Anda di halaman 1dari 242

INTRODUCTION

Today, in mature petroleum provinces, we seldom or never find classical anticlines with simple four-way
closure. All those structures have been found long ago. Nowadays, our standard fare consists of structures
bounded in part by dip and in part by faults. We must become expert in recognizing such structures, and in
deciding whether or not they have potential as hydrocarbon traps.
Conduits and Seals
Where a crack occurs in the earth, we usually call it a fracture. Where the rock layers are displaced across
the fracture, we usually call it a fault. The importance of the fault, as distinct from the fracture, is that the
fault has greater potential to form a trap ( Figure 1 ).


Figure 1


At the time of cracking, the fault is likely to provide an escape for water (and possibly hydrocarbons)
contained in the rocks below; it provides a conduit. For the fault to form a trap, however, the fault must
provide a seal.
An attractive situation arises where a fault is a conduit in its lower reaches, and a seal in its upper reaches.
This may happen, for example, if the conduit formed by the crack allows the escape of mineral-rich waters
from below, and the conditions of chemistry and temperature cause the minerals to precipitate out of
solution as the water rises in the fault. If a seal forms just above a faulted reservoir we have a trap, while if
the conduit remains open below the reservoir we have a path by which the reservoir may be charged from
deep source rock.
The danger in a fault trap, of course, is that the seal may be broken by later reactivation of the fault, so that
the hydrocarbons are lost.
Sometimes, therefore, we need faults to act as conduits, allowing the charge of a reservoir from a deep
source. In other situations, we need faults to act as seals. And our constant enemy is the breach of a fault
seal by reactivation of the fault.
In Figure 1, we depend on the fault to act as a conduit, to charge the reservoir from the deep source rock.
We depend on the offset cap rock at A to provide an updip barrier. But we also depend on a fault seal at B. If
this seal does not exist, we have nothing. if it does, the spill point, C, depends on the thickness and offset of
the cap rock.
In Figure 2, with a thicker cap rock, we do not need the fault seal at B, but it must be present higher in the
section, perhaps at D.


Figure 2


If it is not present by E, the trap cannot be charged unless both the unconformity surface and its overlying
rocks are impermeable-and then only after the development of this impermeability. If the fault has been
reactivated as suggested at F, any charge depends on the subsequent sealing of this breach, and can
represent only hydrocarbons generated since the resealing.
We see, then, the critical decisions we must make in fault interpretation. Is it possible for the fault trap to be
sourced? Does the fault plane seal? Did the seal develop before or after the generation and migration of the
hydrocarbons? Where the thickness and offset of the cap rock imply a spill point, is the trapped volume
sufficient to warrant drilling? Has the fault been reactivated? Is it likely to have resealed? When? Were
hydrocarbons still migrating after this time? Having regard to the burial history, should we expect oil or gas?
In general, there are uncertainties in our answers, and fault traps always carry an additional element of risk.
Some decision-makers, as a matter of principle, will not drill downthrown fault traps, or traps where faults
can be seen extending to the surface. Others point to the large number of proven fields having just these
conditions. As always, the drilling decision should be based on a reasoned assessment of potential reward
and potential risk.
One other seismic decision is very relevant to the assessment of risk: is there evidence of any mobile rock
above the trap, in which any fault displacement could be absorbed? Thus, the presence of thick salt above
the unconformity in Figure 2 would virtually remove any risk from fault reactivation (though in Figure 1 it
might not, because of the silt layer). Similarly, a thick overpressured shale is often seen to absorb quite
major faulting. Where the salt or shale layer is thin, however, the fault may be transmitted; then we face the
judgment of whether it is likely to be sealed by flowage of the mobile material into the fault zone.
All in all, therefore, we cannot afford to dismiss fault traps, but working with them will often intensify our
problems-and occasionally cause us sorrow.
Interpretation, Past & Present
In the past, a great deal of seismic fault interpretation has been grossly in error. Some of the reasons for this
are understandable.
With the poorer quality of old data, reflection continuity was often interrupted by
seismic artifacts-changes in surface conditions, refraction blind spots, incorrect
statics. To a person raised among mountains, every such break was a fault.
In the regrettable days when geologists and geophysicists were kept apart, many geophysicists
did not even know what the geologists knew about the nature of faulting.
What explorationists know about basin dynamics and fault generation has increased enormously in
the last decade.
There is always a temptation to give the seismic data the weight of a measurement, while
regarding the geological input as mere generalization. Today, we resist this, and insist on an
interpretation that satisfies both the seismic measurement and the geological generalizations.
Perhaps the clearest example of this is that we would no longer accept any fault interpretation,
however "obvious," if it did not satisfy conservation of mass before and after faulting. One of the
cheapest ways to find oil today is to reinterpret the old fault maps over producing oil fields in the
light of present geological knowledge, and to discover pockets of oil never tapped by the drill.
Round and Round
In modern practice, therefore, we are much concerned with establishing the tectonic regime of the area-and
how that regime has changed over time-before we attempt final picks on the faults. Here we put ourselves
into an interpretive loop: we cannot pick the faults until we know the tectonic history, but we learn the
tectonic history, very largely, from the faults. So no fault can be picked in isolation; we search first for faults
that show the tectonic message most clearly, we make regional and local inferences about the history, and
we go back into all the faults until a coherent pattern emerges.
2-D versus 3-D Seismic Data
Fault interpretation on 2-D seismic data is much less safe than on 3-D data. One reason for this is that 2-D
lines are seldom sufficiently close to give definite fault correlations from line to line. Another is that fault
interpretation is jeopardized whenever a 2-D line is not perpendicular to the fault. However, we must still be
ready to do the best we can on 2-D data.
Fault interpretation on 3-D data has given us new insight into the geometry, distribution and behavior of
faulting. Once, where we mapped a few long continuous faults using 2-D data, we now see more numerous
short, discontinuous faults separated by some sort of transfer zone.
Similarly, mapping around faults has changed. The use of 3-D data has allowed us to see the details of
structure along and near the fault planes. We have learned that structural contours at the faults are not
necessarily the smooth and flowing artistic lines that so many of us were taught in the past.
General Issues in Interpretation
Figure 1 illustrates the fault-picking operation at a simple level.


Figure 1


We join the apparent breaks or steps in the reflections with a geologically plausible fault trace, continuing the
pick upwards and downwards until the faulting becomes mere flexure, or disappears. Simple enough. But
even in this simple case, we are looking for additional messages. Above the fault we are always vigilant for
evidence of reactivation (possibly in the opposite direction), because such reactivation may have broken a
seal. However, we also remember the possibility that mere compaction, particularly in shales, can suggest
minor reactivation of a fault where none has occurred. Further, the figure illustrates that either type of
movement can produce significant local changes in stratigraphy above a fault zone.
Figure 2 goes to an opposite extreme; it shows us that, even with fair record quality, fault interpretation can
sometimes be extremely problematical.


Figure 2


In the face of the changing intervals between major reflections, the requirement for geological plausibility
becomes the key; whatever picks we make, we must have an explanation for them. Just joining the
reflection breaks in some arbitrary manner will not do.
Even when the fault interpretation appears simple, we have to satisfy ourselves that other faulting is
compatible with that interpretation. Are the other faults of compatible type and the same age (acceptable),
or of different type and different age (acceptable), or of different type and the same age (problematical)?
As we have agreed earlier, many of the problems of fault interpretation on 2-D data arise because the lines
are not at right angles to the faults. This problem is made worse by the fact that many areas have been
subjected to several tectonic episodes, each generating a system of subparallel faults-but at different
orientations. Figure 3 is a fairly simple illustration of this, with a system of curved extensional faults crossed
by a system of vertical strike-slip faults.


Figure 3


If we lay out a rectangular grid of 2-D lines over such an area, some or all of the lines must be at
unfavorable angles.
Figure 4 illustrates fault picking at a basin edge.


Figure 4


We can see very readily the likely geological development, and the picking is straightforward. However, the
earth is also capable of some exceedingly grotesque contortions, and sometimes our courage is taxed (
Figure 5 ,


Figure 5


Figure 6 , and Figure 7 ).


Figure 7





Figure 6


In some cases, situations that would otherwise tax our courage are eased when we can recognize the
presence of mobile rocks such as salt or overpressured shale. Thus, the major salt-induced fault in Figure 8
gives us few problems.


Figure 8


An essential step in fault interpretation is the recognition of episodes of extension, of neutral subsidence, and
of compression. Thus, in Figure 9, we recognize an orderly subsidence at the edge of a rift basin, and the
repeated reactivation and adjustment of the main basin-bounding fault.


Figure 9


On the other side of the basin, Figure 10 shows the formation of a major left-tilted faulted block as a
consequence of extension at rift time, before the filling of the basin; such blocks, of course, may represent
attractive petroleum targets.


Figure 10


We would also read extension (followed by stable subsidence) into the tilted fault blocks of Figure 11.



Figure 11.


It is equally important that we recognize the evidence for compression. At a regional scale, this may be fairly
clear, and confirmed by the mountainous nature of the surface. Occasionally we may even see such thrust
situations at prospect scale ( Figure 12 ); the associated rollover can yield a very important structural trap.


Figure 12


A more complex example, involving both compression and extension, is given in Figure 13.



Figure 13.


Many petroleum provinces have been subject to episodes of both extension and compression. When these do
not act in the same direction, some sideways movements and adjustments occur, and these create
distinctive fault patterns, both in plan and in section. Figure 14 braces us for the difficult decisions that we
shall have to make, when we are working in such provinces.


Figure 14




Stress
A rock is said to be in a state of stress when a force is applied to it. Earth stress is not simple, but involves
total force-per-unit-area for a particular point. What makes the measurement of stress complex is the variety
of magnitudes and directions of force that occur within a single volume of rock. Basically, however, we can
think of stress as the three-dimensional intensity of force acting at a specified point.
At any moment in its total history, a rock has structure. It is also true that for any such moment, it is subject
to stresses that tend to alter its properties. In reality, only in deep space would this same rock be relatively
free of stress. On earth, the forces that compose stress can be divided into two main typesbody forces,
which act at every point within the crust and surface forces (also called applied forces), which act only at
interfaces between objects and, therefore, are defined only along surfaces.
We can better understand how these forces act by considering a single clastic grain buried within the earth's
crust. Two types of body forces act on this grain at all timesgravity and inertia. We usually ignore the
effects of inertia. The force due to gravity is called the lithostatic pressure. It results from the simple weight
of overburden transferred by grain-to-grain contact. If our particular grain becomes involved in deformation,
two types of surface forces will also exert their influence on it normal forces, which act perpendicular to the
surface of the clast, and shear forces, which act parallel to these same surfaces.
Figure 1 (Components of stress on a single grain), shows diagrammatically the breakdown of all these forces
that compose the total stress on this single grain.


Figure 1


Since gravity and inertia, by definition, act on every particle at every point in the crust, it is the surface
forces that are primarily responsible for the creation of geologic structure and with which we as
explorationists are concerned.
A rock undergoes deformation when stress causes the displacement of particles within it. This stress can
result from body or surface forces, or both. In nature, deformation is almost never simple, since it results
from a complex interaction between the chemical and physical properties of a rock mass, its immediate
environment, and the rate and intensity at which the displacing forces are applied.
We commonly make use of the concept of a stress field which refers to the distribution of stress acting within
a defined body. Such a body can be a single folded layer or a sizable portion of a continent. A stress field is
described as homogeneous if the stress at each point is equal. The only situation that normally approximates
this near the earth's surface is when stress is almost totally due to lithostatic pressure. "Approximates" is the
appropriate term to use here, since there is nearly always some component, however small, of lateral stress
in the crust.
As a measure of force-per-unit-area, stress is a vector quantity and may be expressed as the sum of various
components. If we trade our clastic grain for a hypothetical cube of uniform composition and subject this
cube to a progressive deformation ( Figure 2 , Resolution of unidirectional force (F) acting on a cube face into
the basic normal (N) and shear (S) components of stress.


Figure 2


Also shown is the idealized physical effect of each component), we see that the total force (F) can be
resolved into contributions that act perpendicular to the faces of the cube (N, or normal stress), and parallel
to them (S, or shear stress). Normal stresses can be either compressional and tend to compact, or tensional
and tend to separate particles within a body. The effect of shear stresses, on the other hand, is to move
particles past one another (as in the bottom right of Figure 3 , Generalized orientation of the principal axes of
stress for three basic geologic structures.


Figure 3


In (a) and (b), horizontal compression can be thought of as the actual applied force, while (c) shows the
effect of horizontal extension). During deformation, the relative magnitudes of these two stress types will
change, such that one type may decrease as the other increases.
Principal Axes of Stress
Material behavior science, through its detailed analysis of stress conditions, has derived an important
conclusion that has proven very useful in structural geology: for any point in a homogeneous stress field,
there exist three mutually orthogonal planes along which all shear stresses vanish and only the components
of normal stress exist. These three planes are known as the principal planes of stress, and the axes of their
intersection are thus the principal axes of stress. These axes are used to describe what are referred to as the
three principal stresses. This ideal triaxial system makes everything simpler, since it allows us to speak in
terms of only normal stresses, i.e., compression, or "squeezing," and extension, or "pulling apart" ( Figure 3
).
In geology, principal stress is usually spoken of in terms of compression, which is taken as positive, and
tension as negative. (For materials and engineering science, the opposite is true-i.e., tension is positive,
compression is negative.) The three principal axes, or stress directions, are correspondingly written as 1
(maximum principal stress), 2 (intermediate principal stress), and 3 (least, or minimum, principal
stress). For our purposes, it is useful to understand four special states of stress:
1. uniaxial stress, where two principal stresses are zero and the other is nonzero
2. biaxial stress, where two principal stresses are nonzero and the other is zero
3. triaxial stress, where all three principal stresses are nonzero
4. pure shear stress, where 1 equals 3 and is nonzero, while 2 is zero. This is actually a special
case of biaxial stress.
Within the earth's crust, the most common stress situation is triaxial, with 1 >2>3>0.
In terms of actual physical effects, al can be thought of as representing the relative compression that acts to
deform a rock body by compacting particles into each other, while a represents the relative tension and the
stretching of particles away from each other. We say "relative" because these stresses are often defined only
in relation to each other. For example, a strong extensional event in the crust will have a al associated with it
that is secondary i.e., the compression is passive, related mainly to overburden pressure in this case.
The relationship between the three principal axes of stress can be pictured by considering the triaxial
examples shown in Figure 3 , which place our hypothetical cube in evolving fold and fault structures. Again,
1 represents the direction of maximum effective compression; 3, extension. At this point, we should not be
concerned with the specific angle that the faults make with respect to these directions of stress. Note,
however, that these faults and 2 are normal to the plane of the paper. In physical terms, 2 is parallel to
the surface of the earth and basically represents the presence of essentially infinite neighboring material.
Thus, with these three axes, the conditions of stress at any point or as geologists often apply it at a certain
location within a defined plate tectonic regime, can be described and related to actual geologic structures.
Though there are no shear stresses acting on the surfaces of the cube in our example, the fact that 1, 2,
and 3 differ in magnitude and direction means that shear stress can be resolved within the cube. In fact,
the quantity 1 -3 (1 minus 3), called the differential stress, is sometimes used as a general indicator of
shearing stress. If we were to measure the shearing stresses generated within our cube, we would find that
they reach a maximum along planes that are inclined at 45 degrees to 1 and 3 and intersect in the plane
of 2 ( Figure 4 ) This represents the ideal case, and (as we will see later on) it helps explain why rocks do
not fracture randomly when stressed to the point of rupture.


