection offsets.
Based on an idea
presented by Marfurt et
al. (1998).

FIGURE 3.50:
Factors affecting reflection
coherency across a fault

3.8 Coherency (Semblance) Processing


Bahorich and Farmer (1995) introduced coherency
processing on 3-D seismic cubes as a way to enhance faults
and stratigraphic features. Since then, other processing
and software companies have developed similar attributes
that go by different names (e.g., semblance5) but,
simplistically, they are measures of the lateral variability
in trace shape i.e. how similar or dissimilar the traces
are. Hill et al. (2006) described and compared the
principal algorithms. In brief, portions of the data where
reflections are continuous have high coherency, whereas
abrupt changes in reflection shape, such reflection offsets
associated with faults, have low coherency
(FIGURE 3.50). Abrupt lateral facies changes, such as
at reef or channel margins, can also generate coherency
features.
FIGURE 3.51 shows two versions of a vertical transect
through a 3-D seismic cube. The original amplitude
version (part A) shows offsets of reflections that correspond
to faults (see CHAPTER 6). The coherency version shows
that the areas corresponding to reflection offsets correspond
to low coherency (black). FIGURE 3.52 shows a
coherency cube. The black lineations show the locations
of faults in both vertical and horizontal planes through the
cube.
The images shown in FIGURE 3.53 should help to
demonstrate the usefulness of these attributes. Part A
shows a timeslice through a 3-D seismic cube from the
Western Desert of Egypt. After some inspection, it might
be possible to recognize curvilinear trends (reflection

Back to Chapter

Figure 3.51: Comparison of A) seismic transect through a 3-D


seismic amplitude volume, and B) corresponding
transect through a coherency (semblance) volume

Original seismic data shown as variable


area wiggle display to illustrate trace shape.
Note how lateral changes in trace shape
correspond to low coherency areas (black)
that show the locations of two faults.

FIGURE 3.51:
Comparison of seismic
transect through a 3-D
seismic amplitude volume,
and corresponding transect
through a coherency
(semblance) volume
Back to Chapter

Figure 3.53: Comparison of A) amplitude timeslice and B) coherency (semblance)


timeslice for a structurally complex area

The locations of the faults are much more


clearly seen in the coherency timeslice.
From Hart (2007).

FIGURE 3.52:
Sample coherency cube
showing the locations of
faults (black)

__________________________________________________________________
5

For simplicity, I will use the term coherency to refer to these processing products, even though there are differences
in the way the attributes are derived mathematically.

Page 31

About this Disc

Disc Contents (PDF format)

Show/Hide Bookmarks

Previous Page | Next Page

Search

Print

Close

Navigation
Show Hide

An Introduction to Seismic Interpretation Chapter Three Acquisition and Processing


truncations, changes in amplitude, etc.) running from upper
left (NW) to lower right (SE) that indicate the locations
of faults. Part B of the figure shows exactly the same
timeslice, but this time through a version of the data that
has been processed to show coherency. Discontinuities,
i.e., areas of low coherency, are shown in black, and it is
clear that this attribute is clearly delineating the location
of the NW-SE striking faults. The reader should compare
the number of faults visible in the coherency display with
the number that might be interpreted in the amplitude
timeslice of the amplitude version at top. It is clear that the
coherency timeslice will be more useful for mapping faults
than the amplitude version of the same data.

Back to Chapter

Figure 3.53: Comparison of A) amplitude timeslice and B) coherency (semblance)


timeslice for a structurally complex area

The locations of the faults are much more


clearly seen in the coherency timeslice.
From Hart (2007).

FIGURE 3.53:
Comparison of A) amplitude
timeslice and B) coherency
(semblance) timeslice for a
structurally complex area

To generate a coherency volume, the interpreter must


specify a minimum of three things:
Back to Chapter

Choice of algorithm (coherence, semblance,


etc.). In practice this choice is commonly
dictated by the software available to the
interpreter or by the capabilities of the
company doing the processing.

Choice of time window. Comparing


amplitudes from trace to trace at a single
TWT is not an adequate way of capturing
changes in trace shape. As such, a sliding
window of fixed length (in ms) is used for
this purpose. A larger window will be less
sensitive to noise, but will tend to smooth
out subtle structures more than a shorter
time window.

Search pattern. The trace must be compared


to its neighbors, but the choice of how many
neighbors, and their locations, is up to the
user to decide. For example, a trace might
be compared to two, four or all eight of its
immediate neighbors.

Additionally, coherency calculations are generally able to


account for dipping reflections. Removal of acquisition
footprint prior to generating the coherency volume by poststack processing is recommended (Marfurt et al., 1998;
Chopra and Marfurt, 2008).
Coherency is a type of seismic attribute (CHAPTER 4)
that is commonly used in structural (CHAPTER 6) and

Page 32

Figure 3.54: Comparison of timeslices through coherency volumes before (left)


and after (right) post-stack processing using principal component
filtering, a technique designed to attenuate random noise

Note the improved crispness of the


faults in the image on the right.

