You are on page 1of 37

1

Well Control 5
Geophysics

CONTENTS

1. WHY IS GEOPHYSICS IMPORTANT ?

2. TYPES OF GEOPHYSICAL INVESTIGATIONS

3. BASICS OF WAVE PROPAGATION

4. REFLECTION SEISMOLOGY PRACTICES

5. INTERPRETATION OF REFLECTION
SECTIONS

6. SUMMARY

7. EXERCISES
1

LEARNING OBJECTIVES

The aims of this Chapter are to give the Petroleum Engineer:

• An understanding of the basic geophysical concepts as used in the petroleum


industry

• An overview of the applications of seismic data in reservoir description

After studying this Chapter the student should be able to:

1. Explain the rationale for geophysical investigations

2. List the characteristics of the main geophysical methods

3. Describe the propagation of P- and S-waves

4. Explain how waves are altered at a material interface

5. Describe the nature of seismic reflections

6. Define: acquisition, source, receiver, processing, interpretation and migration

7. Discuss the process of interpreting seismic data

8. List the roles of seismic in reservoir imaging

9. List what properties may be determined with seismic data

10. Describe the controls on, and limits of, seismic resolution

2
Geophysics 5
1. WHY IS GEOPHYSICS IMPORTANT?

Petroleum Engineers will find themselves working alongside Geophysicists, and they
will make use of geophysical data. A basic understanding of geophysical data - its
strengths and weaknesses - and how the data are used for interpreting the subsurface,
are therefore useful preparations for those experiences.

What can an Engineer gain from geophysical data? Geophysical data can provide the
Engineer with vital information concerning the planning of drilling activities, and the
planning for field development. This information includes:

• The geometry of the rock bodies that comprise the petroleum system, including the
over-burden, the under-burden, the side-burden, and the reservoir itself (the
boundary of the hydrocarbon-bearing region can be called the reservoir envelope),
along with the characteristics of the petroleum kitchen and its associated migration
routes

• The distribution of reservoir properties (its internal architecture), and the properties
of the surrounding rocks

• The size and distribution of the aquifer

• The presence of internal compartments or disruptions of the reservoir (such as


faulting)

• Possibly, direct indications of hydrocarbons

• Sometimes, the orientation of fractures

• In some cases, indications of high pore pressures

However, for reasons that will be discussed in this Chapter, these benefits are not
always realised in practice. The un-fulfilled potential of Geophysics is largely related
to problems with data quality, but sometimes the reduced benefits are related to
inappropriate techniques and limited interpretation skills. Understanding the basic
controls on data quality (resolution) will help the Engineer appreciate the limits of the
data - although these limits are being consistently pushed back as technology develops
in all areas of the geophysical method (acquisition, processing, interpretation and
visualisation).

The main motivation for undertaking geophysical surveys is to develop an image of


the subsurface geology. The secondary motivation is to determine the properties of
the rocks and to detect their contained fluids. These goals are met in practice by close
integration throughout the entire Geoscience team, and between it and the Engineer-
ing team.

The rationale underpinning all geophysical methods is that rocks are not all alike, and
that these differences are manifest by differences in the physical properties of the
rocks. In Chapter 1, you discovered that there are many kinds of rocks, with a wide

Department of Petroleum Engineering, Heriot-Watt University 3


1

range of compositions. In Chapter 3, we saw that sedimentary rocks are deposited by


processes that result in spatial heterogeneity of composition, grain size, sorting,
packing, etc, and that these processes also result in the stacking of layers with differing
characteristics. In Chapter 4 it was shown that natural processes that are active in the
Earth disrupt the original rock-body distributions, producing translations, rotations,
and distortions of the original simple layered arrangements. Other natural events, such
as fluid flows that can produce variations in the cementation of the grains, also add to
the heterogeneous distribution of rock properties. Even the accumulation of hydro-
carbons has an effect on the bulk physical properties of the reservoir rock.

Each of these changes in physical character can produce changes in the physical
properties. For the purpose of developing an image of the subsurface, some of the
important physical properties of rocks are:

• The sonic velocity (speed of passage of acoustic waves)

• The density

• The resistivity (or its converse, the conductivity)

• The magnetic compliance, and/or the natural magnetism

The premise of Geophysics is that these physical properties have effects “at a
distance”. In other words, their variations can be detected in locations that are
removed from the site of the actual property variation - such as at the ground surface.
The challenge, and success, of Geophysics is to develop and improve the methods by
which these detected variations can be related to the spatial locations from which the
effects arise. Useful interpretations of the data depend on both careful calculations and
visualisation methods that create “geological” displays that correspond to our mental
image of the subsurface geology.

Scope of this Chapter


In this Chapter, we will briefly consider the range of geophysical methods, but the
majority of the Chapter focuses on seismic techniques. We describe the way that
acoustic waves travel through solids, and especially the way that these waves are
altered as they pass from one material to another. We then focus exclusively on active
seismic investigations, in which sonic energy is deliberately propagated into the
subsurface, and where the reflected energy is collected and processed to develop an
image of the subsurface configuration of the rocks. A final short section discusses how
seismic methods are being used in a new area of activity known as Reservoir
Geophysics.

2. TYPES OF GEOPHYSICAL INVESTIGATIONS

An appreciation of the physical properties of rocks has been used to create the four
primary geophysical investigation techniques:

4
Geophysics 5
• Gravity

• Electrical

• Magnetic

• Sonic

Gravity methods are based on the fact that changes in density of parts of the Earth (e.g.
a body of rock that has a different density than the surrounding rocks) can be detected
because of their effect on the gravity field. Of course, the Earth itself is responsible
for the majority of the observed gravitational field (it produces an acceleration at the
surface of approximately 9.8 m/sec2), while perturbations associated with changes in
1-10 mgal (1 mgal = 103 gal; 1 gal = 1cm/sec2, so gravitational accelaration ≈ 980 gal).
So, measurements require accuracy of better than one part in a million. Gravitational
effects associated with density (mass) changes decline with distance. Small density
changes that occur near the observation point (usually, the Earth’s surface) can have
as much impact on the observed gravity field as a much larger density (mass) change
that occurs at a larger distance. In practice, gravity surveys are used primarily for
detecting large rock bodies that have a significant density anomaly (e.g. a salt diapir).

Electrical methods are based on the fact that rocks have varying resistivities. In most
cases, the resistivity is actually a “bulk” value representing the resistive rock
framework and the (often lower) resistivity of the contained porefluids. In practice,
an electrical potential (voltage) is applied across widely-separated electrodes located
at the ground surface, and another electrode pair is used to determine the electrical
potential at intervening sites. Source rocks, because of their high carbon content, can
have lower resistivities. Sometimes a survey in undertaken to assess the regional
extent of a source rock layer. However, electrical techniques are not commonly used
for reservoir studies.

Magnetic methods are based on the fact that rocks can alter the observed magnetic
field of the Earth. We all know about the main field of the Earth (e.g. compass needles
pointing to the magnetic poles). Rocks that have their own magnetic fields (these can
be produced by some forms of iron- and titanium-containing minerals) can add or
subtract from the main field. If these changes can be detected, the location of the
magnetised rocks can be deduced. Strongly-magnetised rocks are not very common,
but weakly-magnetised rocks occur in many places (red-beds, basalts, etc). All rocks
affect the local magnetic field through the way that the magnetic “forces” are
transmitted through them. This effect is called the magnetic compliance, and it too
can be used to infer the spatial distribution of rock units. The spatial resolution of
magnetic investigations is usually very coarse, and the uncertainties render this
technique of only limited value. Magnetic methods are not widely applied in
Petroleum Engineering.

In contrast to the above methods, seismic (for seismos - the Greek word for shaking)
investigations play a major role in Petroleum Geophysics. The money spent on
seismic accounts for perhaps 70-80% of the Geoscience budget of most companies.
Because of this strong economic emphasis, it is important for the Engineer to

Department of Petroleum Engineering, Heriot-Watt University 5


1

appreciate both the value and limitations of this method relative to the effort to
understand the subsurface operation of the Petroleum System.