Figure 4



Strain
Strain is said to exist when particles within a body have undergone displacement. Specifically, strain, usually
denoted as 1, compares the change in length, L, of any line element in a rock body to its original length, L,
which increases in some directions and decreases in others. Strain is a measure (qualitative or numerical) of
this displacement, which, if permanent, is called deformation. Changes in shape resulting from strain are
called distortion; those in volume are described as dilatation, which can be either positive (in expansion) or
negative (in contraction or shrinkage). Strain can result from both body and surface forces.
Although strain is commonly conceived of as the effect of stress, stress and strain are inseparable during
actual deformation. Furthermore, if we make a close comparison between Figure 1 (Computer-generated
stress field for a hypothetical fold in a less viscous matrix.


Figure 1


The short lines represent 3 axes) and Figure 2 (Computer-generated strain pattern for Figure 1 .


Figure 2


Short lines are drawn perpendicular to the axes of maximum shortening), we can see that the specific
relationship between stress direction and resulting strain is complex, even in simple situations. We can see,
for example, that the orientations of the short lines in these figures coincide least within the flanks of the
folded layer, where shearing stresses are greatest.
As we move up or down the flanks, toward the "hinges" of the fold, we notice that the attitude of the strain
lines must be explained in terms of a growing combination of normal and shearing stresses. Thus, we would
need to understand precisely the mechanisms involved in the transformation of stress into strain before we
could accurately predict how these two vector quantities are geometrically related. As a result, we often
introduce helpful simplifications into the analysis of strain in nature. The more important of these are based
on the assumption that the strain involved in deformation has been relatively homogeneous.
Much of the terminology derived for understanding stress has also been applied to strain. For example, our
hypothetical cube is said to have suffered homogeneous strain, also called uniform strain, when the strain is
the same at all points within it. This means that originally straight lines remain straight after deformation (
Figure 3 Homogeneous).


Figure 3


Thus, for example, a cube becomes a rhombohedron, while a sphere inscribed within it becomes an ellipsoid.
In fact, when we describe strain, we often do so with reference to an imaginary object called the strain
ellipsoid. In this strain ellipsoid, the principle axes of strain are denoted as 1, 2 and 3 where the maximum
lengthening is designated 1 and the maximum shortening as 3.
This should also help make clear the basic concept of inhomogeneous strain ( Figure 3 , Inhomogeneous).
Inhomogeneous strain is by far the most common in naturally deformed rock. This type of strain involves
some amount of rotation in the position of particles, which means that originally straight lines become
warped and detailed analysis becomes impracticable. It is, therefore, almost always useful to find some way
in which natural deformation can be approximated as homogeneous. The most common approach is to
consider geologic structures as the summation of many localized homogeneous strain fields. This method has
proved especially helpful in the explanation of secondary rock fabrics, such as mineral alignment and
fracturing.
Pure and Simple Shear
There are a number of basic ways in which deformation by homogeneous strain at constant volume occurs.
Those that involve simple flattening and stretching are shown in Figure 4 (Uniform extension),


Figure 4


Figure 5 (Unifrom flattening),


Figure 5


and Figure 6 (Plane strain: Each are basic types of homogeneous strain imposed on a cube of ideally uniform
composition.


Figure 6


In each case, the inscribed circle and ellipse represent cross sections through the strain wllipsoid before and
after deformation). To understand how we treat natural deformation, however, it is also necessary that we
look at the two basic types of shear strain-pure shear and simple shear ( Figure 7 , Hypothetical cross
section and diagram to illustrate domains of pure and simple shear in a series of folds that show
progressively greater strain.


Figure 7


The shape of the strain ellipse can be the same for either type of shear and cannot be used to derive detailed
strain history). Both types help us explain a great many large and small scale features seen in rocks.
Pure shear is a form of strain in which no rotation of the strain axes takes place. It is also often referred to
as an irrotational deformation. It results from uniform extension in one direction and contraction
perpendicular to it ( Figure 8 ).


Figure 8


In simple shear, all particles within a body are displaced in one direction. This is our cube pressed into a
rhombohedron again; this time, however, we need to take note of the rotation in the strain ellipsoid. Simple
shear can be visualized by imagining the result of placing our cube with its inscribed sphere between the two
surfaces of an active fault. The shearing motion created by these two surfaces stretches and flattens the
sphere into a strain ellipsoid whose long axis is progressively rotated until it is nearly parallel to the fault
plane itself. Displacement within such a body takes place by slippage along closely spaced planes ( Figure 8
).
In actual materials, this can be accomplished in a number of ways for example, by slippage between grains
or crystals, or by actual flow at elevated temperatures and pressures. As we shall see, this style of
deformation has widespread application to geologic structure.
Rock Strength, Brittleness & Ductility
A substantial amount of literature exists with regard to experimental rock mechanics in the laboratory, the
intent of which is to simulate and analyze the process and effects of deformation. Studies are usually
performed on cylindrical, core-like samples, which are subjected to compressive or tensile stresses in a
chamber whose pressure and temperature are regulated. From these studies, scientists have determined that
deformation generally progresses through three main stages (Nadai, 1950). These stages are defined by the
behavior of the deformed material, and are most clearly and simply shown by the use of stress-strain
diagrams ( Figure 1 ).


Figure 1


In order, the three stages are as follows:
1. Elastic During this initial stage, the stress-strain relationship is linear. If stress is
removed, the body reverts to its original dimensions. No deformation (permanent
strain) results; strain is said to be completely reversible ( Figure 1 , segment A).
2. Plastic As stress continues to increase, it will eventually reach some limit beyond which the body
suffers permanent strain. This is termed the elastic limit, or yield stress. Beyond this limit, material
is said to behave plastically: any increase in stress brings a corresponding increase in deformation (
Figure 1 , segment B).
3. Rupture With continued increase of stress, the body will eventually fracture, rupture or fault (
Figure 1 , point C).
Elastic strain can also end in rupture; a material simply fails without having undergone any
measurable deformation.
Models of Deformation
These three stages through which deformation normally progresses are idealized as the behavior of three
hypothetical "bodies" that are subjected to stress, as shown in Figure 2 ,


Figure 2


Figure 3 , and Figure 4 (Stress/strain relationships for several ideal materials: (a) Hookean (elastic) body;
(b) St.


Figure 3


Venant (plastic) body; (c) Newtonian (viscous) body).


Figure 4


Each style of behavior is denoted by the name of a well-known mathematician and should be apparent from
the stress-strain graphs shown. A Hookean body knows only elastic strain before rupture, and is
approximated by a simple elastic spring attached to a fixed body. A St. Venant body, in contrast, shows
elastic strain up to a yield stress and then deforms indefinitely by shear strain at that same stress. This type
of behavior is approximated by a weight that is pulled across some surface by an attached spring; the spring
stretches elastically up to the point where friction on the table top is overcome and the weight begins to
slide. The last example, called a Newtonian body or, sometimes, Newtonian liquid, has no shear strength at
all and therefore exhibits no elastic strain. It deforms by what is called viscous strain. In Figure 4 , this type
of body is represented as a porous piston pulled through a fluid-filled cylinder. A Newtonian body, then, will
deform indefinitely in response to any shear stress, with the total strain being directly proportional to the
amount of elapsed time.
Strain Hardening
These three types of ideal bodies help describe the components in progressive deformation when it is caused
by increasing amounts of stress. We might compare them to the graphs shown in Figure 5 , which describe
two generalized stress/strain relationships for known materials.


Figure 5


The principal difference between the ideal
( Figure 2 , Figure 3 , and Figure 4 ) and reality ( Figure 5 ) lies in the behavior known as strain hardening.
Such hardening is often the result of complex readjustment, or even recrystallization, of the material
suffering strain. We can envision it on one level as being due to the compaction and realignment of particles
during progressive deformation. In sandstones, for example, stress concentrations along grain boundaries
result in extensive grain fracturing and dissolution. This causes the progressive filling of pore space with
grain fragments and recrystallized quartz. All other conditions (e.g., temperature and pressure) being equal,
this will increase rock strength-that is, greater and greater amounts of stress become necessary to impose
the same increment of deformation ( Figure 5 , upper curve). We expect certain clastic lithologies to respond
in this manner.
On the other hand, strain hardening can occur up to some ultimate strength, after which the stress necessary
to cause a given strain decreases continually ( Figure 5 , lower curve). Lithologies such as salt and gypsum-
anhydrite may behave this way under certain conditions.
Brittleness and Ductility
On the basis of the differences between elastic and plastic behavior, we are able to characterize the general
response of materials as either brittle or ductile.
Brittle materials rupture before any significant plastic deformation occurs. Such behavior in rocks is marked
by the development of breakage discontinuities along the planes that represent maximum shear strain.
These are not necessarily faults or fractures visible to the unaided eye, but may take place between
individual grains or within crystal lattices. In contrast, material is described as ductile if it is able to undergo
a large amount of plastic deformation before failing.
It should not be assumed that brittleness guarantees faulting, or that ductility inevitably leads to folding. In
every situation, whether a rock responds in a brittle or ductile manner depends on several parameters-
composition, effective confining pressure, temperature, strain rate and anisotropy. It is, in fact, relatively
meaningless to speak of the true "strength" of a rock without reference to these parameters.
Figure 6 ,


Figure 6


Figure 7 ,


Figure 7


and Figure 8 ,


Figure 8


(A series of stress/strain diagrams showing the effects of confining pressure ( Figure 6 ), temperature (
Figure 7 ), and strain rate ( Figure 8 ), on rock deformation on homogeneous (Solenhofen) limestone. Note
for ( Figure 8 ): a strain rate of 10-7 per sec is roughly equivalent to 1 year of time; most geologic
deformation occurs at strain rates of 10-14 per sec) indicate how laboratory analysis has shown deformation
to vary as a function of pressure (a), temperature (b), and strain rate (c) on homogeneous limestone.
These laboratory tests tell us that near the surface, at low temperatures and pressures, rock will tend to act
in a more brittle manner. With growing overburden, which increases both pressure and temperature, ductility
generally increases. At the same time, however, the element of time is crucial. If small amounts of stress are
applied over sufficiently long periods of time, almost any rock will deform plastically; only those lithologies
with very low resistance to shear, such as basalt, may not. But given normal rates of tectonic deformation,
there is some transitional depth range over which the response of a particular rock type will grade from
dominantly brittle to ductile.
Figure 9 is a schematic representation of the spectrum from brittle to ductile behavior in limestone, as
determined by laboratory tests.


Figure 9


Such testing applies a uniaxial stress (either compression or tension) to a cylindrical sample that is sealed
within a chamber whose temperature and confining pressure can be regulated. Limestone and marble have
been favored as samples, since these lithologies are more isotropic than clastic rock types.
In addition to pressure, temperature and strain rate, the other factor that determines a rock's response to
stress is, in one sense, the most obvious-its composition. Figure 10 is a diagram derived by Handin et al.


Figure 10


(1963) from extensive lab experiments on natural rock. Shown are measured ductilities for four common
sedimentary lithologies in a water-saturated condition.
One interesting point made by this diagram is that, with only one kilometer of burial, marked differences in
ductility already exist between limestone and other lithologies. This is partly the result of pore pressure
effects that strongly reduce ultimate strength. It is also, however, directly related to the mineral structure of
calcite, which allows for significant intracrystalline gliding, even at moderate pressures and temperatures.
The same trend of ductility increase is true for sandstones, and is often aided by pore pressure effects. We
might also note that dolomite, while slightly more ductile than quartzite at depths below two kilometers,
shows the tightest range of permanent strain before rupture. Due to its mineral structure, dolomite does not
deform readily by intracrystalline gliding.
What is normally referred to as "rock strength," then, is not a fixed property, but a relative response,
determined by a specific set of immediate environmental conditions. Since both engineers and geoscientists
make use of the terms compressive strength, tensile strength and shear strength, two major points should
be made in relation to these qualities. First, almost all materials are weaker under tension than compression
( Table 1). And second, it is very often the shearing strength of materials that determines when and how
they will fail.


Average crushing
strength
Tensile
strength
Shearing
strength
Sandstone 740 10-30 100-200
Limestone 960 30-60 150-250
Quartzite 2020 30-90 150-250
Granite 1480 30-50 150-300
Basalt 2500 - 50-150

Table 1: Measured strength of various rock types at standard temperature and pressure
Principal Stress Directions and Faulting
Figure 1 ,


Figure 1


Figure 2 ,


Figure 2


and Figure 3 display the idealized orientation of the principal stress axes during reverse (a), strike-slip (b)
and normal (c) faulting. Because faulting and fracturing both represent the brittle rupture of rocks, they are
often discussed in relation to similar stress systems.


Figure 3


Figure 4 shows the generalized orientation of actual fractures formed in a experimental triaxial test, in which
a block of Solenhofen limestone was shortened by 1 percent at room temperature.


Figure 4


The two categories of fractures that explorationists usually speak of, shear fractures and extension fractures,
were generated by this test and are indicated. This is how rock generally ruptures in the laboratory. Note
that a broad correspondence exists between ideal faulting and shear fracturing with relation to stress
orientation. As we have previously noted, rocks are weakest in shear.
Two mutually orthogonal types of extensional ruptures are shown in Figure 4 . These have been related to
the stresses generated during loading and unloading, and, in each case, they develop normal to 3. Notice
that during unloading, 1 and 3 exchange the orientation that they had during loading. This tells us that
faults can conceivably develop at very low angles to 1, and that we should expect the relaxation of tectonic
stress to generate late-stage features, especially fractures, in a rock body.
More generally, it can be understood from Figure 1 , Figure 2 , and Figure 3 that a region undergoing a high
degree of faulting is characterized by numerous local stress fields that change in both orientation and
magnitude as diastrophism progresses. The geometry and attitude of particular fault planes, therefore, may
not always appear directly related to the stress axes that apply to large-scale structural trends. The stress
orientations shown in these figures do not, by themselves, always explain the specific angles at which faults
develop or the shape of fault planes. For example, many fault planes are curved and cannot be explained
simply in terms of one stress axis alone we need several at varying depths.
At smaller and smaller levels of scale, many faults reveal increasingly complex components of displacement,
usually involving shear. couples. Figure 5 shows the detailed resolution of shear fractures in an evolving
monocline. The fractures shown are those which immediately precede propagation of a high-angle reverse
fault from basement into the overlying strata.