Reproduced with permission from


Chopra and Marfurt (2008).

FIGURE 3.54:
Comparison of timeslices
through coherency volumes
before and after post-stack
processing using principal
component filtering

About this Disc

Disc Contents (PDF format)

Show/Hide Bookmarks

Previous Page | Next Page

Search

Print

Close

Navigation

An Introduction to Seismic Interpretation Chapter Three Acquisition and Processing

Show Hide

stratigraphic (CHAPTER 7) interpretations. Its derivation is presented here because, at least in


the petroleum industry, 3-D seismic cubes are routinely processed to generate coherency volumes.
Those volumes then become available to the interpreter for use at all stages of the interpretation.

3.9 References
Abriel W.L., J.P. Stefani, R.D. Shank, and D.C. Bartel, 2004, 3-D depth image interpretation, in A.R.
Brown, Interpretation of 3-D seismic data (6th ed.). AAPG Memoir 42, p. 449-475.
Accaino, F., G. Bhm, and U. Tinivella, 2005, Tomographic inversion of common image gathers:
First Break, v. 23, p. 39-44.
Albertin, U., M. Woodward, J. Kapoor, W. Chang, S. Charles, D. Nichols, P. Kitchenside, and W.
Mao, 2001, Depth imaging examples and methodology in the Gulf of Mexico: The Leading
Edge, v. 20, p. 498513.
Bahorich, M., and S. Farmer, 1995, 3-D seismic discontinuity for faults and stratigraphic features:
The coherence cube: The Leading Edge, v. 14, p. 1053-1058.
Bednar, J.B., 2005, A brief history of seismic migration: Geophysics, v. 70, p. 3MJ-20MJ.
Burger, H.R., Sheehan, A.F., and Jones, C.H., 2006, Introduction to applied geophysics, exploring
the shallow subsurface. W.W. Norton and Company, 554 p.
Chopra, S. and K.J. Marfurt, 2008, Gleaning meaningful information from seismic attributes: First
Break, v. 26, p. 43-53.
Cordsen, A., M. Galbraith, and J. Peirce, 2000, Planning land 3-D seismic surveys: SEG Geophysical
Development Series, 9, 204 p.
Crawley Stewart, C.L., 1995, 3-D solution to a 2-D pitfall: seismic detection of carbonate buildups,
Kiva Field, Paradox Basin, San Juan County, Utah, in High-definition seismic 2-D, 2-D swath,
and 3-D case histories (R.R. Ray, ed.), Rocky Mountain Association of Geologists Guidebook,
p.177-183.
Cummings, D.I. and H.A.J. Russell, 2007, The Vars-Winchester esker aquifer, South Nation River
watershed: Geological Survey of Canada, Open File 5624, 68 p.
Dessler, J. F., 1992, Marine seismic data acquisition in Development Geology Reference Manual
(Ms. Morton-Thompson and A. M. Woods, eds.): AAPG Methods in Exploration 10, p. 361-363.
Dix, C.H., 1955, Seismic velocities from surface measurements: Geophysics, v. 20, p. 68-86.
Duncan, P.M., 1992, Basic seismic processing in Development Geology Reference Manual (M.
Morton-Thompson and A.M. Woods, eds.): AAPG Methods in Exploration 10, p. 364-271.
Evans, B.J., 1997, A handbook for seismic data acquisition in exploration. SEG Geophysical
Monographs Series, 7, 305 p.
French, W.S., 1974, Two-dimensional and three-dimensional migration of model-experiment
reflection profiles: Geophysics, v. 39, p. 265-277.

Page 33

About this Disc

Disc Contents (PDF format)

Show/Hide Bookmarks

Previous Page | Next Page

Search

Print

Close

Navigation

An Introduction to Seismic Interpretation Chapter Three Acquisition and Processing