Seismic waves in rocks are periodic, oscillatory motions; they are a form of acoustic
wave. Seismic waves can be created naturally (e.g. by an earthquake), or they may be
intentionally induced as part of an investigation (this is our primary emphasis in this
Chapter). Because the sonic properties of rocks vary from place to place (and perhaps
over time as a consequence of producing a reservoir), the passage of seismic waves
is locally affected by the rock pattern and the distribution of fluids. These effects are
used by Geophysicists to develop an image of the subsurface distribution of rocks and
their contained fluids.

There are two main categories of seismic techniques. One is termed (seismic)
refraction. In this method, the goal is to determine the position, and velocity and
density, of rock bodies. It is effective at “seeing” situations where the lateral
dimension is very much greater than the depth. Its depth resolution/accuracy is rather
poor (at least for problems of interest to Petroleum Engineering). This approach has
only limited applications to reservoir problems (but note that refraction can be used
to image the extent of major basalt flows, and it is important in making static
corrections; see below), and we will not discuss it further here.

The other seismic method is called reflection seismology, or reflection seismic, or


often just “seismic”. In this method, sonic (vibrational) energy is deliberately
propagated into the ground, and the approach is to detect the sonic energy that is
reflected back to the surface, and to use these data to create an interpretation of the
subsurface. Using the ideas developed in the remainder of this Chapter, the returned
seismic data is processed and presented in a form suitable for interpretation. The usual
approach is to display the processed data in a way that looks similar to a geological
cross section (Fig. 1). Figure 1
A seismic section from the
southern North Sea,
demonstrating how seismic
data can image geological
structures. In this example,
a salt diapir has deformed
the overlying sedimentary
rocks, and we can also see
packages of sedimentary
rocks deposited into local
paleo-depo-centres that
The remainder of this Chapter will concentrate on the seismic reflection method, since evolved as the salt moved.
it is the primary geophysical technique that will be encountered by a practising
Petroleum Engineer.

6
Geophysics 5
3. BASICS OF WAVE PROPAGATION

Above, we said that a seismic wave is an acoustic wave passing through a rock
medium. Such waves are called body waves. They are considered to be elastic waves
(see Rock Mechanics Appendix) because there is no permanent distortion caused by
their passage. There are two types of body waves:

• P-waves - ComPressional, longitudinal, Pressure, or Primary waves (Push-Pull)


• S-waves - Shear, transverse, tangential, or Secondary waves (Side-to-Side)

These body waves consist of particle motions that involve oscillation (periodic
motion, hence the term wave) about a fixed point. In other words, the body itself does
not change its position in space, but its constituent particles vibrate about their
positions. In the normal situation that applies to the Earth, rock bodies are not
vibrating, because any earlier acoustic waves that may have existed will have
dissipated (due to attenuation; see below). The initiation of vibration at a point,
especially the way that such initiation of movement sweeps through the rock mass, is
called the propagation of a wave.

3.1 P-waves
A P-wave is an elastic body wave in which the particle motion involves oscillations
in the direction of propagation of the wave (Fig. 2). Perhaps a simple thought-
experiment will clarify this process. Imagine several students standing in a line, all
equally spaced apart. Between each student, there is a spring that allows both
shortening and elongation of the distance between the students. To start the acoustic
wave, one of the students (assume it is the one at the end of the line) moves towards
the student next to her (and away from the wall). Because of the springs, the student
ahead of the initial student must move forward, and the one ahead of him must then
move forward in turn. To someone watching this process, there is an advancing
disturbance that alters the positions of the students. This disturbance represents the
propagation of the wave.

Time = 0.0 Spring Compressed

Spring Lengthened

Time = 0.0 + ∆t
Springs Compressed

Figure 2
Spring Lengthened
A one-dimensional
Time = 0.0 + 2∆t
demonstration of the Springs Compressed

propagation of a
compressional wave, using
students as “particles”
Springs Lengthened

Department of Petroleum Engineering, Heriot-Watt University 7


1

Now let us consider what happens after the initial movement has passed a particular
student. Because of the springs “behind” that student, the student is pulled back
towards his/her original position. In fact, the stretch of the spring will cause an “over-
recovery” that induces him/her to move further back than where he/she started. Thus,
once a vibration is initiated, there is a tendency to continue the oscillation to and fro.

If we adopt a sign convention such that student spacings that are close together are
considered positive, we can draw a curve representing the density of the students at
any particular time (Fig. 3). This curve shows places where the students are bunched
close together, and places where they are spread apart. The curve has a sinusoidal
form. The places where there is a bunching-together are called compression, and
places where the students are separated are called rarefaction (dilation; sometimes
written as dilatation). Importantly, the position of the compressional and dilational
spacings is not fixed. Instead, each student is alternately (indeed, periodically) part
of the compressional or dilational state. (You may have seen a “wave” at a sports
event: spectators stand and then sit down in such a way that an observer sees this
process sweep around the stadium. If we let this standing/sitting process represent the
compression/rarefaction of particles, then it is clear that a waveform moves past any
particular point.)
Compression

Figure 3
Position Plot of compression/
rarefaction of student
Dilation

spacings (at some


Wavelength particular time) as a
function of distance

Now let us consider how the speed of propagation of the wave is related to the distance
between successive compressional maxima. This distance is known as the wave-
length. Depending on the “springiness” of the springs, there is a characteristic
velocity of propagation. The velocity and wavelength are related by the frequency
of the waves (the number of waveforms passing an observation point in a given time
interval). This relationship is:

V
λ=
f
where Equation 5.1
λ = wavelength Relationship between
V = velocity frequency, velocity, and
f = frequency wavelength

Of course, our interest lies with rocks, instead of with thought-experiments involving
lines of students. The principles outlined above apply equally to the propagation of
acoustic waves in rocks, and the students and springs can be thought of as symbols
for the atoms, and their associated bonds, that make up the mineral constituents of the
rocks.

8
5
Geophysics

The preceding narrative treated a single line of particles (e.g. the students). This one-
dimensional conceptual model must be extended to higher dimensions in order to
address other important aspects of wave propagation. Now, imagine - not just a line
of particles, but a two-dimensional array (Fig. 4). If an acoustic wave propagates
across this array in a fashion so that the initiation of the oscillations (the wavefront)
excites vibrations in particles along a (moving) straight line, this mode of propagation
is called a plane wave. Although this situation can be readily understood as a concept,
plane waves are somewhat rare in seismic investigations. Instead, most induced
seismic waves propagate away from their source point with a (nearly) spherical
wavefront (Fig. 5). This is known as spherical radiation. The reason for the non-
sphericity actually observed is that rock velocities are not constant everywhere.
wavefront

propagation
direction

compressed zone dilated zone

Figure 4
Example of a plane wave
affecting a regular array of
patrticles are
particles spaced together
particles are
spaced apart

seismic source
in shothole

Figure 5
D il a ti o n
Spherical radiation of a sio n
C o m pres

compressional wave from Dilation

io n
its point source C o m press

Department of Petroleum Engineering, Heriot-Watt University 9


1

3.2 S-waves
An S-wave is an elastic body wave in which the particle motion involves oscillation
in planes that are perpendicular to the direction of propagation of the wave. If we
return to the image of a line of students (Fig. 6), this mode of propagation involves the
students having springs attached to their heads and feet. The connections between the
students are not springs, but rigid bars, so the students do not move forward and
backward, but up and down, as the wave propagates along the line. This type of motion
is known as shear. In this example, the initiation of the wave is associated with a shear
disturbance. Another way to initiate shear might be where a P-wave impinges on the
line at an angle from the side. We will see below that this process occurs at interfaces
between materials that have different properties.