Figure 5


While in their basic geometry monoclines represent one of the simpler geologic structures, the details of the
deformation associated with them are not simple at all.
We should expect, then, that rocks in the vicinity of a fault or fault zone will be complexly sheared for some
distance on either side of the fault plane itself. This distance may be measurable in centimeters to
kilometers, depending on the lithologies, the type of fault, and the amount and environment of displacement.

General Terminology
Much of the basic nomenclature relating to faults was derived from coal mining in the British Midlands during
the late eighteenth and early nineteenth centuries. In fact, the word "fault" itself was originally used by
miners to describe the sudden, unexpected, and vexatious termination of a coal seam. Thus, "fault" then
carried much of its vernacular sense-some sort of mistake had been made.
Early geologists like Murchison and Lyell, however, were quick to realize that a fault was, in reality, a fracture
where displacement had occurred, and that simple geometric methods could be used to predict where a
particular seam might be found again, either above or below. The bounding surface of a fault presented a
"wall" to the disgruntled miner, who was normally forced to continue his heading a short way into solid rock
and then start a new shaft in order to relocate the seam. The wall of the fault plane was almost always
inclined, which meant that the miner could hang his lamp from the rocks on one side above the fault and rest
his foot on those on the other side below it ( Figure 1 ).


Figure 1


Thus, the terms hanging wall and footwall then, as now, simply label the two sides of the fault and imply
nothing about displacement. Due to its more common occurrence in the coal-mining area of Britain, a fault
was normal to the miner's experience when it was inclined toward the hanging wall. A coal seam that ended
against such a normal fault could be found again if the miner continued his tunnel a short distance in the
same direction and then sunk a shaft downward. At times, the reverse was true, and the corresponding fault
was designated as a reverse fault. These terms remain in use today.
By modern definition, rocks are said to be faulted when they have suffered observable displacement along a
plane or interval of rupture. As noted previously, such rupture occurs mainly by shear. The fault plane can be
relatively simple ( Figure 2 )


Figure 2


or it may consist of a large number of individual offset surfaces and thus be more accurately described as a
fault zone ( Figure 3 ).


Figure 3


(In the true mathematical sense, the term "plane" is incorrectly used, for rarely does a fault exist without
horizontal and vertical kinks, bends and changes of direction. Because of this, some of our colleagues protest
the use of "fault plane" and instead support the term fault surface. Although technically incorrect, fault plane
is more commonly used, so that is the term we will use in this text.)
Less frequently, rocks may be displaced by a form of shearing that causes loss of internal cohesion but not
actual rupture of lithological layers ( Figure 4 ). We use the term shear zone to label fault zones in which the
individual planes of displacement are extremely closely spaced.


Figure 4


The great majority of faults, however, more closely approximate planes of slip along which shearing has
taken place as a result of movement.
Faults are generally classified on the basis of their relative sense of displacement. Yet, as for most of
geology, certain settings have encouraged the development and use of more specific terminology, often
related to proposed mechanisms. Each major type of fault discussed immediately below is given a more
specific nomenclature when discussed in relation to a particular structural style.
Displacement varies along a fault, being greatest near the middle of the fault and decreasing to zero at both
ends. If we contour the amount of displacement, the resulting map describes a displacement ellipse. We can
construct displacement ellipses in both plan view ( Figure 5 ) and along the fault trace.


Figure 5


Net slip along a fault is measured by a vector that traces the displacement between originally adjacent points
( Figure 6 , Basic terminology for fault offset, showing strike-slip (ss) and dip-slip (ns) ).


Figure 6


It is most often resolved into dip-slip (measured in the dip direction) and strike-slip (measured in the strike
direction) components. Faults in which one or the other of these components is dominant are
correspondingly named dip-slip faults and strike-slip faults.
For nonvertical dip-slip faults, we find it useful to again divide the displacement into vertical and horizontal
components ( Figure 7 , Frequently used terminology for faults in cross section: (a) definition of throw,
heave, and hade.


Figure 7


(Note: the latter two terms are more common in mining geology); and (b) definition of normal stratigraphic
separation). From coal mining terminology, the former is sometimes called the throw, the latter, the heave,
and the angle between the fault plane and the vertical, the hade. We frequently refer to the vertical and
horizontal separation across a dip-slip fault, while the inclination of the fault plane is simply said to be either
high-angle or low-angle with reference to a horizontal datum. Also, we often use log data to determine the
normal stratigraphic separation, as shown in Figure 7.
When "Throw" is Not Throw
The definitions noted in Figure 7 are fine when the beds are horizontal. However, when they are dipping, we
must introduce an additional term, called vertical separation, and distinguish it from throw.
Figure 8 shows a block diagram of a normal fault in which the formations are dipping from right to left.


Figure 8


In this figure, we can see that when we project dipping beds across the fault plane, we can measure the
displacement in two ways-we can project the bed across the fault perpendicularly to the vertical plane, or we
can project the bed in the same dip angle across the fault. The former is the throw, the latter is defined as
the vertical separation.
There are only two cases where the vertical separation equals the fault throw. The first is the case as shown
in Figure 7 in which the stratigraphic units are horizontal-i.e., zero dip. The second is when the fault is
vertical.
When we measure the fault "throw" using well logs, we are actually measuring the vertical separation.
However, when we measure the fault "throw" using a seismic section, we are usually measuring the throw,
although we are actually measuring the apparent throw.
The apparent throw is a measure of the fault displacement as a function of the angles between the direction
of the seismic line and the strike of the fault as well as the dip of the beds. Thus, if several seismic lines are
oriented at different angles relative to the strike of a fault, each line would yield a different value of the fault
displacement. The fault throw," as measured on a seismic line, is the true fault throw only if the direction of
the seismic line is perpendicular to the fault strike and the bedding horizontal.
Ignoring this question of relative orientation, we can use seismic data to determine the vertical separation
(actually apparent vertical separation) to allow us to integrate seismic measurements with well log
measurements. Figure 9 (Determination of vertical separation by projecting dipping reflections across the
fault plane), shows that we can do this by projecting the interpreted horizons across the faults.


Figure 9


Back to Displacement
Figure 10 (Major types of dip-slip faults.


Figure 10


H and F refer to hanging wall and footwall) shows the four types of dip-slip faults. In most cases, it is
assumed that simple shear has acted as the principal strain within the fault plane ( Figure 11 , Diagram
illustrating the progressive simple shear within an ideal fault plane ).


Figure 11


We should note here that the fault type may not always be obvious. Where separation cannot be determined
or appears to change along strike, determining the nature of the fault can be difficult. In some regions, a
diversity of interrelated faulting styles exists in close juxtaposition; differentiating specific structures can
prove very difficult.
In addition, many intracontinental basins are characterized by near-vertical, high-angle faults of small
displacement that cannot be easily identified as normal or reverse. Furthermore, multiple episodes of
tectonism often affect a single region. These episodes may involve contrasting stress regimes, and
displacement along early faults can be reversed. One of the more important occurrences of this phenomenon
involves provinces initially dominated by normal faulting being subjected to compressive stresses as a result
of changing plate boundary interactions.
Normal Faults
Normal faults can either be planar or listric (concave upward). On seismic sections, planar faults often appear
curved, due to the effect of increasing velocity with depth. In general, however, normal faults are more
easily identified on seismic sections than other types of faults. This is because of their frequent occurrence in
deep, dominantly marine basins characterized by otherwise relatively undeformed sediments.
Normal faulting on a regional scale is most often referred to as block faulting, since it results in the creation
of horst (high) and graben (trench) structural topography ( Figure 1 ).


Figure 1


Refer to Figure 2 ,


Figure 2


Figure 3 , Figure 4


Figure 4


, and Figure 5


Figure 5


.


Figure 3


The normal faults in these figures are clearly visible. In these seismic sections, it is relatively easy to
correlate the reflection horizons across the faults and distinguish the hanging wall from the footwall, as well
as the (apparent) throw in time. (The throw in depth may be a little more difficult.)
Thrust Faults
A thrust fault has often been defined as a reverse fault dipping less than 45 degrees. However, today we use
"thrust" as a generic term to imply near-horizontal, tangential compression and a zone of movement
dominated by simple shear strain.
In Figure 1 and Figure 2 , we see seismic examples of thrust faulting.


Figure 1


In Figure 1, the effects of compression are documented by the presence of older rocks overlying younger
ones.


Figure 2


We noticed the "rollover" of beds caused by the thrusting in Figure 2.
Thrust faults can (and do) dip at any angle, and individual thrust planes often show complex, curving
geometries. They may be essentially horizontal (along incompetent bedding planes), or flat, for kilometers,
then "ramp" to a higher structural level in competent rock and, finally, when reaching another incompetent
bed, flatten out again .
Thrust faults may curl up at their termini to become nearly vertical or even overturned. In the latter case,
thrusts become apparent "normal faults," geometrically speaking.
Often, due to complexity, it is very difficult or even impossible to distinguish between underthrusting, in
which the lower, relatively undeformed "block" is active, and overthrusting, in which the upper, deformed
block actually moves. We may only know that shortening is involved, and this shortening can be as great as
tens of kilometers or more for single faults.
Rocks that have been transported from their original location (or root zone) are said to be allochthonous
("other earth"), while those that remain in place are called autochthonous ("same earth"). The allochthon is
also variously known as a thrust sheet or plate (not lithospheric), or, in certain circumstances, a nappe.
Some thrusts define a listric plane that flattens with depth while others are of a ramp-and-flat geometry.
Several or more subsidiary thrusts commonly occur within a single allochthon, either as splays off the sole
fault (the deepest and controlling major thrust fault in an area) or as local ruptures in the cores of anticlines.
In general, the thrusts in foreland belts are very rarely simple listric planes, but are themselves folded and,
at times, truncated by younger thrusts.
As discussed by Dahlstrom (1970) and Elliot (1976), the geometry and location of thrust faults in thick
sedimentary sequences is largely determined by the distribution of competent and incompetent layers. Four
simple rules summarize this influence:
1. Thrusting cuts up-section in the direction of displacement. (This is often called the
direction of tectonic transport or, in older nomenclature, the facing direction.)
2. Thrusting tends to parallel bedding in incompetent layers, occurring near contacts with
competent units, and to cut obliquely up-section in thicker, more brittle units.
3. The age of major faults is younger in the direction of thrusting.
4. Major thrust faults do not overlap appreciably.
Evolution of Thrusting
The overall evolution, therefore, is for thrusting to begin at deeper levels and to progress upward and
outward (i.e., away from the root zone) from a major sedimentary basin. Ductility contrast in the
stratigraphic section encourages a stair-step evolution of thrusts, which we know as ramping. A relatively flat
portion of a thrust plane, particularly where it remains parallel to bedding within a single lithology, is referred
to as decollement. If this flattening of the thrust plane occurs at the boundary between the sedimentary
section and basement, it is called basal decollement.
The progressive stacking of thrust faults may develop in piggyback fashion, where younger thrusts form in
the footwall or, alternatively, in overstep fashion, where thrusts become younger toward the root zone.
Piggybacking seems prevalent on a regional scale and is by far the more significant progression. At the same
time, both piggyback and overstep thrusting occur on a more local level, often as a result of imbrication.
Imbricates occur most often in two structural positions of high stress concentration-near the toe of a major
thrust and above ramps in a thrust plane. They dip steeply as they approach the surface, and stack slice
after slice of the same stratigraphic section along faults, which sole out into a major thrust plane. Continued
movement along this plane after the imbricates have formed will rotate them, so they can become vertical
and overturned.
As discussed by Dahlstrom (1970), imbrication actually offers a basic model for foreland thrusting-as the
scale of a cross section is increased to become more regional, major thrusts themselves become imbricates
of the largest faults (i.e., those with the greatest displacement). These, in turn, can be thought of as
subsidiary faults to a basal detachment or decollement plane that marks the structural boundary between
basement, usually crystalline metamorphic or plutonic rocks, and sedimentary cover.
Thin- versus Thick-Skinned Tectonics
The concept of basal decollement is the fundamental structural principle in the hypothesis known as thin-
skinned tectonics. This theory is often contrasted with the thick-skinned hypothesis, which postulates no sole
fault and, there-fore, direct involvement of basement in each major thrust.
The debate between these two schools of thought is a historical one that continues today. On the basis of
drilling and seismic data, both of which have proved the flattening of thrusts at depth in foreland areas, most
explorationists now favor the thin-skinned hypothesis for at least the more medial and distal portions of
thrust belts. However, toward the metamorphic core of many such belts, basement rocks are known to be
heavily involved in thrusting. This involvement can be very extensive, as in the case of the Himalayas, and
may, in fact, control the overall style and evolution of thrusting.
Yet some recent studies based on deep-reflection seismic profiles (Cook et al., 1979; Cook, 1982) have
strongly favored the thin-skinned hypothesis for basement thrusting as well. Thus, the debate has expanded
to focus on two major questions whether thick-skinned faulting occurs at all in foreland belts, and, if it does,
what is the nature of the transition between it and the decollement tectonics that characterize the
sedimentary cover. The controversy represents one of the major areas of research in contemporary
structural geology.
Strike-Slip Faults
Strike-slip faults can be simply classified as left-lateral (also called sinistral) or right-lateral (also called
dextral) on the basis of the displacement sense ( Figure 1 ).


Figure 1


When facing the fault, the direction in which the opposite side appears to have moved indicates this sense. In
cross section, the symbols "A" and "T" are used to mark which side of the fault has been displaced away from
and which side toward the observer.
The San Andreas ( Figure 2 ) is one of the best examples of an active right-lateral strike-slip fault.


Figure 2


Others include the Alpine Fault of New Zealand and the Atacama system of Chile. Left-lateral faults of this
scale are well-represented by the Philippine Fault ( Figure 3 ). These large strike-slip faults are often called
wrench faults in general, and transcurrent faults more specifically if they cut across regional structural
trends.


Figure 3


Transform faults are strike-slip faults that connect convergent and divergent plate boundaries. Basically, they
serve to "transform" the interaction between plates into strike-slip motion. The motion along these faults
often includes components of compression or tension. Transform plate boundaries, therefore, are the links
that unify the world's spreading centers, subduction zones and collision zones into a single mosaic of
movement.
In Figure 4, Figure 5


Figure 5


, and Figure 6 , we see the variety of structural types associated with large strike-slip faults.


Figure 4


Later in this text, when we discuss strike-slip faults and their relationship to structural styles, we will see
several examples of these structures in seismic section.