Show Hide

Gadallah, M.R., and R.L. Fisher, 2005, Applied seismology. PennWell, 473 p.
Galbraith, M., 2001, 3D seismic surveys past, present and future: CSEG Reservoir, p. 9-12,
available at: http://www.cseg.ca/publications/recorder/2001/06jun/jun01-3d-seismic.pdf.
Gendzwill, D.J., M.H. Serzu, and G.S. Lodha, 1994, High-resolution seismic reflection surveys
to detect fracture zones at the AECL Underground Research Laboratory: Canadian Journal of
Exploration Geophysis, v. 30, p. 28-38.
Hardage, B. A., 1997, Principles of onshore 3-D seismic design: The University of Texas at Austin,
Bureau of Economic Geology Geological Circular 97-5, 23 p.
Hart, B.S., 1999, Definition of subsurface stratigraphy, structure and rock properties from 3-D
seismic data: Earth-Science Reviews, v. 47, p. 189-218.
Henry, G., 1997, La sismique rflexion, principes et dveloppements. ditions Technip, 172 p.
Henry, S., 2001, Understanding the seismic wavelet, AAPG Search and Discovery Article #40028,
http://www.searchanddiscovery.net/documents/geophysical/henry/index.htm.
Hill, S.J., K. Marfurt, and S. Chopra, 2006, Searching for similarity in a slab of seismic data: The
Leading Edge, v. 25, p. 168-177.
Krey, T., 1987, Attenuation of random noise by 2D and 3D CDP stacking and Kirchhoff migration:
Geophysical Prospecting, v. 35, p. 135-147.
Lansley, R.M., 2004, CMP fold: a meaningless number?: The Leading Edge, v. 23, p. 1038-1041.
Lawton, D.C., J.H. Isaac, R.W. Vestrum, and J.M. Leslie, 2001, Slip-slidin awaysome practical
implications of seismic velocity anisotropy on depth imaging: The Leading Edge, v. 20, p. 70-73.
Leetaru, H.E., 1996, Seismic stratigraphy, a technique for improved oil recovery planning at King
Field, Jefferson County, Illinois. Illinois State Geological Survey, Illinois Petroleum 151, 37 p.
Lehmann, B., 2007, Seismic traveltime tomography for engineering and exploration applications.
EAGE Publications BV, 273 p.
Liner, C.L., 2004, Elements of 3-D Seismology. PennWell, 608 p.
Marfurt, K.J., R.M. Scheet, J.A. Sharp, and M.G. Harper, 1998, Suppression of the acquisition
footprint for seismic sequence attribute mapping: Geophysics, v. 63, p. 1024-1035.
Mosher, D.C. and P.G. Simpkin, 1999, Status and trends of marine high-resolution seismic profiling:
data acquisition, Geoscience Canada, v. 26, p. 174-188.
Nedimovi, M.R., S. Mazzotti, and R.D. Hyndman, 2003, Three-dimensional structure from
feathered two-dimensional marine seismic reflection data: The eastern Nankai Trough: Journal of
Geophysical Research, v. 108, p. doi:10.1029/2002JB001959
Ogiesoba O.C., and Hart B.S., 2009, Fault imaging in hydrothermal dolomite reservoirs: A case
study. Geophysics, v. 74 , p. B71-B82.

Page 34

About this Disc

Disc Contents (PDF format)

Show/Hide Bookmarks

Previous Page | Next Page

Search

Print

Close

Navigation

An Introduction to Seismic Interpretation Chapter Three Acquisition and Processing

Show Hide

Ray, R.R., 1995, Utilizing seismic for higher definition of geologic prospects: 2-D, 2-D swath, and
3-D seismic in High-definition seismic 2-D, 2-D swath and 3-D case histories (R.R. Ray, ed.):
Rocky Mountain Association of Geologists Guidebook, p. 1-6.
Rawlinson, N., and M. Sambridge, 2003, Seismic traveltime tomography of the crust and
lithosphere: Advances in Geophysics, v. 46, p. 81-198.
Roberts, G., K. Rutherford, and C. OBrien, 2008, Observations on the petroleum potential of deep
offshore west coast India from newly reprocessed 2D seismic data: First Break, v. 26, p. 77-86.
Schuck, A., and G. Lange, 2008, Chapter 4.6 Seismic Methods, in K. Kndel, G. Lange and H.-J.
Voigt (eds.), Environmental Geology, Springer, p. 337-402.
Seeber, M.D., and D. Steeples, 1986, Seismic data obtained using .50-caliber machine gun highresolution seismic source. AAPG Bulletin, v. 70, p. 970-976.
Shabrawi, A., A. Smart, B. Anderson, G. Rached, and A. El-Emam, 2005, How single-sensor seismic
improved image of Kuwaits Minagish Field: First Break, v. 23, p. 63-69.
Sheriff, R.E., 2002, Encyclopedic dictionary of applied geophysics. Society of Exploration
Geophysics, Geophysical References Series, 13, 429 p.
Sheriff, R.E., and L.P. Geldart, 1995, Exploration Seismology (2nd Ed.). Cambridge University
Press, 592 p.
Steeples, D.W., and R.D. Miller, 1998, Avoiding pitfalls in shallow seismic reflection surveys:
Geophysics, v. 63, p. 1213-1224.
van der Veen, M., and A.G. Green, 1998, Land streamer for shallow seismic data acquisition:
evaluation of gimbal-mounted geophones: Geophysics, v. 63, p. p. 1408-1413.
Veeken, P.C.H., 2007, Seismic stratigraphy, basin analysis and reservoir characterization: Elsevier,
509 p.
Vermeer, G., 2002, 3-D seismic survey design, SEG Geophysical References Series 12, 205 p.
Wu, J., 1992, Potential pitfalls of crooked-line seismic reflection surveys: Geophysics, v. 61, p. 227281.
Yilmaz, O., 2001, Seismic Data Analysis. SEG Investigations in Geophysics, 10, 2027 p.

Page 35

Das könnte Ihnen auch gefallen