Spring Lengthened Spring Compressed

Figure 6
Vertical polarisation of a
shear wave affecting the
Spring Compressed Spring Lengthened
line of students

Although we imagined a vertical shear in our example, the students might alterna-
tively be caused to move from side to side (if the springs were to either side of them)
as the wave propagates (Fig. 7). The plane containing the shear movements is called
the polarisation plane. In fact, its orientation is not restricted to the vertical or
horizontal, but it can be at any inclination. The direction of polarisation is important
because it relates to the shear movement of the initial disturbance, or to changes in
polarisation induced by rock-property changes. Fluids (like air and water) do not
transmit shear waves (they lack the “rigid” connecting bars of our analogy).
Spring Lengthened

Spring Compressed

Spring Compressed

Figure 7
Horizontal polarisation of a
Spring Lengthened
shear wave

10
Geophysics 5
3.3 Some additional characteristics of waves
As spherical waves propagate, they must use their energy to excite the movements of
ever-increasing numbers of particles as the wavefront expands. The surface area of
a sphere is determined by its radius (depending on the radius (r) squared), so the energy
of a spherical wave degrades as it propagates proportionally to 1/r2. This cause of the
loss of energy is called divergence. Other losses of wave power occur because energy
is needed to make particles change their motions (i.e. to overcome their inertia). This
converted energy is physically expressed as a small increment of heat energy. (In
practice, the heat is so small that it is simply “lost” in the system.) Because the
conversion of motion to heat is similar to the energy conversion that occurs as
materials slide along a frictional surface, this internal loss is sometimes called internal
friction, but the technical term is absorption. Higher-frequency waves lose their
power more readily over a given distance than do lower-frequency waves; this is
largely because there are more oscillations per second for the higher-frequency wave,
and each oscillation has a “loss” attached to it. Additional energy is lost from a given
wave as it is converted into other wave types (at interfaces with different materials;
see below). These reflections and scattering reduce the power of an advancing wave.
Together, these losses of energy mean that waves eventually attenuate. The
attenuation of waves is why the Earth is normally “at rest” and not still vibrating from
the passage of previous waves. (However, there may be some vibrations occurring at
almost any time, caused by many natural and human-related causes, and these
represent “noise” that interferes with seismic investigations.)

P-waves travel faster than S-waves. As a rule of thumb, P wave velocities (Vp) are
usually about 1.7 times as fast as those of S-waves (Vs). This difference is related to
the mechanics of the process (see Rock Mechanics Appendix). The P-wave velocities
of some common rock types are given in Table 1. Note that there is a range of
velocities for any rock type. This range is related to variations in porosity, cementa-
tion, composition, and other factors.

Rock Type Vp (km/sec)


Anhydrite 4.1 - 5.0
Basalt 5.0 - 6.4
Chalk 2.1 - 4.2
Dolomite 3.5 - 6.9
Limestone 1,7 - 7.0
Table 1 Salt 4.4 - 6.5
Compressional wave Sandstone 1.4 - 4.3
velocities for common rock Mixed sand/shale 2.1 - 4.5
types Shale - Slate 2.3 - 4.7

Because they have a polarisation, S-waves can be significantly affected by anisotropies


of the medium (i.e. a medium whose properties differ in different orientations). If one
orientation is much faster than the other ones, then it may be possible to detect this by
observing different propagation velocities in the different polarisation planes. This
aspect underpins efforts to identify open fractures through shear wave variations.

Acoustic waves are affected as they cross the interface between different materials
(e.g. passing from one rock layer to another that has different properties). Referring

Department of Petroleum Engineering, Heriot-Watt University 11


1

to our student analogy, this represents springs of different stiffnesses attached to the
students at one end of the line versus at the other end of the line. These variations in
stiffness relate to variations in velocity. If the frequency of the wave is conserved, the
wavelength must change as it propagates into the new material. Equation 5.2
V1 V2 Change of wavelength
=f = ,
λ1 λ2 caused by change of
if V 2 > V 1, λ 2 > λ 1 velocity at constant
frequency (material 1 to
If we move away from our one-dimensional analogies, we need to be aware that it is material 2)
wavefronts that cross material interfaces, and that the change in velocity produces a
change in the orientation of the vector representing the wavefront movement (Fig. 8).
The energy of the incoming wave is partitioned, with some of it possibly continuing
across the interface, some being reflected, and some may move along the interface.
There is also a potential that waves can be converted from one type to another (eg P-
waves converting to S-waves, and vice versa). This process is called mode conversion.

Reflected S-wave
Incident
P-wave Reflected P-wave

Interface Figure 8
Transmitted Propagation of wavefronts
Transmitted P-wave at an interface
S-wave

Let us now consider in greater detail what happens when a propagating acoustic wave
encounters an interface (e.g. a change to a separate rock body with different
properties). The simplest configuration is one in which the interface is horizontal, and
the wave (let us assume a compressional, or P-wave) is propagating downwards (Fig.
9). (For this explanation, it is useful to represent the advancing front of the wave by
an arrow. The paths indicated by the arrows are called the raypaths of the wave.) In
general, a wave can approach a point from any direction. When the point is on an
interface, we can identify the angle between the wave’s direction of advance, and the
orientation of the interface. This angle is called the incidence angle. When a wave
is travelling at right angles to an interface, this is called normal incidence, so the left
example is a vertical, normal-incidence case. The question is: what will happen as the
wave (vibrations of particles) encounters the interface?

successive
incidence angle compressional
wavefronts
10
=
e
ti m 11
e
= Figure 9
ti m 12
= Incidence angles of an
θi θi e
ti m 13
θi = approaching wavefront
e
ti m
interface
normal
incidence
(perpendicular)
12
Geophysics 5
The first concept that needs to be covered (to enable this question to be answered) is
that of acoustic impedance. This parameter is determined as the product of the density
and the compressional velocity:

V
λ=
f
where
λ = wavelength
Equation 5.3
Definition of Acoustic V = velocity
Impedance f = frequency

The acoustic impedance can be associated with the ease of transmission of vibrations.
In principle, the density and velocity of a rock (which determine the acoustic
impedance) are independent properties, but empirical data indicates that they are
correlated (i.e. if one increases, so does the other). However, density variations are
less than velocity variations (expressed as a percentage). Rock densities range from
about 2.2 kg/m3 (g/cm3) to about 2.8, with most values being nearer 2.5 or 2.6. The
table of common rock sonic velocities presented above shows a variation of at least
a factor of two. Therefore, the dominant control on acoustic impedance is the velocity.

The second necessary concept is that of the reflectivity of an interface. In our example,
the rock above the interface is termed medium number 1, and the rock below it is called
medium number 2. The reflection coefficient for a wave passing from medium 1 to
Equation 5.4 medium 2 is:
Definition of Reflection
Coefficient between V 2 ρ 2 -V 1 ρ 1
R12 =
materials 1 and 2 V 2 ρ 2 +V 1 ρ 1

The energy reflected (“bounced back”) from the interface is determined by multiply-
ing the amplitude of the incident wave by the reflection coefficient. The remaining
energy is, in our example, transmitted downwards into the second medium. The
reflection coefficient is greater than 0.3 for the “best” interfaces, but is generally less
than 0.1 for typical geological interfaces.

There are conversions of wave energy that occur if the incidence angle is not
perpendicular to the interface. (These situations are the “norm”, and vertical, or
normal, incidence is the rarity.) For each interface, there is a critical angle, determined
by the velocities of the two media, that controls the behaviour (Fig. 10). Snell’s Law
dictates the angles between the raypaths of the inbound and reflected/transmitted
wave. At incidence angles less than the critical angle, an incoming P-wave is partly
reflected as another P-wave, partly reflected as an S-wave, and some (most) of the
energy is transmitted to the next medium as both a P-wave and an S-wave (Fig. 10a).
When the incidence angle equals the critical angle, most of the incoming energy is
refracted along the interface (Fig. 10b), but there is also a reflection and transmission
of S-waves. At angles greater than the critical angle, all of the incoming energy is
reflected (Fig. 10c). The equations that represent these energy and mode conversions
are more complicated than we need to address in this Chapter.

Department of Petroleum Engineering, Heriot-Watt University 13


1

Incident P-wave
P
Figure 10
Incident P-wave Reflected P-wave Consequences ofwaves
θi
S
θi
Reflected P-wave
approaching an interface at
Incident P-wave
θc θc differing incidence angles.
Reflected Waves
θi
a) Incidence angle less than
critical angle; b) Incidence
(A) (B) Refracted Waves (C) angle equal to critical
angle; c) Incidence angle
P greater than critical angle.
S
Transmitted Waves

θi < θc θi = θc θi > θc

When waves are reflected, refracted, or converted at an interface, there are important
changes that can be induced. For example, their sign can be changed, and for shear
waves, their polarisation direction can be altered. If the acoustic impedance of the
second medium is greater than that of the first, the reflection coefficient is positive
(refer to Equation 5.4), and any reflected wave has the same sign as the incoming
wave. In other words, an initial compressional waveform is reflected as a compres-
sional waveform. On the other hand, if the second medium has a lower acoustic
impedance, the reflection coefficient is negative, and an incoming compression is
reflected as a rarefaction (dilational waveform). If the acoustic impedances are
identical, there is no reflection.