Figure 6



Simple Fault Traps
More often than not, structural traps containing hydrocarbons result from or are associated with some form
of faulting. Although there are numerous, and sometimes unique, types of fault traps, most generally have
an appearance similar to one of the simple models shown in Figure 1 and





Figure 2.



Figure 1




Introduction
What is a fault, to the seismic method?
A single, clean fault, in massive layers, has the effect of replacing one continuous
reflection by three discontinuous ones, all of the same reflection coefficient (
Figure 1 ). Clearly, the best measure of the throw of the fault is given by the two
stratal reflections, while the best measure of the location of the fault is given by
the fault plane reflection.



Figure 1


In modern practice, we are much concerned to record the fault plane reflection,
and so to obtain this precision of fault location. Traditionally, however, the
recognition of faults on seismic sections has relied on the "terminations" of the
stratal reflections at the fault. These terminations, are not abrupt; however
abrupt the termination of the stratal reflector, the corresponding stratal reflection
smoothly tails away into a curved diffraction. Because of this, the actual fault
location on a seismic section is often difficult to pin-point.
Of course, it is the business of the migration process to collapse these
diffractions, and so to restore abruptness to the faulted reflection But we also
remember that 2-D migration cannot be perfect. In the interpretation of faults,
therefore, we are well advised to use both the migrated and unmigrated sections.
For this we need to be comfortable with the generation and meaning of
diffractions.
Faults and Diffractions
We need only Huygens' principle to tell us to expect a diffraction from a fault. In
Figure 1 (Looking at diffractions in terms of Huygens sources), at point A on a
continuous reflector, one Huygens source interacts with those each side of it to leave
only the reflected signal at normal incidence.


Figure 1


At B, the presence of the last Huygens source at the edge tells us that a signal must
be observed in the direction C, while the absence of further Huygens sources to the
right tells us that the "backward" radiation in direction D will not be "canceled."
Exactly the same message is conveyed, though in a different form, by considering the circular
reflection zone on the reflector ( Figure 2, Looking at diffractions in terms of the relfection zone).


Figure 2


The reflected signal at surface point P includes contributions from all of the reflection zone R. That
at point Q includes contributions from only a semicircular zone S, so that we would expect only half
the amplitude. The diffracted signal at point T includes contributions from a zone that gradually
decreases in size as the offset increases.
The classical seismic expression of a fault is obtained by summing these contributions. For the
assumed model of a terminating reflector, the result is as shown in Figure 3 (Seismic response to a
terminating reflactor)-a plane reflection that decreases in amplitude to one-half at the fault, and a
hyperbolic diffraction with its apex at the fault.


Figure 3


The diffraction comprises a forward branch of the same polarity as the reflector and a backward
branch of the opposite polarity.
The physical necessity for the two branches of opposite polarity is easily shown by bringing
together two such models, as in Figure 4.



Figure 4.


In the limit, when the reflector becomes continuous, the diffractions must cancel to zero.
The diffraction amplitudes make sense. Since the reflection melds smoothly into the forward
branch, and since the amplitude of the reflection at the fault is one-half that for a continuous
reflection, the amplitude of the forward branch near the fault is one-half, and therefore so is that of
the backward branch. In this sense, the half-amplitude of the reflection at the fault may be viewed
as the whole amplitude minus the backward branch. Therefore, canceling the backward branch with
the forward branch (by summing along the diffraction hyperbola, as in migration) not only removes
the entire diffraction but restores the reflection amplitude at the fault.
The above model of a terminating reflector is geologically unreal; a reflector cannot just
terminate. There must be an additional contribution from the fault face. However, for a vertical
fault, it is still true that there must be forward and backward branches of equal amplitude and
opposite polarity, to provide cancellation when two blocks are brought together in the manner of
Figure 4 , (The physical necessity for backward and forward branches, as the gap between reflector
terminations decreases) .
One situation that is close to the classical model is that of a vertically faulted thin horizontal layer
( Figure 5, The differentiated form of the classical response, obtained from a thin layer (here
assumed to be hard) ), where the signal from the fault face is very small.


Figure 5


We remember that if the layer is thin, relative to the seismic wavelength, the reflection response
approximates to a differentiated version of the incident pulse; then the forward and backward
diffractions also approximate to differentiated versions of the classical response.
The response changes if the fault plane is inclined. Although the diffractions remain grossly
hyperbolic, the pulse shape and the amplitude of the forward and backward branches may no
longer be the same. However, as suggested in Figure 6, (If the fault is inclined the details of the
diffractions change,


Figure 6


but bringing the fault blocks together (a) still requires cancellation of the forward branch of one by
the backward of the other (b) ), it is still necessary that the forward branch from the left block
should cancel the backward branch from the right block (and the reverse) when they are brought
together; therefore, these must have the same amplitude and form, but opposite polarity. The
reasonableness of this is confirmed when we remember that the tails of the diffractions contain the
fault plane reflections, which must also cancel.
If we complete the fault, as in Figure 7 (The reflection from the inclined fault plane in (a) is found
among the tails of the forward branch of the upblock and the backward branch of the downblock
(b).


Figure 7


The impression is therefore of stronger diffractions on the downthrown side), there is ordinarily a
strong forward branch from the upthrown block and a strong backward branch from the
downthrown block, and the lateral offset means that these must approach each other. Again, these
tails contain the fault-plane reflection.
The backward branch from the upthrown block and forward branch from the downthrown block
ordinarily remain well separated, as in the figure. However, if there is a rapid increase of velocity
with depth, the lower diffraction is flatter, and the two could again approach each other. In this
case the zone of confluence would contain the reflection from the underside of the fault
planeobtainable because the rapid increase of velocity with depth turns the downgoing rays
upward again.
The above observations apply to sharp rupture of the reflector at the fault... as occurs with brittle
rocks. With more plastic rocks, there is likely to be drag into the fault This may be normal or
reverse, and may be modified by reactivation or by isostatic rebound (footwall uplift). For any type
and degree of drag, we can synthesize the response using Huygens hyperbolas.
For normal drag into an inclined fault, as suggested in Figure 8 (Likely appearance of a normal
fault with normal drag),


Figure 8


a common effect is that only the forward branch from the upthrown block and the backward branch
from the downthrown block remain visible ( Figure 8 ).
Similarly, the dominant diffraction from the layers pierced by a salt plug is often the backward
branch of reversed polarity ( Figure 9, The termination of a stratal reflector (a) at a salt plug often
yeilds (b) an amplitude increase caused by the concave focusing,


Figure 9


and a diffraction dominated by the backward (whose tail contains the reflection from the salt-
positive or negative, as the case may be).
The preceding figures illustrate the diffractions obtained when the faulted layers are horizontal. If
the faulted layers dip, the diffraction hyperbolas are tangent to the reflection at the point of
apparent faulting, and have their maximum amplitude at this point; then the true position of the
fault is through the apices of the diffraction hyperbolas ( Figure 10, The situation where the faulted
layers dip, and the distinction between apparent, unmigrated and migrated positions of the fault
plane ).


Figure 10


The figure also shows that the apparent position of the fault plane, joining the terminations of the
stratal reflections, does not correspond to the unmigrated position of the fault-plane reflection,
which is contained within the diffraction tails. The fault pick that we make on the unmigrated
section-from diffraction apex to diffraction apex-is the migrated position of the fault.
On an unmigrated section, an inclined fault-plane reflection must always appear to cut across the
reflections it truncates; Figure 11 (The fault plane corresponds (approximately) to the apices of the
diffractions. The fault-plane reflection is contained within the tails of the diffractions, and therefore
must cut across stratal reflections) is a case in point.


Figure 11


A full hyperbola without any change of polarity at the apex ( Figure 12, The message of a full
hyperbola with no change of polarity is a tight syncline-not in itself a fault ) generally indicates a
buried focus (often, a tight syncline)-not a fault (though it may also signal a shallow source off the
line).


Figure 12


In principle, diffractions can help us to distinguish between normal and reverse faults, in that the
diffraction seen most clearly has the same polarity as the faulted reflection for a normal fault, and
the opposite polarity for a reverse fault ( Figure 13 , Signature of an inclined normal fault-plane
reflection (for a positive stratal reflection).


Figure 13


(b) The corresponding reverse fault, with its different-polarity diffractions melding into a negative
fault-plane reflection.). In practice, we have our usual problems in being sure of pulse polarities.
Diffractions cease to be hyperbolic if there is a lateral change of overburden velocity ( Figure 14 ,
Lateral change of velocity (introduced, in this figure, by dipping layers) distorts the diffraction form
hyperbolic form). (Of course, so does NMO.) The apex of the diffraction curve may no longer
correspond to the fault edge.


Figure 14


Diffractions generally survive the stacking process quite well near their apices, though their tails
may be weaker and lower-frequency than would be expected from simple Huygens synthesis.
Perfect and Imperfect Migration
The ideal solution to diffractions is perfect migration, which collapses each of them to its corresponding
reflection termination, and thereby restores the full amplitude of the reflection at that point. Drag into faults
is correctly represented, and flexures not involving actual rupture are distinguished from faults ( Figure 1,
Unmigrated) and Figure 2 , Migrated).


Figure 1





Figure 2


However, 2-D migration can be perfect only if the strikes of all stratal reflectors and of all faults are
perpendicular to the seismic line. This requires first that the reflector strikes and the fault strikes should be
the same (which happens, but not always), and second, that the dip lines were laid out straight and on dip.
In this happy situation, our only remaining problem is that the dip lines must be linked into strike lines-for
which 2-D migration must be in error. We also know that 2-D migrated sections do not tie at the line
intersections, and that we must demigrate the dip lines to achieve this fundamental and powerful check.
However, we can work with 2-D migrated sections on such a grid and produce a correct map.
But if the strikes are not perpendicular to the lines, that our general approach would be different. Our
preference would be to do the reflection picking on the unmigrated data, to proceed to an unmigrated
contour map, and then to migrate the map. The problem with this, of course, is that we cannot afford to
sacrifice the clarification of the faults provided by section migration where it is valid. Interpretation therefore
becomes an iterative process:
We make a first assessment of the fault correlations from line to line, and of the
gross reflection contours.
We thereby identify those faults on which migration is likely to be correct.
We use the migrated sections over these faults to clarify the fault interpretation.
We improve our picking of the unmigrated sections accordingly.
We proceed to an improved unmigrated contour map.
We migrate the map.
As a check, we can see whether the migrated map explains the invalid migrations that must
be apparent on the migrated sections crossing faults and other features at unfortunate angles.
To this end, we need to be able to think out the signatures of incorrect migration. As the
reference, the diffraction in Figure 3, from a fault perpendicular to the line, is perfectly
collapsed (b) by migration with the correct velocity.


Figure 3


The undermigration in (c) can be caused only by the use of too low a velocity. The
overmigration of (d) can occur either by the use of too high a velocity or if the fault is not
perpendicular to the line. The compact mustache of Figure 4 (e), superimposed on genuine
reflections in an impossible manner, indicates a small source off the line (in marine work, a
shipwreck or extrusion on the sea floor); migration turns the moustache into horns.


Figure 4


A more extended pattern, such as (f), indicates a fault edge locally subparallel to the line, but
again not in the plane of section ( Figure 5 ); the event is a sideways diffraction, having a
travel time that must not be interpreted as vertical.


Figure 5


If the fault edge is not parallel to the line, very odd events appear on the section (g). The
application of migration to these events is clearly wrong; it finds apparent dips that originate
merely in the varying distance of the fault edge from the line, and steepens them, in a manner
that is quite false.
Figure 3 (d) reminds us that when a fault (or other feature) crosses the line at an angle other than 90
degrees, the correct migration to a sharp edge can be obtained by using a lower migration velocity. If the
angle between line and fault is 0, we need to decrease the migration velocity by the factor sin 0. In particular
cases, where the details of one fault edge are critical, or where the whole section has the same strike, it may
be legitimate to use this artifice. We must remember, however, that the migration velocities then fail to tie at
the line intersections. And, in general, the failure of reflection times to tie at the line intersections is made
worse.
Fault-Plane Reflection Coefficient
Fault interpretation is much more positive if we have fault-plane reflections as well as stratal reflections.
Fault-plane reflections are extremely desirable.
There are two basic problems. The first is concerned with the reflection coefficient of the fault plane, and the
second with the seismic problems of recording very steep dips.
Do They Reflect?
Figure 1 [(a) A potential fault trap, and (b) the reflection coefficient of the fault plane] illustrates the problem
of reflection coefficient.


Figure 1


In (a) we see a potential trap in a limestone reservoir, and in (b) we see the reflection coefficients produced
by the fault offset . . . zero, strong positive, zero, weak positive, negative, strong negative and zero. The
length of each reflecting segment, in the figure, depends on the thickness of the layer and the throw of the
fault. Significant lengths of zero reflection occur if the layer thicknesses exceed the throw. So, in general, a
fault-plane reflection contains positive segments, negative segments, and zero segments. This does not look
promising. When we remember that only reflectors of a large area can return strong reflections, it looks even
worse.
In just two situations might we expect significant fault-plane reflections.
The first occurs where the layers are thick and the throws are large. Figure 2 (The
type of fault-plane reflection generated when the layers are thick and the throw is
large) gives two unmigrated illustrations.


Figure 2


In this situation, the sign of the reflection coefficient may be either positive or
negative, and is different for normal and reverse faults.
The second does not demand thick layers, but occurs where the throw is so large that the fault
juxtaposes layers of markedly different age or consolidation ( Figure 3 , The type of fault-plane
reflection generated when the throw is sufficient to juxtapose rocks of vastly different age and
compaction).


Figure 3


In this case, the reflection coefficient may be uniformly positive; all layers in the downthrown block
are so much younger than their juxtaposed counterparts in the upthrown block that the effect of
consolidation outweighs the variations in the layering itself.
In fact, fault-plane reflections seem to occur more often than we would expect from either
situation. In Figure 4 (Fault-plane reflections generated in a sequence of thin layers faulted
with small throw; squinting along these reflections reveal surprisingly constant polarity),


Figure 4


for example, some of the fault-plane reflections seem to be more continuous and stronger,
and of more uniform polarity, than we would expect from the sand-shale sequence involved.
This raises the possibility that the fault-plane reflection may be affected by more than the
simple acoustic contrast between materials across the fault. Specifically, it may be telling us
something about the nature of the fault about the width of the fault zone, or the degree of
pulverization, or recrystallization under pressure, or cementation, or perhaps even the
saturant in the fault zone.
This raises our interest, because it contains the possibility of a seismic distinction between a conduit and a
seal.
Conduits and Seals
The fact that faults sometimes act as conduits, over periods long enough to generate large accumulations of
hydrocarbons, tells us that the enormous pressures acting to cause intimate closure of a fault are not always
sufficient to do so. Therefore, we can visualize the situation shown in close-up in Figure 5 (A close-up of a
fault-plane area across which there is intimate contact over only a fraction of the fault-plane),


Figure 5


where the particles that are in contact across the fault, though individually under great stress from the
weight of the overburden, make up no more than some fraction of the fault-face area. What is the effective
reflection coefficient of this?
Everything depends on the saturant. If it is water or oil, the downgoing seismic wave scarcely sees the break
. . . the fault-plane reflection is negligible. But if the void is full of gas, the local reflection coefficient of the
rock/gas interface is almost -1, while that of the rock/rock contacts is zero or small; the overall effect is a
significant negative reflection coefficient whose magnitude depends on the fraction of the fault-face area
having solid rock-to-rock contact. There is no difficulty in visualizing values for this fraction which would
account for a significant reflection.
This gas-induced negative reflection does not depend on the throw of the fault or the layer thickness (as do
the classical fault-plane reflections we considered earlier). It can persist over the entire length of the conduit
with unchanged polarity (though we can expect variations of strength as different rock materials close
differently under pressure). Where the reflection is strong, the conduit is very open; where the reflection
ceases, the conduit becomes a seal, or the saturant changes to liquid.
Looking again at Figure 4, we have to ask: What other explanation could there be for those fault-plane
reflections of apparently unchanging polarity-surely they are gas conduits? And dare we hope that the left-
hand fault conduit becomes a seal in its upper reaches, where the faultplane reflection disappears? As we
build up a coherent picture in a known area, this type of thinking is no longer fanciful.
Such fault-plane reflections do not require any thickness to the fault zone-a millimeter of void will do it. The
other possibility for a fault-plane reflection arises if the fault zone is thick-perhaps tens of meters thick (
Figure 6 ).


Figure 6


Then we can visualize a negative-positive or positive-negative reflection doublet from the upper and lower
limits of this broken zone, having a time spacing sufficient to retain some amplitude as the reflection pulses
interfere. The doublet is negative-positive if the saturant is liquid, but it could be positive-negative if the fault
zone is filled with a cement harder than the faulted sediments. However, the reflection coefficients are likely
to be small in both cases.
A related situation can occur where waters in the fault conduit cause either cementation or leaching in the
faulted strata, out to some distance from the fault (wall-rock alteration). Again the reflection coefficient of
this zone is likely to be fairly small; it may also be transitional.
Summarizing this discussion, we see three situations in which we expect significant fault-plane reflections:
1. Thick layers with large throws; the fault-plane reflection is likely to be discontinuous
and to alternate in polarity.
2. Any layers with very large throws; the fault-plane reflection can be fairly continuous, in which
case it is positive for normal faults and negative for reverse faults.
3. Any layers and any throws, but with gas in open pockets in the fault plane; the fault-plane
reflection is continuous over the extent of the gas, and always negative.
Steeply Dipping Fault Planes
Fault-plane reflections are of value for accurate localization of faults. But fault planes often dip very steeply,
and may even approach the vertical in the shallow section (where we are particularly concerned to see the
transition from conduit to seal). So we must address the problem of recording steep dips.
The following are important considerations:
As always, the location of the lines is the most important variable.
Dip lines must be extended far enough away from the fault to record the fault-plane reflection;
this may be a long way.
Strike lines intended to tie the fault-plane reflection must likewise be the appropriate distance
from the fault.
To accommodate vertical faults, the group interval should not exceed V/2fm, where V is the
velocity to the zone where the fault is vertical (often shallow), and m is the maximum frequency to
be maintained through the field work and processing.
In those areas where it is known that the fault dips do not exceed some value , the group
interval may be lengthened to V/2fm sin .
The group length should equal the group interval; double-length groups are not recommended
(either in the field or by mixing or summing in processing).
The stack-array approach should be used.
The simple practical test of all the above requirements is that diffractions having their apices at
the fault should be preserved all the way out until their tails become nearly straight ( Figure 1 and
Figure 2 ,


Figure 1


The field work and processing in (a) will produce a good fault-plane reflection; those in (b) will not).


Figure 2


As always, good resolution is desirable, and this means a high value of fm But any hope of
distinguishing the polarity of a fault-plane reflection also requires good preservation of the low
frequencies.
However, the terminations of stratal reflections against a fault are best shown by the high
frequencies alone (effectively reducing the diameter of the reflection zone). Therefore it may make
sense to make one section filtered to pass only the high frequencies, and on this section to provide
a red-blue rectified display to accentuate the terminations of the stratal reflections; then we may
make a second section, including the low frequencies, to optimize the fault-plane reflection and the
reflection correlations across the fault.
The steep dips of fault-plane reflections demand appropriate processing routines; at the least,
they demand prestack partial migration (DM0) to provide adequate stacking ( Figure 3 and Figure 4
,


Figure 3


A fault-plane reflection imaged without and with DMO), and a large-dip poststack migration
program.


Figure 4


These considerations become extremely important in the recurring practical situation where a trap
is cut by a fault, the fault provides a conduit from a deep source of gas, but the fault extends to the
surface - steepening as it rises. Everything depends on whether the conduit becomes a seal above
the trap. If we find a good negative fault-plane reflection at depth, we are encouraged to expect the
gas conduit; if this reflection ceases above the trap we are encouraged to expect the seal. But what
if the reflection ceases merely because our field work or processing could not handle the steepening
dip?
We must seize on those few opportunities that afford an internal check on seismic data. One of
these is that, after migration, the fault-plane reflection must fall right on the terminations of the
stratal reflections. This is a powerful check-on aliasing, on programs and on velocities.
Because of the need for small group intervals and detailed processing, fault-plane reflections are
expensive. As always, we have to balance cost and benefit.
Thrust-Plane Reflections
At this stage we should consider a few particularities of thrust planes. Obviously, their smaller dip makes
them easier to record, without the intensive techniques appropriate to steep dips; thus, the compressional
thrust plane of Figure 5 (like its extensional counterpart the glide plane of Figure 6 )


Figure 5


produces visible fault-plane reflections with what we would today call reconnaissance techniques.


Figure 6


In Figure 6 the glide plane maintains constant polarity over a large extent; however, the gliding itself is
probably telling us that the underlying material is an overpressured shale, so that there is no message about
throw (or conduits or seals) in this maintained polarity. In Figure 5 the zone of constant polarity may be
telling us that a conduit exists, and this might be relevant to the prospectivity of the rollover structure.
Where there is rollover, of course, we do not need the thrust plane as a seal. But in situations of less intense
compression we may see a picture like Figure 7 (a), where the rollover has not yet developed and the fault
shows merely as an abrupt change of dip.


Figure 7


In this case we are happy to see the evidence of compression, for while extensional faults rely mostly on
chemical precipitation to seal them (which may or may not be there), compressional faults always have the
additional sealing mechanism of recrystallization under pressure. Then we have the chance of an additional
direct indicator of gas ( Figure 7 (b)); not only do we look for a bright negative reflection over the gas, and a
modest positive reflection from the gas/liquid contact, but we hope also for a modest positive reflection from
the fault plane within the gas.
A Hoax
In fault country, many sections (particularly from land work) show low-angle events of constant polarity
extending over significant distances, and fortuitously these may look very much like thrust-plane reflections (
Figure 8 (a)).


Figure 8


They are near-surface events reflected from faults whose expression at or near the surface makes a small
angle to the line; they may be refractions in a normal near-surface, or direct reflections if the surface
material is thick basalt. The test is to map the events on this assumption ( Figure 8 (c)), and to see if they
are coherent from line to line on the map. Of course, each event must be mapped on both sides of its line,
since the seismic evidence does not resolve this ambiguity unless either the line or the fault is not straight.
Refracted Fault-Plane Reflections
Figure 1, Figure 2


Figure 2


, and Figure 3


Figure 3


illustrate a situation where, for a coincident source and geophone, the downgoing signal is refracted into a
deep high-velocity layer, reflected back at the vertical fault, and "refracted" back to the geophone.


Figure 1


From the geometry and the polarities of the interfaces, the normal-incidence reflection must be positive, and
the phenomenon occurs only on the upthrown side; the refracted fault-plane reflection is negative, merging
smoothly and tangentially into the negative branch of the diffraction at half the critical distance from the
fault.
In practice, of course, we have to accommodate CMP geometry, which changes (a) to (b) in Figure 1 (The
generation of a refracted fault-plane reflection on a near-trace section). Instead of a triangle of reflection, we
have a parallelogram of refraction, but the path lengths above the reflector/refractor are substantially
unchanged. Therefore the refracted fault-plane reflection stacks well in its higher reaches, but at later times
(where it receives the NMO appropriate to later reflections) the stacking may become poor. The event
marked RR in Figure 4 has been interpreted as such a refracted fault-plane reflection.


Figure 4


Figure 3 is the equivalent of (a) for a basin-edge situation where the reflector terminates against a vertical
wall of very hard rock. The backward branch of the diffraction is now positive, as is the refracted fault-plane
reflection; the two are again tangential at half the critical distance. Figure 5 ( Backward diffraction, fault-
plane reflection and refracted fault-plane reflection, from a basin-edge situation) has been interpreted as an
example.


Figure 5


A refracted fault-plane reflection is of some value in interpretation. It establishes that there is a high-velocity
layer of thickness sufficient to support a significant refraction arrival, and gives its velocity. It establishes
positive polarity for the reflection from the top of this layer. It establishes that the layer is faulted by a fault
of large throw, probably near-perpendicular to the layers. If the refraction is of opposite polarity to the
reflection, it establishes that the fault is down-thrown on the side opposite to the event. And it "points"
toward the fault edge (or, strictly, towards the diffraction whose apex is at this edge).
Such refracted reflections may be optimized by a separate stack maintaining the NMO (not the stacking)
velocity appropriate to the top of the refracting layer; dip may be taken into account. A genuine refracted
fault-plane reflection should follow the structure of the refractor. Obviously, it is misplaced (and its refraction
velocity falsified) by migration.
Occasionally, fault-plane reflections and refracted fault-plane reflections are observed on the same section,
as in Figure 5. Figure 6 (a) (unmigrated) is also interpreted as an example; at first glance, we might have
expected that the upper (and stronger) near-linear event would be the fault-plane reflection.


Figure 6


Figure 6 (b) (migrated) shows that it is the lower near-linear event that is the fault-plane reflection; the
migration of this event is very good. The upper event, though distorted and misplaced by the migration, is
interpreted as a refraction along the strong stratal reflection.
Reactivation and Wrenching
On rare occasions, the earth surprises us by allowing considerable movement without the reactivation of
older faulting ( Figure 1 ) In general, however, the earth cracks along preexisting zones of weakness, so that
the reactivation of old faults is very common.


Figure 1


Sometimes we see an example that is clear-particularly when a normal fault is reactivated as a reverse fault.
Figure 2 is one such; Figure 3 is its interpretation.


Figure 2


Faults A and B, normal faults after time PP', were reduced in throw, and later sediments reverse-faulted,
after time NN'. The direction of the fault plane appears to have maintained itself much the same in the upper
sediments.


Figure 3


When the reactivation is in the same direction, the history may not be as clear. Thus a growth fault, for
which the activation is continuous rather than episodic, generally yields a smooth plane (which may steepen
upwards), and major problems for us in correlating the reflections across the fault. Perhaps the ability to
make definite correlations across a fault is our best indicator that the activation was episodic. In this case we
can search for the differences in reflection intervals across the fault that would define the times of
reactivation, and the amount of throw at each of these times.
In this we may be helped by the genesis of new faults at the time of reactivation. Thus in Figure 4 we
visualize that the reactivation of the old normal fault engendered a massive slump in the unconsolidated
sediments above, along one or perhaps two glide planes angled quite differently from the old fault.


Figure 4


Or, we may expect that antithetic faults are generated at the top of an old fault when it is reactivated, and
that when this happens the "continuation" of the old fault is steepened ( Figure 5 ).


Figure 5


In practice, of course, it may be difficult to distinguish this from the commonplace generation of antithetic
faults into a curved normal fault ( Figure 6 ); we sometimes get help (as we did in a previous example)


Figure 6


by making the mental distinction between a fault propagating upwards ( Figure 5 ) and one propagating
downwards (as in the antithetic fault of Figure 6 ).
The reactivation of a vertical normal fault may produce a "fan" of new faults, with both normal and reverse
members ( Figure 7 ).


Figure 7


When we see this we are unlikely to understand the details, but the "V" of the fan is still very useful in
pinpointing the timing of the episodes of faulting, stability and reactivation. And we always have the hope
that what is unclear on one line will become clear as we build the picture from all the lines.
We have to accept that the forces involved in successive episodes of faulting (whether compressional or
extensional) may not act in the same direction. As we saw in earlier sections, the earth may break into
lozenge-shaped blocks, each of which may be rotated-yielding local compressions and extensions, local
uplifts and local strike-slip movements. At the grand scale, we are confronted by a wrench zone ( Figure 8 ).


Figure 8


Quite apart from the inherent complexity of two or more systems of faults, we must admit that the seismic
method is not very good at detecting strike-slip movement. This sometimes surprises surface geologists, who
are accustomed to reading the simple message of offset colors and offset faults on their surface maps.
Seismically, at least in the case of layers of constant thickness, there is no direct indication that an observed
fault has a strike-slip component. However there are some guides:
First, we positively expect strike-slip faulting if the regional tectonic history is known
to include episodes of compression/extension acting in different directions.
We generally exclude major strike-slip displacement on faults that do not involve the basement. If
we can see the basement on the seismic section, and the fault extending down to it, we are braced
for strike-slip movement; if the seismic quality (or the record length) does not allow us to see so
deep, we check the gravity and magnetic data for indication of a basement fault.
We generally exclude major strike-slip displacement on faults that are not vertical or near-vertical.
Strike-slip displacement can occur at a small scale on reactivated inclined faults, but we are unlikely
to be able to detect this seismically; it is the vertical faults that put us on guard.
On land, the surface-geological maps may give clear proof of strike-slip offsets at the surface; we
must honor these at depth. At sea, just off the coast, we may be able to project strike-slip
indicators in the onshore surface geology or the coastline, and may be helped in this by the gravity
and magnetic data.
A marked difference of the character of the section across a near-vertical fault is a sure indicator
of strike-slip movement ( Figure 9 , A section from an area of wrench tectonics.


Figure 9


Before the wrench, and interesting rollover had formed into a thrust plane (which may well be a gas
conduit). This thrust plane (like the stratal reflections) has been offset by the near-vertical wrench
fault. A second thrust plane has formed, probably after the wrench). In this we are not fazed by the
difference of character that can occur across a growth fault, because growth faults ordinarily dip at
a less steep angle.
Strike-slip movement is suggested wherever zones of intense deformation occur in isolation
between undeformed areas. A zone of intense deformation may be seen as a wide vertical swath of
no reflections, or chaotic reflections ( Figure 10 , A section across a major wrench fault, illustrating
both the zone of chaos (at depth) and the ansence of correlation across the fault).