3.4 Imaging Reflections


The principle of reflection of (some) wave energy from an interface is the basis for the
entire seismic industry. Modern advances are now also capitalising on other aspects
such as mode conversions, and additional opportunities exist as a consequence of the
way that seismic data are acquired and processed. Subsequent sections of this Chapter
will address how these approaches are implemented in the modern seismic industry.
First, however, we will see how the basic knowledge outlined above translates into
images of reflections.

Seismic sources (see below) produce an acoustic wave that contains many frequen-
cies. Therefore, a real reflection consists of many different waveforms (each a perfect
sinusoid, but not in phase with each other because their different velocities will have
shifted them in time). If these waveforms are all added together (as they would be in
a real case), a new, non-sinusoidal waveform is created. (In reverse, this process is
what happens when we de-compose an arbitrary waveform into its Fourier compo-
nents.) The non-sinusoidal reflection waveform is called a wavelet, and it represents
the net time-varying amplitude of the returned energy.

In the “perfect” vertical, normal-incidence example that we introduced above, this


principle can be developed further to produce what we call a seismic trace. A trace

14
Geophysics 5
is the amplitude versus time plot of the signal that is received at the surface (in real
practice, it is the composite signal of many receivers; see below). Seismic traces, from
adjacent “receivers”, are plotted next to one another to produce the reflection seismic
section (as illustrated in Fig. 1).

Let us consider the simplest case: a two-layer Earth, with layer 1 being horizontal and
extending from the surface down to some uniform depth (say 1000 m), and layer 2
occupying the region below that depth (Fig. 11). A source and receiver are located at
the surface. At time = 0.0 sec, a seismic pulse is generated at the source. The resulting
acoustic wave propagates (radially) downwards (and laterally, to other locations, but
we are going to ignore that for the moment). The compressional wavefront moves at
the speed determined by the compressional velocity of the material (Vp) (here, let us
say 2500 m/sec). Given this velocity, at time = 0.4 sec (400 ms), the compressional
wave reaches the interface. The physical properties indicate that a positive reflection
coefficient characterises the interface, so the downgoing incident compressional
waveform is reflected (but at lesser amplitude) as an upgoing compressional wave-
form. The reflected waveform reaches the surface 0.4 sec later. The total time of travel
of the wave, from the source, down to the reflector, and back to the receiver, is 0.8 sec
(800 ms). This is called the two-way (travel) time, representing the total time to travel
down and back up. This parameter is ALWAYS abbreviated as TWT. The single
seismic trace that would be observed is shown in Figure 12. This has a wavelet
centered on the TWT of 0.8 sec. The crucial rule for seismic reflection traces is that
they ALWAYS depict the TWT. The other rule is that they always depict compres-
sional waveforms as a deflection of the curve to the right.

Model Configuration

source + receiver location


surface

down-going and
1000 m

Figure 11 up-going waves Layer 1: Vp = 2500 m/sec, ρ = 2.5 g/cc


Simple, two-layer model
illustrating normal-
incidence wave reflection interface
from an interface
Layer 2: Vp = 3000 m/sec, ρ = 2.8 g/cc
transmitted wave

Department of Petroleum Engineering, Heriot-Watt University 15


1

0.0

TWT (sec)

0.5

single
Figure 12
wavelet
Ideal seismic trace
1.0 produced by configuration
of Fig. 11
Seismic Trace

A more-involved example is shown in Figure 13. Here we have several layers, each
having its own thickness and sonic properties. Note, however, that we are still
considering each layer to be perfectly horizontal. Based on the argument above
(concerning the relative importance of velocity and density in determining reflectiv-
ity), we can estimate the positive or negative reflection characteristics of the interfaces
by means of the compressional velocities only. For this problem, we need to calculate
the TWT for each layer (each velocity interval), and from these values, determine the
cumulative TWT for the deeper reflections. These numbers are shown in Table 2.
From the tabulated results, we can construct a seismic trace (Fig. 14), placing an
appropriately signed wavelet at each reflection time (TWT).

Multi-layer Model
600 m

Layer 1 Vp = 2500 m/sec

R12 negative
850 m

Layer 2 Vp = 2000 m/sec

R 23 positive
350 m

Layer 3 Vp = 3000 m/sec


R 34 positive
Figure 13
Layer 4 Vp = 3500 m/sec Multi-layer model

16
Geophysics 5
Table 2 Layer Interval Thickness Interval Velocity Interval TWT Cumulative TWT
Determination of TWT for 1 600 2500 0.480 0.480
2 850 2000 0.850 1.330
reflections expected in 3 350 3000 0.233 1.563
model of figure 13

Seismic Trace

0.0

0.5
TWT (sec)

1.0
Figure 14 Seismic trace
produced by model
illustrated in Figure 13 1.5

This procedure can be used to construct a synthetic seismic trace at several locations
if the subsurface configuration is not horizontally-layered. For example, depths and
velocities may be known from wells spaced some distance apart, and these locations
can be used to construct such traces. The traces can be mentally “connected” to gain
an impression of the form of the seismic reflection section that would be produced by
the subsurface configuration (see Exercises). When traces are spaced closely
together, the positive and negative waveforms appear to merge (visually). This effect
is enhanced by the common practice of colouring-in the positive wavelets. The
resulting bands of dark and light colour are called peaks and troughs, representing the
compressional and dilational portions of the seismic waveforms.

Synthetic traces can be used to address a very important issue: the ability (or not) to
image a “thin bed”. Consider an otherwise-uniform Earth that has a single, horizontal
thin layer of contrasting velocity (Fig. 15). Let us assume that the velocity of the
majority of the Earth is 1500 m/sec, and that the velocity of the thin bed (1 m thick)
is much greater at 10,000 m/sec. Surely, this extreme contrast of velocity will generate
a major reflection?

Department of Petroleum Engineering, Heriot-Watt University 17


1

Thin-Bed Model
750 m

Layer 1 Vp = 1500 m/sec

Layer 3 Vp = 1500 m/sec

R 12 positive
0.2 ms + =
1m

Layer 2 Vp = 10,000 m/sec


apart Figure 15 The “thin-bed”
R 23 negative wavelets "cancel" imaging problem

This situation can be treated as a three-layer problem. The reflection coefficient for
the top of the thin bed will be large and positive, and the reflection coefficient for the
bottom of the thin bed will also be large, but negative. Given the depth to the top of
the thin bed (750 m), and the velocity of Layer 1 (1500 m/sec), the TWT to this
reflector is 1.0 sec. Assuming a dominant frequency of 25 Hz for the seismic signal,
the wavelength in Layer 1 is 60 m. The interval TWT for Layer 2 is very small: 0.0002
sec (0.2 ms). (Note that the wavelength in Layer 2 is 250 m.) Thus, the reflection from
the base of Layer 2 occurs at the TWT of 1.0002 sec (1.0 + 0.0002). The wavelet for
the reflection at the base of Layer 2 is an exact mirror image of the wavelet for the top
of that layer. The two opposite wavelets essentially cancel each other when summed
together. (They are “offset” only by the 0.0004 sec of cumulative TWT.) Therefore,
no reflection is observed, even for this extreme velocity contrast. For more realistic
velocity contrasts, even-smaller-amplitude signals would be created, and the likeli-
hood of imaging the thin bed is essentially nil (any hint of a signal would be lost in the
noise).

By this argument, individual beds do not (normally) produce reflections. But seismic
sections (e.g. Fig. 1) clearly show patterns of reflections that correspond to our view
of the subsurface Geology. So, what do seismic reflections represent? The detailed
answer is more involved than this Chapter can address, but we can say that there is
more to the problem than merely the reflectivity of an ideal interface and raypath
vectors. The acoustic energy of a seismic signal is a wave, and full wave theory is
needed to determine how the vibrational energy moves through rocks, and how some
of it is returned to the surface as a received signal. Interferences arising from multiple
layers (real rock sequences are not single, ideal layers), and tuning effects, are part of
the story. Although the simple-interface approach that we have taken thus far does not
fully represent reality, it is a reasonable approximation, and we can continue using it
to gain further understanding of the seismic investigation process.