Figure 10


Because of the broken nature of the rocks in this zone, the forward branches of diffractions from
the truncated reflectors are not likely to be seen; on migrated sections, false "smiles" may be
generated. A particularly wide swath suggests a zone in which the fault is changing direction in
plan, or in which two separated parallel faults overlap. Such wide swaths can be interpreted only by
approaching them along clearer faults.
We remember that bends in strike-slip faults (in plan) require the creation of a third direction of
faulting; this may show as a distinct fault (typically bisecting the obtuse bend angle), or merely as
the wide swath of chaos discussed above.
Whenever older structure is cut by a strike-slip fault, there must be alternations in the sense of
the fault, or corresponding variations in the throw; thus, a fault may appear to be here down-to-
the-north, there down-to-the-south, then down-to-the-north again.
In practice, it may be difficult to read this simple indicator, because of complication by structure
induced just before and during the strike-slip faulting.
The above considerations make it easier for us to contemplate the interpretation of flower
structures like those of Figure 11 and Figure 12 .


Figure 11


In Figure 11 (A positive flower structure formed by transpression),


Figure 12


for example, we recognize the steep basement-related fault at depth (which we might pick more
nearly vertical than in the illustration), the evidence for sustained extensional subsidence in the
thickness of sedimentary section, and the suggestion of strike-slip motion in the incompatibility of
intervals between one side and the other; these conjure up a picture of a triangular sliver of rock
caught in a compressional pinch between the two sliding rock-masses. We marvel at the ability of
the rocks to bend; we conclude that most are shales, and possibly cap-rocks. Then we ask: Are
there sands present also, with that fault conduit from below . . . ? But does the structure have
sufficient area, in such a depth of water? The stuff of exploration.
In Figure 12 (A negative flower structure formed by transtension), we recognize similar
components, but in a situation where the strike-slip movement produces local tension, rather than
compression; the sliver of rock drops down into an opening chasm. This is of no interest as a
target, but important in telling us that the general regime of Figure 11 -while it certainly has a
wrench component-does not necessarily have a compressional component.
Wherever the nature of a fault allows the interpretation of strike-slip displacement, we search for
structural (or even stratigraphic) markers which might show the amount of displacement. Thus the
offset of a high ridge (or a graben) across the fault may give us this displacement; so may a line of
truncation at an unconformity, or even a well-marked channel.
The other approach, or course, is to identify a reflection interval that changes across the fault, and
then to search for the lateral displacement that yields equal intervals. The most positive way of
doing this is to construct an interval map. This has the important advantage of highlighting any
ambiguity in the reading of the displacement.
It is also worth scanning the structure maps for situations where a longitudinally extended basin is
deeper at one end than at the other; faults subparallel to the basin axis may have been subject to
strike-slip displacement.
Basic Interpretation Rules
From our observations of faults and migration, we can begin to see some basic rules for fault interpretation:
At an early stage, we must form some ideas on the direction of the structural grain
and the fault grain, and on how these relate to the line directions. In particular, we
wish to identify those lines that are approximately perpendicular to particular faults.
Among these, we wish to concentrate the initial interpretation on those that are also
on structural dip; these are most likely to be clear. In the early stages we avoid the
strike lines, which are likely to be very obscure.
To obtain the fault directions we must correlate the faults from line to line. Often it helps to
display the sections at the workstation, in the form of a fence diagram ( Figure 1 ). By rotating the
diagram under software control, we can quickly establish the most-likely correlations, and check
that no fault splays or other complications appear on the strike lines.


Figure 1


In correlating faults from line to line, we check that the faults are of the same type, broadly
explicable by the same stress mechanism, of the same age, and of compatible throw. In this we are
very conscious of the constraints on fault behavior discussed in the previous section; in particular,
the displacement ellipse is the control on what we mean by "compatible" throw.
Where fault correlation between two dip lines is ambiguous, we may need to check other lines
having less favorable orientations. Thus, the situation of Figure 2 (a),


Figure 2


(The fault picks on the two lines, although made at poor orientation, may permit confident
correlation-provided that the sense of the faults is certain: the fault correlation is unresolved),
although not ideal for showing the faults clearly, may allow us to work back from the line
intersection in the two directions, and be quite confident that the two faults are the same fault. It is
important, however, that the faults should be sufficiently clear to leave no doubt about their sense;
Figure 2 (b) leaves the two fault directions quite unresolved.
On unmigrated sections, a true fault-plane pick (that is, one made from apex to apex of the
diffraction hyperbolas) must tie at a line intersection. This is an important check.
As the picture becomes clear on the dip lines, the fault-plane pick can be carried to all lines,
including the obscure lines subparallel to the fault. Again these picks must tie at all the line
intersections on the unmigrated sections. So, to a reasonable degree, should the projections of the
faulted reflections. ( Figure 3 ).


Figure 3


As shown, on unmigrated sections, and in the absence of gross lateral changes of velocity, all
stratal reflections and all fault-plane reflections should tie at the line intersections. Where
reflections are truncated before the line intersection, their reasonable projections-within the same
fault block-should also tie. Further, correct fault picks should tie at the line intersections.
In general, we use both migrated and unmigrated sections. We use the migrated sections to
clarify and locate the faults where the line is near-perpendicular, and we use the unmigrated
sections to check the ties at the line intersections until we are confident of the migrations. Where
we are fully confident of a fault pick on the migrated sections, we transfer that pick to the
unmigrated sections by demigration, with the help of the diffraction apices and use this pick at the
line intersections to trace the fault on to the strike lines. The knowledge of the location of fault
edges on the migrated dip lines also tells us where to expect sideswipe diffractions (and possibly
fault-plane reflections) on the unmigrated strike lines. This is a valuable input to the interpretation
of what may otherwise be strange effects on these lines.
As this work proceeds, we decide whether it is better to make an unmigrated structure map, and
migrate this map, or to make a map direct from the migrated sections. If we decide to map the
unmigrated sections, we know we run the risk of making a wrong reflection pick across a zone
disturbed by diffractions, and of mispicking the faults. If we decide to directly map the migrated
sections, we know we shall have a mistie problem, and a need to move updip the location of all
lines having a strike component. Each survey has its own appropriate balance between migrated
and unmigrated sections; we never use only one, without reference to the other.
We keep in mind the importance of "anchoring" the map at places where there is no dip.
Traps defined by one or more updip faults (either normal or reverse) normally require that we
contour the fault planes and the top-reservoir surface independently, in order to calculate the
trapped volume. Then we can construct an isometric view ( Figure 4 ), which is particularly useful in
checking the reasonableness of the fault/reservoir boundary.


Figure 4


Faults are often not straight. Likewise, seismic lines on land are often not straight. We therefore
face the possibility of seeing the same fault several times on one line ( Figure 5 ).


Figure 5


This may be difficult to recognize, because it happens where the fault indications are blurred by the
poor angle between fault and line; nevertheless, it is very important that we do recognize it. We
should note also that the apparent age of the faulting may be quite spurious.
The age of the faulting (generally interpreted from the stratal level of the topmost visible
displacement) is important to us as a constraint on fault correlation from line to line. However, we
have to allow some tolerance on this constraint. In Figure 6 we see a fault plane interpreted from
three seismic lines, and its expression (hatched) where it offsets the same reflection (A, at
equivalent stratal level).


Figure 6


If now we visualize a contour of the displacement ellipse (such as that corresponding to the
smallest visible displacement, as illustrated), we can see that the age pick on the three lines (p, q,
r) does not correspond to a constant stratal level.
The behavior of a fault at depth is also useful for correlation, since it contains information on the
type of faulting. In one type, the faults (which may be planar) persist to great depth, offsetting the
basement ( Figure 7 (a) ); these faults extend downward until the rocks become plastic.


Figure 7


In another type, the fault planes are curved concave-upward, and the faults sole out to bedding
planes (or perhaps unconformities) at depth ( Figure 7 (b) ); subsequent adjustment faults may die
out at, or just above, these planes ( Figure 7 (c) ). In another type, faults in the shallower section
may be absorbed in salt (or other plastic or mobile material), without continuity into the deep
section; in the case of salt, these faults may be caused entirely by movement of the salt, but this
movement may itself have been triggered by deeper faulting. In the latter case, although individual
faults do not persist through the salt, it remains true that a zone of faulting above the salt broadly
follows a zone of faulting below the salt ( Figure 7 (d) ).
Significant strike-slip motion usually involves the basement, and is therefore not expected unless
the fault is like Figure 7 (a). Further, strike-slip faults are usually near-vertical. These are useful
distinctions.
A recurring problem, in the picking of faults, concerns the degree to which we can expect fault
faces to be irregular; that is, neither planes nor smooth curves. Often, the joining of reflection
terminations on seismic sections seems to need very irregular fault faces ( Figure 8 ), and we are
unsure whether to accept these or to insert additional faults. The following observations assist the
decision:

1.


Figure 8


Growth faults are generally smooth. (This is Earth's compensation to us for
the agonies of trying to correlate across such faults.) We remember that,
under conditions of sustained rapid deposition in an extensional basin, the
fault faces are curved concave-basinward in plan and concave-upwards in
section.
2. On a seismic time section, the usual increase of velocity with depth makes inclined
planar faults appear curved, and listric faults appear more curved. Indeed, an inclined
fault that appears planar must be slightly convex-up. These effects are not removed by
time migration. So, for each survey we interpret, we should take a moment to compute
the degree of apparent curvature introduced into planar faults, over a range of fault dips.
3. The division into basement-related and other faults, discussed above, is fundamental.
An inclined postdepositional basement-related fault propagates upwards, finding the
easiest path; initially the factors deciding this easiest path are rupture strength and
overburden density, and these may lead to abrupt changes of fault angle within the
section. As movement occurs, these changes of direction may be accommodated by
plastic deformation or by pulverization, according to the type of rock; if not, new faults
are to be expected, branching off upwards from the change of direction. On the other
hand, faults involved in crestal collapse over an uplift propagate downward, approximately
along a radius of curvature; there is little or no accommodation problem, and the faults
are likely to be planar.
4. Thrusts tend to follow the bedding in layers that allow sliding (for example, in some
shales), and then to step upward quite abruptly in the brittle layers. Seismically, only the
"risers" of this stair-step picture may be visible. Identification of the "treads" as shales is
feasible, if the migration is trustworthy.
5. Where a fault has become stable, and then sediments have been deposited over it,
reactivation of the fault may involve a change of fault direction at the top of the original
fault. There may also be an antithetic fault branching off at this point. However, we must
be aware of the hoax situation that may arise at an unconformity ( Figure 9 ).


Figure 9


6. The appearance of abrupt changes of fault direction may be caused, erroneously, when
faults are themselves faulted ( Figure 10 ).


Figure 10


7. In view of these and other complications, we try where possible to pick fault faces as
smoothly planar or smoothly curved, but we are ready to introduce abrupt changes in
fault direction where the seismic data quality leaves no doubt, or where we see clear
evidence for fault reactivation, or where thrust planes traverse shales, or where we
recognize successive episodes of tectonic stress in different directions, or where faults
pass close to abrupt changes of rock type.

As the tectonic history emerges, and we recognize the direction and degree of the
tectonic forces acting in particular episodes, we can use analog models to guide our
expectations and our picking. And, at all times, we remain very conscious of the
constraints on fault behavior-particularly the need to conserve volume.
Clearly, fault interpretation can be very labor-intensive, and we cannot afford to study every fault in
detail. But we do need to take care with those that represent potential traps. In extensional areas,
unfortunately, such traps are often in the "shadow" of a fault ( Figure 11 ), and we must be aware
of the problems that this poses:

1.


Figure 11


The fault zone itself is likely to be full of scatterers, decreasing the transmitted
signal and adding noise and imperfectly collapsed diffractions.
2. The different velocities along normal-incidence paths a and b distort the reflection
times.
3. As shown by the CMP paths c, d and e, the stacking is likely to be impaired by local
changes in effective velocity.
4. We particularly need a strike line along the top of the trap, but this is likely to mistie to
the dip lines.
5. Depth conversion imposes discontinuities on reflections in the shadow of the fault,
leading to doubts about the fluid-content reflection.
6. Strange paths, such as f and g, become possible.
If this list appears daunting, we can take some cheer from the fact that we are usually luckier
than we deserve. Sometimes we are amazed at the virtual absence of shadow effects, as in
Figure 12 .


Figure 12


Often we see only minor effects. Only occasionally do we see major disruptions in the shadow
zone, as in Figure 13 .


Figure 13


Incidentally, this figure is of interest as a case where the picking of the main fault is facilitated
by noting the numerous gas accumulations (evident both as bright spots and fluid-contact
reflections), and using them to infer where the fault must be. The reason for the pronounced
fault shadow is almost certainly the heterogeneity introduced by the gas.

Fault Plane Maps
A fault plane map is a contour map in which the strike, dip and throw of the fault plane are presented in map
view. Remember that fault throw and vertical separation are different; and we must be careful as to which
one we are measuring and mapping. Also, keep in mind that fault planes not truly mathematical a plane
because they change in strike and dip directions; we sometimes use the term fault surface map to describe
the same information as on a fault plane map.
A fault plane map is constructed by first measuring the throw (or vertical separation) and depth (or reflection
time) of a fault and indicating these values at the proper location on a base map. We generally use either
well log and/or seismic data as the source for this information. The common procedure is to post the data
values using the notation "throw or vertical separation/depth or time," such as 500/3.200 when the input is
seismic reflection time. This example means that the throw is 500 milliseconds and the depth is 3.200
seconds. Of course, we can also indicate (and contour) measurements in feet or meters. However, to map in
feet or meters using seismic data, we must have good knowledge of the velocity distribution for an accurate
time-to-depth conversion; we know that the presence of faults in the data introduces complications in the
horizontal and vertical velocity distribution. We also are aware that fault shadows create even further velocity
complications.
Contouring Fault Plane Maps
Once the data are posted, they are usually contoured according to depth (or time) only. Sometimes, fault
plane maps have both depth and throw contoured on the same map as we shall see later.
Contouring fault plane maps requires a different set of guidelines than structural contouring. Some of the
more general rules for contouring fault plane maps are (according to Tearpock and Bischke, 1991):
We begin contouring where the data density is greatest. At this point we do not
know the fault plane pattern although we can generally tell something from working
the data, from regional knowledge, and from glancing at the data values. This starting
procedure is contrary to structural mapping where we generally begin at the structural
highs and work our way to the structural lows, regardless of date density.
Contours on a fault plane map may be open-ended. Again, this is contrary to structural maps
where contours must close (except along the map margins due to the limitations in the size of the
map).
Contours are generally smooth except for ramp-and-flat thrust faults and other unusual
circumstances where data clearly indicate that changes in strike and dip may be abrupt.
More than one fault plane may be contoured on a single map. Contours, therefore, may split,
cross, or merge depending on whether the faults are bifurcating, compensating or intersecting,
respectively.
Figure 1


Figure 1


is an example of a fault plane map with two sets of contours-depth and vertical separation (
Figure 2 , Diagrams of seismic lines A and B).


Figure 2


In Figure 1 , the fault plane depths are the solid contours while the vertical separations are
indicated by the dashed contours. Notice that the vertical separation contours are, for the
most part, perpendicular to the fault plane depth contours; this implies that the fault is pot-
depositional. This is due to the absence of growth (fault throws are equal for the shallow and
deep horizons) on the seismic lines.
Bifurcating Faults
A fault, generally a normal fault, can split laterally along strike and continue as two smaller faults dipping in
the same direction. This is called bifurcation and results in a bifurcating fault pattern. Figure 3 is a block
diagram showing the configuration of a bifurcating fault system. We notice in the figure that all the faults are
dipping in the same direction.


Figure 3


When the original fault splits into two bifurcating faults, the vertical separation must be conserved. That is,
the sum of the vertical separation for the new, smaller faults must equal the total vertical separation of the
larger, original fault In our model, shown in Figure 3, we see that the vertical separation of one bifurcating
fault (100 feet), added to the vertical separation of the other bifurcating fault (150 feet), equals the vertical
separation of the original fault (250 feet).
A fault plane map ( Figure 4 ) of a bifurcating system is a plan view of the block diagram shown in Figure 3 .


Figure 4


In the map we see the three points noted in our definition and model:
1. We see that the contours for Fault L intersect the same (value) contours for Fault
K. That is, the 12,000 foot L contour intersects the 12,000 foot K contour at the line of
bifurcation.
2. The vertical separation of Fault K (120 feet), added to the vertical separation of Fault L (65-
105 feet), equals the vertical separation of the original Fault L + K (153-225 feet).
3. All the faults are dipping in the same directionin this case, to the southwest.
Compensating Faults
Two faults, generally normal faults, dipping in opposite directions can merge along strike and continue as one
fault. This is called a compensating fault system. Figure 5 is a block diagram showing the configuration of a
compensating fault system. We notice that the two faults, as stated in the definition, are dipping in opposite
directions.


Figure 5


When two compensating faults merge, the vertical separation in the continuing fault must be conserved, just
as in a bifurcating fault system. In this case, that means that one of the compensating faults must "take
displacement away from," or compensate, the other fault, such that the continuing fault equals the
mathematical difference between the two compensating faults. This compensation occurs because of the
opposing dip directions of the two faults-faults having the same dip direction add throws along their
termination line; faults having opposing dip directions subtract throws along their termination line. In our
model ( Figure 5 ), therefore, we see that the vertical separation of one compensating fault (250 feet), less
the vertical separation of the other compensating fault (100 feet), equals the vertical separation of the
continuing fault (150 feet). Of course, geologically, one fault can not take displacement away from another
fault; this is just a simple means to comprehend the mechanics of a compensating fault system.
A fault plane map of a compensating fault system based on well data is shown in Figure 6.



Figure 6.


Here we see that Fault B, which dips to the northwest, and Fault A, which dips in the opposite direction (to
the south), merge and continue as Fault A (still dipping to the south). Again, we note that:
1. Equal contours of the compensating faults intersect each other along the line of
termination (of Fault B at A).
2. The vertical separation of Fault A (100 feet) at depth below the line of termination, equals the
difference between the vertical separation of Fault A (300 feet) above the line of termination, less
the vertical separation of Fault B (200 feet).
Cross section AA' is shown in Figure 7 to aid in the understanding of this example.


Figure 7


Intersecting Faults
Unlike the previous two special cases (bifurcating and compensating fault systems), where one fault
terminates while the other, in a new form, continues beyond the intersection, intersecting faults both
continue after their intersection regardless of their dip direction. This fault pattern may involve both normal
or reverse faults. Figure 8 shows a block diagram of an intersecting fault system.


Figure 8


The vertical separation for two intersecting faults may or may not change laterally or vertically beyond the
line of intersection. If the two faults occur contemporaneously, then the vertical separation of each fault
remains the same before and after their intersection. However, if one of the faults occurs after the other fault
(in time), then there will be some compensation of vertical separation across the line of intersection. In
either case, conservation of vertical separation must exist. Unless the data strongly support a change in
vertical separation across the line of intersection, it is generally assumed that the vertical separation is the
same on either side of the intersection.
Figure 9 shows a fault plane map for two intersecting faults.


Figure 9


We see that Fault A, noted by solid contours above the line of intersection (to the north) and by dashed
contours below it (to the south), dips to the south, while Fault B, noted by solid contours below the line of
intersection (to the south) and by dashed contours above it (to the north), dips to the northwest. Again we
see that equal contours intersect each other at the line of intersection.
The cross section AA' for the intersecting fault system of Figure 9 is shown in Figure 10.



Figure 10.


Here we see that a horst below the intersection, becomes a graben above the intersection.
Uses of Fault Maps
There are several reasons why we should routinely make fault plane maps as part of the exploration and
exploitation processes:
Obviously, fault plane maps provide us with the location of a fault at all depths in
plan view. This gives us much more detail than a structural map, which tells us only
where the fault is located relative to just one mapped horizon.
Fault plane maps can provide us with strike and dip information at any location along the fault
plane.
Fault plane maps give us a better understanding of the three-dimensional structural picture and
may even identify new prospects.
Fault plane maps can be very helpful in the design of drilling plans, especially for directional and
deviated wells.
Fault plane maps can be used in conjunction with other structural maps and data to establish the
limits of a hydrocarbon reservoir.
Fault plane maps can be used in secondary and enhanced oil recovery projects to postulate the
fluid flow pattern.
Structure Maps: Basic Principles
Historically, the search for oil and gas has been the search for structure. This is the first message of seismic
reflection data. For many years, our task as geophysicists was to find and map structural situations, leaving
to the geologist the task of determining what the rocks were likely to be and whether they were likely to
contain hydrocarbons.
Although our mandate today is much broader, it remains true that our primary function is to delineate
structure. Structure is what seismic data does best, with most assurance.
Our eye is therefore highly tuned to the recognition of structure on seismic time sections, and to the
evidence of whether the structure is genuine. Three questions are fundamental:
1. Is the structure closed in time? in depth?
2. Is the amount of closure in time so great that it would survive any conceivable errors in the
static corrections, or any conceivable lateral variation in the velocity used for conversion from time
to depth?
3. Is the structure, as evident on the sections and particularly its upper part, displayed in the
correct position horizontally?
We, as experts, can look at seismic sections and maps, and quickly assess the type of
structure apparent from them. We can also look at the annotation of the sections, such as
elevations, corrections, velocities and applied processing, and quickly assess the certainties
and the doubts about the structure i.e., the measure of the seismic risk, and where that risk
lies.
In acquiring this skill, it is helpful to have a compendium of many examples. Accordingly, the AAPG has
published an atlas of seismic sections in three volumes illustrating the many types of structural style. These
atlases, by A. W. Bally (1983), are highly recommended. We enjoy returning to them periodically, and seeing
things we missed before. Of course, many of us also keep our own collections of illustrative sections that
have come across our desks and have captured our attention for one reason or another.
Figure 1 illustrates the Keystone field in West Texas-a classical petroleum trap. As we can see, the field is a
simple closed anticlinal dome, with two culminations. The anticlinal axis is shown by a bold line, with arrows
to highlight the dips.


Figure 1


We notice that the contours appear natural almost everywhere on this map.
The delightful simplicity of the Keystone anticline in Figure 1 occurs only with gentle structures. When the
folding is more intense, faulting inevitably occurs. This may take many forms, depending on the totality of
the stresses involved. Most anticlinal traps, therefore, tend to look more like Figure 2 -basically a dome, but
cut and displaced by many faults.


Figure 2


Indeed, in most exploration mapping it is not possible to separate the operation of contouring from the
decisions of fault interpretation.
Tying Faults
Tying faults means that we correlate the same fault, or faults, from one seismic section to another. We do
this for several reasons:
1. It helps us to identify a fault on one seismic section as being the same (or different,
if we are incorrect) fault on another section.
2. It helps us to establish, understand and correctly map the fault system within the study area,
particularly in distinguishing the major from the minor faults.
3. It helps us to assure that the correlation of horizons and their relative positions across the faults
are correct.
4. It helps us to monitor the throw of the faults from one seismic section to another and to assure
that we are consistent in our interpretation.
Figure 3 illustrates how we tie faults across several seismic sections.


Figure 3


In the figure, we immediately notice that faults can apparently change dip directions
depending on the orientation of the seismic line relative to the strike of the fault. In tying faults,
however, we recall that faults on unmigrated sections tie; faults on migrated sections,
generally, do not. We use migrated seismic sections to identify and locate a fault, and
unmigrated sections to interpret it.
Structure Mapping: Example
To illustrate how to structurally map faulted seismic data, let us "pick" the following data set.
Figure 1 is the shot point map.


Figure 1


Figure 2 ,


Figure 2


Figure 3 ,


Figure 3


Figure 4 ,


Figure 4


Figure 5 ,


Figure 5


Figure 6 ,


Figure 6


Figure 7 , and Figure 8 are the seismic sections, Lines 2-6,11 and 12.


Figure 7





Figure 8


On Figure 9 Line 4, we recognize and mark the fault.


Figure 9


This example is made rather easy for us by three circumstances:
1. We have a train of reflections (3.17-3.30 s at the east end of Figure 9 ) generally
stronger than the rest, and with a recognizable character. This makes it easy for us to
correlate across the fault, and so to assess the throw of the fault (about 120 ms,
down to the east). In the following material we shall assume for the sake of illustration
that our potential reservoir is near the top of this train.
2. As we shall see, the fault cuts the line almost perpendicularly. This always makes the seismic
reflections more clearly discontinuous and the faults easier to see. In contrast, a line along the
strike of a fault can be very difficult to see and interpret.
3. Earlier in this module we reviewed the concept of migration. We remember the importance of
this process, which moves dipping reflections an appropriate distance updip, in locating faults
correctly. In other modules, we learn that migration also sharpens and clarifies the indication of
faults, provided that the faults are approximately perpendicular to the line. In the present case,
then, the fault on Figure 9 is particularly clear. This section is a migrated section as are all sections
in the figures in this example.
Now that we have recognized the large normal fault as our target, we take our usual broad
look at sections. In Figure 10 and Figure 11 , we see a montage of the relevant parts of the
east/west sections, aligned approximately on the fault.


Figure 10


As we go north (upwards), the fault develops, increases to a maximum, and then attenuates
to nothing again.


Figure 11


We now mark the position of the fault on the lines of the map as shown in Figure 12.



Figure 12.


The temptation is immediately to join the marks together, and to indicate one major fault. However, before
things become too fixed in our minds, we must consider the alternative interpretations: is it possible that we
are wrong in assuming we have only one fault? So we check on the nearest north/south line (Line 12, Figure
8 ) to see if it has any indication of faults. It has not. We are lucky; we can proceed, at least tentatively, on
the basis of the single fault.


Figure 8


So we commence the picking. As always, we prefer to start where the picking is clear. In general, also, we
prefer to start on a dip line, because we learn the structural style more quickly, and so can accommodate our
picking to it. Where there is faulting, however, we have another consideration: if there is any doubt about
the visual correlation of reflections across the fault, and in general there is, we must try to tie our picks
around the fault. We pick to the fault on one side, then round on tie lines to the other side, and so confirm or
correct the visual correlation across the fault.
In the present case ( Figure 12 ), we would like to pick from the west to the fault on Lines 3, 4, and 5, tying
our picks to the north/south Line 11, then tie Line 11 to the unfaulted east/west Line 6, and so be able to
approach the fault from the east. But unfortunately Line 6 does not tie to Line 11. So we have to use Line 2,
to the south. This does not have the big down-to-the-east fault which we are trying to map, but it does have
a small down-to-the-west fault. Fortunately, there can be very little doubt about the correlation across this
fault. So we decide to start by picking the west end of Line 3, then south on Line 11, east on Line 2, north on
Line 12, and so to the east end of Line 3.
Now we have to decide which part of the reflection train to pick. As we look at the lines to the west of the
fault (for example, Figure 3, Line 3, near SP 600), we feel reasonably confident that at least two major
reflections are present in this train. Let us say that we will pick the upper one. Where is the maximum of its
envelope? It is difficult to be sure, but the first major trough between two major black peaks seems a
reasonable first guess. Before we proceed, however, we check all the lines to see whether this picking
decision will be tenable over the whole area. We conclude that no picking decision is without its problems. In
some places we shall need to smooth, or to estimate, or even to ignore small complications which are not
relevant to our task of mapping the fault. But we will probably stay with the decision to pick that first major
trough.
In Figure 13, we pick Line 4 west of the fault.


Figure 13


For present purposes, we ignore the complication between SP 1570 and 1590, being quite confident of our
pick on the two sides of it.
At the junction with Line 11 we check the tie ( Figure 14 ). This is acceptable, and we turn south. Thereafter
we turn east on Line 2, north on Line 12, and west again on Line 4.