The thin-bed case outlined above illustrates the principles of tuning and interference
that are important for understanding how well seismic techniques can image a rock
layer that is being truncated. This situation is known as the “wedge” problem (Fig. 16).

18
Geophysics 5
In the example shown here, an upper horizontal layer overlies two layers below that
are tilted. The middle layer is truncated beneath the top layer, forming a wedge shape
(this situation might be an angular unconformity, or a depositional on-lap). A series
of synthetic seismic traces are created along the profile, showing clear images of the
top and bottom of Layer 2 away from its truncation point, but these distinctions
become lost as the truncation point is approached. Although the ability to “map” (see
below) the top and base of the middle layer is lost, it is nevertheless possible to detect
the disturbance of the seismic reflections and to infer the truncation position. Thus,
there is a difference between the ability to resolve a geometry, and the ability to detect
that a geometric change has occurred.

The Wedge Model Input

Layer 1: Vp = 2000 m/sec, 2.0 g/cc

Layer 2: Vp = 2500 m/sec, 2.5 g/cc

Layer 3: Vp = 3000 m/sec, 3.0 g/cc

Synthetic Seismic Traces


Distance

0.55

0.56
TWT (sec)

0.57

0.58

0.59

0.60

The Two Events Actually Observed


?

Note; Not Planar

Figure 16
The “wedge” problem

Department of Petroleum Engineering, Heriot-Watt University 19


1

4. REFLECTION SEISMOLOGY PRACTICES

A seismic investigation can be described in simple terms as follows:

1. Mechanical energy is injected into the ground (by means of a seismic source)

2. The acoustic waves that are generated travel through the rocks

3. Some of the energy is reflected from a variety of interfaces and returns to the
surface at varying times

4. There, the incoming waves are detected by sensors (geophones or hydrophones)


and the data recorded

5. Sophisticated computer algorithms manipulate these data and produce images


that convey the spatial distribution of physical properties of the rocks. These
distributions of properties are inferred to be directly related to the distribution
of the rocks themselves.

The first four steps are called acquisition, and the fifth step is called processing (and
this merges with a task called interpretation).

4.1 Seismic Acquisition


We might imagine an unrealistic Earth in which all rock layers are horizontal, with
horizontal interfaces. In that idealistic situation, we could place a seismic source and
a receiver at the same point on the surface, and these devices would propagate an
acoustic wave and then sense its reflections. The reflected wave energy would arrive
back to the receiver at times determined by the velocities of the rock layers and their
thicknesses (as in the example above).

In reality, the Earth is more complicated than this simple situation. (If it were not, then
we would have little need to conduct seismic surveys in the first place!) Acoustic
waves generated by a seismic source (often called a seismic “shot”) propagate in a
roughly spherical fashion, and they reflect off of a variety of surfaces that may not be
located directly beneath the source point. The returning energy is, therefore, scattered
over a wide area of the surface. In order to capture more of this returning energy, and
for other reasons that we describe below, multiple receivers are placed across an area
of the surface. Thus, the energy of a single seismic shot is usually recorded by multiple
receivers. A schematic drawing of this arrangement illustrates the main elements of
the acquisition process (Fig. 17).

20
Geophysics 5

Source Geophones / Hydrophones


Surface
Dynamite

Reflected Energy
Vibroseis

Rock Interface
Transmitted Energy

Air Gun

Sea Bed

Figure 17 Filter
Schematic illustration of Source Receiver
The Earth
seismic acquisition

On land, seismic sources can consist of:

• explosives (often called “dynamite”, even if other explosive materials are actually
used). These are usually buried in shallow - approximately 5 m deep - boreholes
that are called “shotholes”, or

• vibroseis (a special, very heavy truck that has a vibrating plate which is placed
against the ground).

In the marine environment, sources tend to be air guns (sometimes water guns). These
impart a short-lived mechanical pulse to the water, and the resulting shock wave
travels to the sea (or lake) bed, and then into the sediments and rocks below. Energy
reflecting from deep rock interfaces must also re-cross the water column (unless sea-
bottom receivers are in use; see below). Because the returning signal passes through
the water column, no shear waves can be recorded.

Each source location is called a shotpoint, regardless of the source type, and
independent of whether the survey is on land or under a body of water.

Air guns and explosives produce a seismic signal that contains many different
frequencies (from 10 Hz to more than 100 Hz). A vibroseis source also produces a
range of frequencies, although in this case, the truck vibrates the ground in such a way
as to “sweep” through a set of frequencies (perhaps from 20 Hz to 75 Hz) over a finite
time interval. It is the range of frequencies that exist in a seismic source that causes
a sharp return signal from a reflection interface (see discussion above).

Seismic receivers in the marine environment are called hydrophones (these detect
pressure changes caused by compressional waves, converting them to electrical

Department of Petroleum Engineering, Heriot-Watt University 21


1

signals; Fig. 18). A single streamer of 16-24 equally-spaced hydrophones (channels)


is common, but nowadays, multiple parallel streamers are used, producing many
100’s of channels of data. In the onshore environment, seismic receivers are called
geophones. Geophones are clamped to the ground and detect motion through
vibrations of a coil of wire moving through a magnetic field, producing time-varying
electrical signals (similar to the operation of standard microphones, as used in
telephones etc). Geophones are also placed in arrays, and one of the costs of onshore
seismic acquisition is the careful surveying of geophone and shotpoint locations.

Sea level

1 Hydrophones (channels) 16

Streamer
Source (plus additional paths
to each hydrophone)

Interface

Figure 18
Refractions
Marine seismic acquisition
Reflections

The configuration of geophone/hydrophone arrays is dependent on the design of the


particular survey. Two-dimensional seismic surveys (2-D seismic) are acquired by
widely spaced (100 m apart, and often much more) lines of receivers that are parallel
(or sub-parallel) to each other. The in-line spacing of receivers is, however, much
smaller (often 25 metres). Nowadays, it is more usual (for both development and
exploration) to acquire three-dimensional (3-D) surveys. In these, the spacing
between lines could be of the same order as the spacing of the in-line receiver array
(25 meters, or, increasingly, 12.5 meters). This arrangement allows the data to be
processed along the acquisition line orientation (in-line), and normal to it (cross-line),
providing high-resolution coverage of the subsurface (this is particularly useful in
resolving faults, Fig. 19). Horizontal sections (time slices) can also be generated from
3-D data cubes.

Figure 19
Example of improved
interpretation enabled by 3-D
seismic surveys and
improved processing

22
Geophysics 5
Seismic waves that are detected by each geophone or hydrophone are converted to
electrical signals. These signals are transmitted by wires (along what are called
seismic cables), or sometimes by radio, to a central control facility where they are
recorded. Occasionally, data are recorded onto a local tape or disk sited with the
receiver, and later assembled into a larger dataset. The start time for recording is the
time of the shot, and the duration of recording is anywhere from one or two seconds,
to perhaps as much as six or eight seconds. In order to ensure fidelity of the signal,
recording is universally digital (making use of analog-to-digital signal converters).
The digital sample interval is usually either 2 ms or 4 ms (two or four one-thousandths
of a second). At 4 ms, there are 250 data points per second, producing some 1500 data
points for a six-second recording. Each data point may be assigned two bytes of
storage (16 bits to record 65,536 different amplitude levels).

Because each receiver eventually records returning waves from many different
shotpoint locations (perhaps several hundred), and because there are many receivers
in a typical survey (at 25 m spacing, there are 40 each way per square kilometre, times
several tens of km2 in a typical seismic survey), the quantity of data is enormous. In
addition, vibroseis methods involve a source “sweep” that takes perhaps 20 seconds.
The receiver must record for this amount of time, PLUS the desired imaging time (e.g.
the six seconds in the above example). The costs of managing such huge data volumes
represent a significant factor in the decision to acquire seismic data, and they impact
the design of every survey that is undertaken.

Seismic acquisition is conducted by service contractors. Occasionally, a survey is


conducted by these companies “on spec”, where the contractor expects to be able to
sell the data at a later time to oil company purchasers. More often, an oil company
arranges for a contractor to conduct a survey for a specific purpose (upcoming lease
sale, development decision, production monitoring, etc). In all cases, the surveys are
conducted either with oil company representatives or consultants monitoring every
aspect of the operation.

Seismic surveys entail sophisticated logistical support and often are associated with
high mobilisation costs. Surveys need to be co-ordinated with other activities (in the
marine environment, this entails other seismic surveys, associated marine operations,
shipping, platform installations) that may be active within the area. The weather is
often a major consideration. Safety (in remote tropical or desert areas, and in the
deeper oceans) and environmental issues (environmentally sensitive zones such as
breeding grounds, locations where there is a risk of ground water contamination,
potential impact on fishing, etc) have to be considered in seismic surveying.

4.2 Seismic Processing


The aim of processing is to emphasise primary reflections (signal) and to reduce
spurious received energy (noise). (A view that is gaining popularity is that “noise”
is also signal - but a signal that we just don’t yet understand.) The energy of the seismic
waves that return to receivers is very small in comparison to the source energy. This
is because energy was lost during spherical radiation, and because only small fractions
of THAT energy were reflected back, and because THOSE reflections also radiated
spherically, and some of their reflected energy was reflected back down when they
crossed the interfaces in the rocks above.

Department of Petroleum Engineering, Heriot-Watt University 23


1

The strategy that has been developed to address the weak signals is to add together (to
“stack”) separate received signals (from different receivers), since this should
(hopefully) reinforce the true signal, and the noise would cancel out (being “random”).

Modern acquisition techniques are designed to capitalise on this notion. Consider the
geometry of shots and receivers illustrated in Figure 20. Multiple pairs of shot/
receiver locations generate reflections from the same subsurface point. This point,
midway between the source and receiver, is called the common mid-point (CMP).
Source Points Receiver Locations
7 6 5 4 3 2 1 2 3 4 5 6 7 Surface

Figure 20
Interface Illustration of Common
Mid Point geometry of
Common Mid Point source/receiver locations

Although each of the source/receiver pairs “images” the same subsurface point, the
paths of the waves for each pair have different lengths, which means that they travel
different distances. The longer distances represent greater times of travel, so the
notion of stacking the data together requires that an additional numerical manipulation
be undertaken to account for this effect.

Another example of the way that distance of travel impacts the time of a reflection can
be seen if we consider a single source point and multiple receivers arrayed beside it
(Fig. 21). Here, a shot sends a wave into the subsurface. The paths taken by the
wavefronts that return to the receivers are as shown. The source/receiver paths are
longer for greater distances on the surface between the shotpoint and receiver. (These
distances between source and receiver are referred to as the “offset”). In fact, there
is a well-defined relationship (known as “moveout”) between offset and path length,
and if the velocity is considered, between offset and TWT. This relationship plots as
a hyperbola whose form is a function of the velocity of the upper layer and the depth
of the reflecting interface. (A plot from a single source point shooting into multiple
receiver locations is known as a “gather”.) Because the form of this plot is sensitive
to these parameters, moveout plots are used to deduce subsurface velocities. In

24
Geophysics 5
practice, this is done by assuming a range of possible velocities, and comparing the
resulting observed moveout to the expected form, thus enabling a constraint to be
placed on the actual velocity of the overlying layer.

25m Ray Paths from Shotpoint 21 into Array of Recievers

15 16 17 18 19 20 21 22 23 24 25 26 27

100m

Layer 1: Vp = 1800 m/sec

Reflector

Layer 2: Vp = 2200 m/sec

Shotpoint Number
15 17 19 21 23 25 27
0.00

source point
TWT (sec)

0.10

Figure 21
Example gather illustrating
Moveout for a horizontal 0.20
reflector Gather for Shotpoint 21

The understanding of the moveout phenomenon gained for horizontal reflectors can
be extended to circumstances where the reflector is dipping (not horizontal). The
geometry of the wavefronts in this case (Fig. 22) depend additionally on the magnitude
of the tilting, and there is no simple plot to use for estimating velocities (because the
tilt is also not known). The complex wave-path geometry illustrated by this situation
necessitates a more-involved processing effort. In simple terms, the underlying goal
is to relocate the reflections to a position that is more nearly where the “should” be in
a distance-time plot - i.e. a normal reflection section.

Department of Petroleum Engineering, Heriot-Watt University 25


1

Ray paths from shotpoint 101 into array of receivers


25 m

95 96 97 98 99 100 101 102 103 104 105 106 107 Surface

100m

Layer 1: Vp = 1800 m/sec

Layer 2: Vp = 2200 m/sec Dipp


ing R
eflec
tor

Shotpoint Number
95 97 99 101 103 105 107
0.00

source point
TWT (sec)

0.10

Figure 22
Example gather illustrating
0.20 Moveout for a dipping
Gather for Shotpoint 101 reflector

The processing methods that are intended to accomplish such relocations of the
returned signals can be grouped under the heading of “migration”. Migration
algorithms require that velocity distributions be given as “input” data. Since velocities
are usually believed to be directly associated with the rock bodies, this task requires
that the answer be known before the answer can be derived. In practice, additional
information on velocities (such as from well logs), and other techniques that make use
of other aspects of the reflection data, can be combined in an iterative way to develop
good estimates of the velocity distribution. The end goal of a migration effort is to
produce a “depth” section - that is, one where the time axis is replaced by a depth scale.

One of the major successes of migration is to remove (actually, move to its correct
location) the diffracted data that arises at subsurface velocity discontinuities (such as

26
Geophysics 5
where beds are truncated by faults). The diffractions have the same hyperbolic form
as the moveout gathers, and they can represent a severe interpretation problem if they
have high amplitudes and thus obscure other data.

Above, we have emphasised the fact that the amplitude of seismic reflections will be
very small in comparison with the amplitude of the seismic source. Indeed, the
amplitude of later arrivals (reflections from deeper in the TWT section) is much less
than the amplitude of early (shallow) arrivals. In order to make interpretation easier,
seismic data are subjected to a gain adjustment (during processing) such that the
largest peak-to-trough amplitudes of each region are scaled to appear equal on the
final plot.

Another processing modification that is usually applied to the data is filtering. In this
activity, the recorded data are subjected to a frequency filter. The objective of this step
is similar to that of the gain adjustment: to make interpretation easier. Recall that
higher-frequency signals are degraded over distance more than lower-frequency
signals. Thus, data from deep in the section will (naturally) have a lower frequency
content. It is sometimes helpful to remove high-frequency returns from the shallower
portions of the dataset so that peaks and troughs have similar appearances regardless
of their location in time. Other reasons for performing filtering do exist, but we do not
need to address these specialist issues here.

Additional processing steps are needed to make other corrections to the data. For
onshore surveys, the locations of the shotpoints and receivers are never exactly as
planned. The actual positions need to be used to “move” the data to reflect reality. In
particular, the elevations of the sources and receivers are extremely critical. Such
alterations of the data are called “static” corrections. Direct arrivals (seismic energy
travelling along the surface, or through the water, from source to receiver) must be
removed, since these do not contribute useful information regarding the deep
subsurface, and their presence will detract from interpretation. In the marine
environment, water-bottom multiples (acoustic waves bouncing off the sea/air
interface, and then again off the water bottom) can be a serious interpretation problem
(because their amplitude can remain large). These various processing issues are noted
here to indicate some of the complexity that occurs “behind the scenes”.

5. INTERPRETATION OF REFLECTION SECTIONS

The main objectives of a seismic interpretation are to determine:

• two-way travel times and/or depths to rock interfaces

• the dip of interfaces

• the location of discontinuities in rock interfaces (faults)

• changes of stratigraphy (due to lateral facies changes) or unconformities

• rock and/or fluid properties from seismic velocities or impedances within layers

Department of Petroleum Engineering, Heriot-Watt University 27


1

These objectives can only be met if the acquisition and processing activities have been
performed in a way that produces a result that is suited to the task at hand. In reality,
interpretation often reveals aspects that can be used to return to the processing stage
to gain an improvement in the quality of the image. Such iterative loops can be highly
rewarding in terms of their resulting enhancements to the understanding of the
subsurface.

The usual end product of a seismic interpretation effort is a map, or series of maps, that
depict the shape(s) of geological horizons or the distribution of some characteristic
(such as high- versus low-amplitude reflectivity). Increasingly, these maps are being
used to develop a full 3-D representation of the subsurface (this is often called a geo-
model, or shared Earth model, when the regions between the surfaces are filled with
information about the rocks that are there). Such a model can be used to depict the
spatial variations in rock characteristics, including petrophysical properties (these
might be derived from seismic attributes; see below). The information from a model
of this sort can be used to undertake reservoir simulations, or basin-process simulations
(basin modelling).

Most interpretation work relies on the classical seismic reflection section (Fig. 23). In
this display, the vertical axis is TWT, while the horizontal axis is “distance”. (The
individual shotpoint locations shown on this axis can be related to a map.) Notice the
presence of white-black bands dipping to the left, away from the well location. These
bands are termed “events”, and they are composed of the troughs and peaks,
respectively, of adjacent seismic traces in which the reflections are at different TWT.
The close spacing of the traces (on the print) enhances the clarity of the events. Better
data quality, and improved processing, make interpretation easier (Fig. 24).

Figure 23
An example of a North Sea
2D line from 1966 showing
the quality of the data upon
which early, large
discoveries were made

Figure 24
An example of modern
seismic data from the North
Sea. This line is taken from
a 3D survey conducted in
1986. It is in the same
location as the line shown
in Figure 23
28
Geophysics 5
The basic task is to “pick” events on the seismic section (Fig. 25). At each shotpoint,
the TWT of these events is noted, and transferred to a map. These data are then
contoured to produce a “time map” (refer to Chapter 7). In modern practice, the
seismic events are usually picked with the assistance of a seismic workstation. This
computer has special software that “understands” seismic rules (basic interpretation
procedures are programmed in), and hence can perform what is called an “auto-pick”
of events (once the interpreter has made initial picks). The picks are followed
throughout the data volume (either manually or with the assistance of the workstation).
Discontinuities of the seismic events are usually interpreted as faults. The faults are
often mapped as separate surfaces, although sometimes they are merely treated as
gaps or breaks in the mapped horizon.

A
B C

Figure 25
Events that have been
picked on the seismic line
shown in Figure 24

In order to be useful for planning wells, or for calculating reserves, time maps need
to be converted to depth maps (actually, elevation maps; see Fig. 26). This task
requires that the velocities be known. As noted above, velocities are usually estimated
during seismic data processing. Additional information can be gained from other
sources. Well logs (in particular, the sonic log) measure the thickness and velocity of
rock layers. The interval velocities can be integrated to give an average velocity to
any depth in the well. However, there are uncertainties in this technique due to a
variety of small problems. A better technique that is available after a well is drilled
is called a check-shot survey (or sometimes, a velocity survey). In this technique, a
sonic source is placed at the surface, and a geophone is progressively lowered along
the wellbore to determine the actual average compressional velocity from the surface
to any depth.

Department of Petroleum Engineering, Heriot-Watt University 29


1

Figs 24 -26
A B C
Points A, B, C
refer to Fig 25

;;;;;;;;; Un

;;;;;;;;;
conf
or mit
y

;;;;;;;;;
Brent
D u nli n Fa
d
St atfjor u
lt

Structural Cross Section

Figure 26
Structural contour map of
the field over which the
seismic lines shown in the
preceding figures were
obtained

In the earlier stages of exploration and development, it may not be clear what the
various observed seismic events mean. In other words, we may not know if an event
represents a reservoir unit or a seal. If a well is near the seismic line, and its rock
succession has been interpreted, the sonic data from the well-log suite can be used to
create a synthetic seismogram (something like a sophisticated seismic trace). This
plot can be used to decide which events on the seismic section relate to which rock
units.

There are a few common pitfalls associated with converting time images to depth.
Unexpected lateral lithological variations (facies change from shale to limestone),
missed compaction trends (due to differential burial), or unidentified high-velocity
materials (e.g., salt, volcanics), or low-velocity materials (gas chimneys above
reservoirs), particularly in areas of sparse well control, can all lead to significant errors
in depth prediction. These, often understandable or unavoidable errors, can lead to

30
Geophysics 5
depth discrepancies - particularly in exploration - and are the reason for many changes
in drilling programmes and the outcome of wells. Luckily (or unluckily!) the target
can be both shallower, or deeper, than prognosed.

5.1 Other Seismic Methods

Direct Hydrocarbon Indicators (DHI)


Under certain (often optimal) conditions, the seismic data can give direct indications
of the presence of hydrocarbons. Free gas in the reservoir can give rise to large
impedance contrasts (“bright spots”) by reducing the impedance in the reservoir.
These, when recognised (e.g. Troll Field), can lead to very high licence bids by the
operators as they are evidence (pre-drilling) of significant hydrocarbon accumula-
tions. However, bright spots can be related to other phenomena - such as coals, tuning
effects, non-hydrocarbon gases or non-recognised, spurious noise - and thereby lead
the industry astray. A bright spot associated with a gas-water contact should be flat.
The effect of free gas on seismic is large and non-linear - small gas saturations may
have very large effects. Where fields are blessed with a DHI, caused by the gas-water
contact, these can be used to monitor production performance if subsequent surveys
can detect position changes of the DHI (e.g., Frigg Field).

Time-Lapse seismic
Repeat 3-D surveys after a passage of time are called 4-D, or time-lapse, seismic.
Time-lapse seismic is becoming increasingly used for reservoir monitoring. In some
fields, where there are DHI’s (e.g. Foinaven Field in the Atlantic Margins), Ocean
Bottom Cables (OBC) are being laid on the sea-floor to allow for repeat seismic
monitoring of the production. In this case, the reservoir sands are relatively shallow,
and the oil-bearing sand is clearly distinguishable from the water-bearing sand. The
fixed receivers eliminate some of the repeatability problems (resulting from different
processing and navigation schemes) between successive seismic surveys. They also
allow for the recording of shear-wave data. However, at this stage, OBC is a
significant investment with unproven, but potentially high, rewards.

Wellbore seismic
The Vertical Seismic Profile (VSP) is an extension of the velocity (or check-shot)
survey. The data are recorded in a similar way as the velocity survey except that an
array of geophones is used to record at many depths simultaneously, and the full
waveform is recorded. Arrivals can be separated as upgoing (from below the receiver)
and downgoing (from above it). The direct arrival is a downgoing wave.

The main uses of the VSP are:

• velocity calibration (as in a velocity survey)

• distinguishing primary reflections from multiple events

• prediction ahead of the bit (in detailed planning of complex high-angle wells, it can
be useful to further define the well path)

• imaging near-well-bore discontinuities - faults, salt wall margins - which may


influence sidetrack locations or production performance

Department of Petroleum Engineering, Heriot-Watt University 31


1

When the source is placed in one well and the receiver array is placed in another well,
cross-well seismic can be acquired. Tomography is an established medical technique
to reconstruct a section through a body from measurements around the body. In
seismic cross-well tomography, we can use a combination of well-to well, surface-to-
well and surface-to-surface techniques on reflection or refraction energy to image the
inter-well reservoir structure and properties. At a smaller scale, full waveform sonic
logging within a borehole (a horizontal borehole) can be used to image bed boundaries
and assess bed continuity.

Amplitude versus Offset (AVO), and Anisotropy


Seismic reflection data is normally presented as a stacked section. Any variation in
amplitude with offset (amplitude versus offset, AVO) is normally “lost” in the
processing (recall, this is the moveout correction). AVO anomalies can be caused by
the changes in Poisson’s ratio in the rocks on either side of an interface (see Rock
Mechanics Appendix). (The Vp/Vs ratio varies with the ratio of compressional to
shear velocity of the rock.) These variations are due to changes in the rock matrix or
the fluid content. AVO effects are recognised on the CMP gathers. AVO can be used
to detect fluid changes during production. This is a specialised technique with its own
costs and benefits, along with pitfalls.

Seismic wave velocity depends on the direction of propagation in anisotropic media.


An S-wave travelling through such a rock will split into two waves. Fractures produce
anisotropy, and shear-wave splitting can be used to determine fracture orientation and
intensity. There are additional techniques involving P-wave anisotropies, but these are
more specialised than we can cover here.

Seismic resolution
Seismic resolution is a function of the wavelength of the acoustic energy. At usual
seismic frequencies, the wavelength varies from about 30 m (shallow depths) to 300 m
(deep in basins) in typical hydrocarbon systems. Seismic resolution (the ability to
determine the top and base of a bed by observing distinct peaks or troughs - the events)
is usually taken to be equal to a quarter of a wavelength. Therefore, resolution ranges
from about 8 to 60 m. The resolution range of a variety of subsurface acoustic tools
is given in Figure 27. Seismic detection (of a bed, or changes between surveys in 4-
D seismic) does not require imaging of the top and base of beds, and therefore can be
expected at thicknesses up to one thirtieth the wavelength (if the impedance contrast
is sufficiently large). If this rule holds, then detection might be possible for beds
ranging from 2 to 10 m in thickness (refer to discussion above regarding the “wedge”
problem).

Attributes
Above, we have commented on the way that various “anomalies” affect the quality
of seismic data. We have noted that changes in fluid content (as might happen in one
place during the production history of a reservoir, or differences between places, such
as the hydrocarbon and water legs of a reservoir) can affect the seismic signal. We also
noted that variations in cementation and porosity can have subtle impacts on the
seismic data. Historically, the goal of processing has been to eliminate these effects,
so as to produce a “cleaner” section for interpretation. Increasingly, methods are being
developed that capitalise on these changes in an effort to identify the spatial locations

32
Geophysics 5
of different rock and fluid types. The characteristics of the seismic data that can be
used for making these distinctions are grouped together under the name “attributes”.
It is now common to see seismic data being processed to emphasise such variations,
and for maps to be created showing their spatial patterns. The premise of this work
is that there is a direct correlation between an attribute and the properties or
characteristics of interest in the rocks. The term attribute can also be applied to derived
parameters (e.g. attribute 1, plus attribute 2, divided by attribute 3). The growing
significance of seismic attributes “proves” the adage that “noise is merely signal we
don’t understand”.

Range of Typical Vertical


Investigation Frequency Resolution
(metres) (Hz) (mertes)

Earthquake seismology 10000000 2 1000


BOREHOLE

Seismic refraction 25000 10 100

Seismic reflection 5000 30 20

Vertical seismic profiling 2500 60 10

Crosswell tomography 500 1000 1


SURFACE

Figure 27 Wireline sonic logs


1 15000 0.1

Comparison of scale and


Acoustic image logs
resolution of various 0.05 1000000 0.001

acoustic techniques

6. SUMMARY

In this Chapter the Engineer will have learned that:

• the seismic method is the study, by means of imposed acoustic waves, of interfaces
between beds or formations of contrasting impedance, and that the reflection and
refraction of energy from these interfaces allows for the mapping of these surfaces

• seismic data allow the interpretation of surfaces in time. Well data are needed for
accurate depth conversion

• seismic resolution is a function of frequency bandwidth (wavelength)

• there are a variety of surface and borehole techniques and, with time-lapse
operation, these can provide cost-effective reservoir monitoring

Department of Petroleum Engineering, Heriot-Watt University 33


1

In the “old days” (many decades ago), seismic methods were very direct: source and
receiver were located together, and each receiver trace was simply plotted next to its
neighbours. The resulting sections were not very good, but the interpreter could be
confident of how the data were collected. Now, with the stacking of data from
multiple receivers, and the associated requirement for sophisticated processing,
seismic sections can seem slightly mysterious, and rather removed from any potential
for a reality check. Some people have turned this situation into an opportunity for
humour (Fig. 28), implying that seismic data are purely imaginary. The amount of
Industry money expended on seismic, and the successes that have resulted from its
use, indicate that this technique is of crucial importance for exploration and
development. Petroleum Engineers should expect to be involved with Geophysicists
at an increasing level throughout their career.

Figure 28
Tongue-in-cheek view of
the seismic method

Acknowledgements

The authors acknowledge Mike Cox, Philip Ringrose, Robin Westerman and Colin
MacBeth who provided assistance with this Chapter.

7. EXERCISES

1. This exercise expands on the understanding developed in the Chapter regarding the
creation of synthetic seismic traces. The subsurface configuration of a hypothetical
region is given by Figure 29. Note that the layers are horizontal, but that they have
velocities that vary from one side of the section to the other. The task is to create a
trace for each location, A and B. This requires that you create a table for each location
in which the depth, velocity, interval TWT, and cumulative TWT are calculated. Then
assess the reflection characteristics of the interfaces (positive or negative). Place a
wavelet at the correct TWT for each trace, and, draw correlation lines between the
wavelets.

34
Geophysics 5

several km
A B

500 m
600 m
Vp = 1800 m/s Vp = 1900 m/s

Vp = 2000 m/s Vp = 2100 m/s


200 m 700 m

Vp = 1700 m/s Vp = 1750 m/s

Figure 29 Vp = 1900 m/s Vp = 1950 m/s


Simplified seismic section
Vp = 2200 m/s Vp = 2300 m/s
for Exercise 1

2. In this exercise, you will determine depths from a time section (simplified). You
are given a “seismic section” in Figure 30. This section is, as usual, in TWT. The main
reflectors are both peaks and troughs, representing positive and negative reflection
coefficients, respectively. Estimated P-wave velocities are given for each interval for
locations X and Y (assume that these were derived from moveout gathers). Your task
is to generate a depth cross section. This requires that TWT be converted to interval
TWT, and this converted to interval thickness. Use the interval thicknesses to
construct the cross section.

5 km

X Y
Vp = 1800 m/s
0.4

Vp = 2000 m/s Vp = 1900 m/s

0.6
Vp = 1700 m/s Vp = 2100 m/s

Vp = 1900 m/s Vp = 1750 m/s


0.8 Vp = 1950 m/s
Vp = 2200 m/s
Vp = 2300 m/s
Vp = 2300 m/s
Figure 30 1.0
Cross section information Vp = 2400 m/s
TWT (sec)

Major Reflection Events, and Velocities, as Given


for Exercise 2

Department of Petroleum Engineering, Heriot-Watt University 35


36
Solution to Chapter 5
1
Exercise 1 (Given data in bold)
Well A
Interval Thickness (m) Interval Velocity (m/s) Time interval (one way) Time interval (two way) Depth(m) Time (s)

500 1800 0.277777778 0.555555556 500 0.555556


600 2000 0.3 0.6 1100 1.155556
700 1700 0.411764706 0.823529412 1800 1.979085
200 1900 0.105263158 0.210526316 2000 2.189611

Well B
Interval Thickness (m) Interval Velocity (m/s) Time interval (one way) Time interval (two way) Depth(m) Time (s)

500 1900 0.263157895 0.526315789 500 0.526316


600 2100 0.285714286 0.571428571 1100 1.097744
700 1750 0.4 0.8 1800 1.897744
200 1950 0.102564103 0.205128205 2000 2.102873

Exercise 2
Well X
Interval Thickness (m) Interval Velocity (m/s) Time interval (one way) Time interval (two way) Depth(m) Time (s)

432 1800 0.24 0.48 432 0.48


110 2000 0.055 0.11 542 0.59
93.5 1700 0.055 0.11 635.5 0.7
57 1900 0.03 0.06 692.5 0.76
110 2200 0.05 0.1 802.5 0.86

Well Y
Interval Thickness (m) Interval Velocity (m/s) Time interval (one way) Time interval (two way) Depth(m) Time (s)

541.5 1900 0.285 0.57 541.5 0.57


115.5 2100 0.055 0.11 657 0.68
87.5 1750 0.05 0.1 744.5 0.78
68.25 1950 0.035 0.07 812.75 0.85
115 2300 0.05 0.1 927.75 0.95
Geophysics 5

Well A Well B
Q1

1.0
TWT
(Secs)

2.0

Q2 X Y

250

Depth
(M)

500

750

1000

Department of Petroleum Engineering, Heriot-Watt University 37

You might also like