Figure 14


This completes the interpretation of the data. Now we must contour it to produce our structure map.
The contouring of the western part of the area is straightforward; mostly, it represents simple regional dip.
But the contouring of a fault requires some thought. Generally, one should follow these guidelines when
contouring faulted data:
Often it helps to regard the fault merely as a steep dip. Then contours that enter the
fault turn and follow the fault, finally emerging on one side or other from the fault.
Contours that enter a fault must also leave it.
The throw of the fault follows from the contours on the two sides of it. The variation of this throw
along the fault must be geologically plausible.
Contours beyond the end of a fault should show the flex likely to be associated with an incipient
fault.
The spacing of the contours on the two sides of a fault should take into account the likely shape of
the surface before it was faulted.
Figure 15 shows one version of the contoured structure map. It shows, superimposed on the
regional dip to the southwest, a south-plunging anticlinal nose whose east flank is cut by the
down-to-the-east fault. It also shows a much smaller down to-the-west fault, which intersects
the down-to-the-east fault in a compensating fault system. This assumption may be typical of
the judgments which we must make many times a day. There is no escape from making these
judgments. However, each of us evolves some personal scheme for marking our working map
to show the confidence of these judgments. For example, the continuity of the major down-to-
the-east fault across Lines 3, 4, and 5 is highly probable; it is the only interpretation to be
seriously considered on the present data. The continuity of the minor down-to-the-west fault
across Lines 2 and 3 is possible and plausible, but is certainly not the only interpretation. If it
were critical to a prospect, we would need to try several alternatives and consider their risks
to the viability of the prospect.
It is an important duty of ours to convey to management not only the appeal of the prospect, but also its
uncertainty the interpretational risk.
We notice the offset between the steep south gradients on the two sides of the fault, and how the implied
variations in the throw of the fault are expressed in the contours. If we wished, we could draw a side-view of
the fault plane between the two surfaces (for example, a view from the east).
The final contour map, of course, is a disappointment to us. We had hoped for a drillable prospect-a closed
trap against the fault. Certainly a first look at the sections allowed this hope. But now we find that there is no
closure. Both upthrown and downthrown sides of the fault rise steadily to the north or northeast and there is
no trap.
It should be mentioned at this time, and will be discussed in a later section in more detail, that the
development of 3-D seismic data has led to a new perspective on mapping faults. In the past, faults were
thought to be long continuous features that gently meandered through the subsurface. This is essentially the
same concept as that applied to the previous exercise.
With the advent of 3-D seismic data and workstations, this perception is quickly being replaced. The mass of
3-D data has given us the ability to see much more detail than ever before. These new data show that most
faults extend only a short distance and that many more faults exist than had been previously thought.
Instead of a few long faults, there are numerous short ones. Similarly, 3-D data have changed the way we
contour faulted structure maps. As noted above, when using 2-D data, contours on both sides of a fault
appeared to flow sinuously. 3-D data, however, have shown that contours on the two sides of a fault are not
as sinuous or flowing as once believed.
Examples of the 3-D seismic expression of faults and associated maps will be shown later.
Role of Reflection Time
The message of reflection time includes much more than the location of traps and
faults: first, there is the critical importance of reflection time in unraveling the
geological history of an area; second, the change in a reflection time interval may
also have a message about the type of rock to be expected; and third, the
message of reflection time often contains information on the maximum depth of
burial of a layer. We will focus on the first of these three relationships.
The example we will use is that of Figure 1, from which we will reconstruct the
history as shown in Figure 2,



Figure 1





Figure 2


Figure 3 ,


Figure 3


Figure 4 ,


Figure 4


Figure 5 ,


Figure 5


Figure 6 , and Figure 7 .


Figure 6


The key is to draw conclusions from the change of a reflection time interval.



Figure 7


As indicated in Figure 2, Figure 3 , Figure 4 , Figure 5 , Figure 6 , and Figure 7 ,
there are six major stages of movement and deposition of the seismic section
shown in Figure 1. In the first stage (f), the parallelism of reflections suggests
stable subsidence and matched deposition from Base Zechstein (Ze) salt through
Lower Jurassic (JL). In the second stage (e), truncation of the Lower Jurassic by
erosional unconformity suggests uplift, caused by salt thickening in the left half of
the section, in Upper Jurassic or Lower Cretaceous time. At this time, some minor
faulting occurs below the salt on the left side of the section due to the uplift. Any
sediments deposited during these times and before this movement were removed
by erosion. Parallelism of the erosion surface and lowest Upper Cretaceous (KU)
reflections suggests salt movement had ceased by early Upper Cretaceous time.
The third stage (d) is characterized by subsidence and deposition from Upper
Cretaceous through Lower Triassic (TRL) time while salt thickening occurred in
the right half of the section in the early Upper Cretaceous. During this third stage,
minor faulting occurs below the salt and in the Upper Triassic-Lower Jurassic
rocks in the right side of the section due to the uplift.
Pronounced salt thickening occurred in two episodes during and after Lower
Tertiary (TL) in the fourth stage (c). At this time, the faults on the right side of
the section become more numerous. In the fifth stage (b), there is continued
uplift, increase in the intensity and growth of faulting and erosion of some of the
'Tertiary sediments. The faulting above the salt now extends through the whole
Mesozoic section. Finally, in the last stage (a), salt movement, uplift and faulting
ceases. Sedimentation proceeds uneventfully to present day.
The conclusions gained from this exercise regarding the structural and faulting
context of the prospects are:
1. Minor uplift and faulting begins in the Lower Cretaceous time.
2. Major uplift and faulting occurs in the Tertiary.
3. Faulting of the right-hand structure would probably lead us to prefer to
drill the left-hand prospect first.
4. Faulting of the right-hand prospect could block the migration of
hydrocarbons to the apex of the structure and trap them downdip where
fault closure exists.
Contrary to the saying that "timing is everything," in the search for hydrocarbons,
timing is only a piece of the puzzle. But it is an important piece that cannot be
overlooked. Without knowledge and understanding of the timing of any faulting
within a prospect, we run a greater risk of drilling a dry well.
Exercise
1/ Draw a bifurcating fault system in map view where the throw of the original
fault is 190 ms before bifurcation and 150 ms afterwards. What is the throw, in
ms, of the newly split fault?
Solution



Figure 1


The correct answer is 40 ms ( Figure 1 ).
2/ Draw a compensating fault system in map view where the throw of the two
original faults are 200 ms and 70 ms. What is the throw, in ms, of the continuing
fault?
Solution



Figure 1


The correct answer is 130 ms ( Figure 1 ).
3/ Draw an intersecting fault system in map view where the throw of the two
faults is 150 ms and 75 ms. What are the throws, in ms, of the faults after
intersecting, assuming that the faults occur contemporaneously?
Solution



Figure 1


The fault throws remain unchanged-150 ms and 75 ms ( Figure 1 ).
4/


Figure 1


Using the fault plane map in Figure 1, if a well is drilled exactly half-way between
existing wells number 11 and 14, at what depths would you expect to encounter
Faults A and B?
Solution
Fault B occurs at -6,860 ft.
Fault A occurs at -7,200 ft.
5/


Figure 1


In Figure 1, list the chronological order of faults, from oldest to youngest, with
Fault A being on the far left and Fault E on the far right. What type faults are
they? Has the area undergone a change in tectonic regime and, if so, what was
this change?
Solution

Oldest Fault A normal
Fault E normal
Faults C and D normal
Youngest Fault B thrust
Yes, the basin has undergone a change in tectonic regime. Prior to Fault B, the area was under an
extensional stress regime. At the time Fault B occurred, the area changed to a compressional stress regime.

Introduction
The concept of structural style is based on comparative tectonics. Its greatest utility lies in identifying certain
basic patterns of deformation and faulting that are repeated in geologic provinces throughout the world.
Studies of structural style have the considerable advantage of relating these patterns to present-day plate
tectonic habitats and thereby to predictive models of origin and evolution. This means that prior to
exploration, the types of source and reservoir rocks, migration paths and hydrocarbon traps can be
predicted, or at least anticipated.
In classical structural geology, surface mapping, well data and laboratory experiments were the main sources
of information for understanding tectonic features and their genesis. This has changed dramatically within
the past several decades. Seismic reflection profiling has provided us with a tool that can reveal continuous
structural data to considerable depths. It has opened up both submarine provinces and the deep continental
subsurface to eager scrutiny.
As developed by Harding and Lowell (1979), Bally et al. (1983), Bally and Oldow (1983), and Lowell (1985),
structural styles are first differentiated on the basis of whether basement is involved or uninvolved directly in
deformation. In the case of noninvolvement, structures primarily develop within a "detached" sedimentary
cover.
These two basic criteria are, in turn, applied to the four major types of tectonic provinces-compressional,
extensional, strike-slip (or wrench) and cratonic (or vertical). For the most part, these last two types of
province are thought to involve basement in nearly every case.
First of all, we need to define what is meant by basement. For petroleum geology, this is usually taken to be
structural basement, which means rigid crystalline igneous or metamorphic rock. The degree and manner of
its involvement in the creation of structures are important to understand, since these determine the overall
context for structural entrapment in overlying sediments. In fact in many hydrocarbon provinces, basement
structure is the key to both overlying deformational and depositional patterns.
These provinces and the structural styles within them are related genetically to one or more of the following
plate boundary classifications:
1. divergent, which includes intracontinental rifts, proto-oceanic (new ocean) basins
and oceanic basins
2. convergent, which includes intra-oceanic arc (oceanic/oceanic) systems, continental margin arc
(oceanic/continent) systems, closing ocean (continent/continent) basins and both forearc and
backarc zones
3. strike-slip, which includes any of the above basinal systems where the dominant deformational
forces are oblique to each other rather than in the same plane and the resulting movement is
predominantly horizontal rather than vertical
Likewise, just as basins and structural styles are a product of a tectonic process, faulting
within a basin is generally a product of that same tectonic process. Thus, the dominant type of
faulting within a particular basin is the result of the dominant deformation force. The
classification of structural styles-whether referring to the overall basin, the individual
structures or the faults-is in fact the same.
Divergent Basins: Overview
Divergent or rift basins are closely linked with plate divergence and seafloor spreading. These basins
generally form as down-faulted failed arms, where continental plates are stretched but do not completely
separate. They are usually fairly small, linear basins that can be either deep or shallow ( Figure 1, A typical
rift basin. Generalized cross section through the Suez Basin of Egypt). Horst-and-graben tectonics generally
give them an irregular profile.


Figure 1


Not all divergence results in failed rifts. Some rifts will evolve into pull-apart basins, as the seafloor continues
to spread and basaltic ocean crust is added ( Figure 2, A typical pull-apart basin.


Figure 2


Generalized cross section through the Gabon Basin of west Africa). These basins form next to the trailing
edges of continents that have been split apart. For this reason, they are also sometimes known as passive
margin basins.
Pull-apart basins are generally long, linear, asymmetric basins with a one-sided continental source. They
occur mostly offshore on intermediate sub-basin crust. Most of the world's pull-apart basins are located on
both sides of the Atlantic and Indian Ocean margins.
Di vergent Basins: Basement-Invol ved Faul ting
Regional block faulting is perhaps the most widespread structural style in the earth's crust, characterizing a
majority of the world's passive continental margins ( Figure 1 ),


Figure 1


as well as numerous, linear intracratonal grabens that occur on nearly every continent ( Figure 2 ).


Figure 2


In most cases, block faulting is directly related to the process of rifting (i.e., continental breakup). In a few
possible exceptions, e.g., the Basin and Range province of western North America, crustal extension still
demonstrates the principal features of rift tectonics and thus may, in fact, represent unique or anomalous
circumstances of this same basic process. It should be noted, however that block faulting also occurs along
some convergent plate margins and in backarc settings.
Tectonism in these provinces may be thought of as dominantly tensional in nature. At the same time, rifting
is a complex process; rarely, in fact, during its initial stages does separation occur exclusively at right angles
to the central rift axis. Components of wrenching and compression are common; the former may be related
to transform offset of ridge segments. individual faults, or fault assemblages, may therefore show
considerable strike-slip displacement. Local folding and reverse faulting may also exist.
Perhaps the most important characteristic regarding normal faulting is its deceptive simplicity. Though
normal faults most often appear relatively simple in cross section, they are most often extremely complex in
map view. Individual faults can be straight or cuspate, or can alternate between the two. Individual blocks
bounded by faults can vary substantially in size and can be tilted or rotated in different directions to different
angles within a few square miles. In places, faults may appear to die out in ways that seem contrary to the
predictions of geometry or rock mechanics. In a broad sense, much of this can be thought of as resulting
from the intricate readjustment that occurs due to the creation of "extra space" by extension. In many cases,
the precise pattern of structure related to a particular prospect or play may not become clear until years of
detailed work have been done.
Figure 3 shows three basic models that have been proposed to explain the geometry of rift-related normal
faultingstraight, listric and domino. Mach controversy continues to surround this subject.


Figure 3


The degree to which basement-involved faults show listric geometry appears to vary between regions. Such
variation is believed to result from differences in how rifting evolved and in the type and thickness of crust
involved.
Due to the excellent exposure and relative youth of its features, the Red Sea graben is often used to point
out the principal aspects of this structural style. In a study of the evolution of the Red Sea ( Figure 4 ),
Lowell et al.


Figure 4


(1975) recognized that the larger, regional faults vary in strike and frequently intersect, but demonstrate a
dominant orientation that parallels the central graben. Two important secondary fault trends appear to form
a conjugate set that is bisected by the primary trend. Displacement along them, however, is almost entirely
dip-slip. Both trends appear to have developed simultaneously during various stages of rifting. Together,
they represent the principal style of rift-related faulting. Active extension, and thus faulting, appears to have
been episodic. As a result, considerable overprinting of older structures by newer ones has occurred. The
cumulative effect is great complexity on a local scale, despite the overall consistency of major structural
trends.
Most faults in the Red Sea graben are interpreted to show listric geometry at depth in cross section ( Figure 4
) and a curving trace in plan view. Seismic profiles, however, have been interpreted to show more or less
straight, relatively high-angle normal faults, at least in several areas. Since reflection profiles generally
exaggerate the curved geometry of faults at depth due to velocity increase, the linearity of these fault planes
appears all the more distinct. The question of detailed fault geometry, therefore, remains partly unanswered
in this region.
In basement-involved divergent basins, we often see fault-bounded blocks rotated in stair-step fashion
toward the central graben axis. Seldom, however, is the structure this simple. Antithetic faults, those that dip
opposite to the major faults (i.e., away from the rift axis), are also common. These, too, can be thought of
as conjugates. Fault wedges and basement blocks can be rotated in any number of directions, and both the
geometry and magnitude of displacement often vary considerably, even between adjacent blocks. Sufficient
rotation can cause reverse drag folding in hanging-wall sediments, such that beds "roll over" and thus imitate
the structure of many growth faults.
In addition, a phenomenon known as footwall uplift, where the footwall is physically displaced upwards as a
result of isostatic and elastic rebound, has been identified as accounting for about 10 percent of the total
offset along basement block faults (Jackson and McKenzie, 1983). Such uplift can lead to local thickening of
sediments over the hanging-wall block, thinning over the footwall block, and to local bathymetric highs that
can localize reef growth or other shallow marine sedimentation.
A look at the Piper Oil Field in the central North Sea graben ( Figure 5 ) reveals several other major features
consistently seen in basement-involved divergent basins.


Figure 5


Figure 6 ,


Figure 6


Figure 7 ,


Figure 7


and Figure 8 are taken from Maher (1980) and show the relationship of production to block fault structure in
the Piper Field.


Figure 8


Note the location of the oil/water contact. Given that migration and entrapment most likely occurred after
faulting and that erosion tilted and sealed the Piper Formation, the faults within the sand body are probably
nonsealing.
Cross section ( Figure 7 ) and seismic data of the Piper Oil Field ( Figure 8 ) reveal how strata are draped
over basement blocks. These blocks exercised a dominant control over depositional environments. In
particular sedimentation of the high-energy, marginal-marine Piper Sand was largely determined by the
location of structurally positive horsts. Moreover, later movement along faults created migration paths into
the sand body. Thus, rifting was instrumental in creating conditions conducive to both the deposition and the
favorable structural geometry of source and reservoir rocks.
Basement block faulting also occurs at plate convergence and in association with transform faults. This
structural style characterizes backarc basins and sometimes the inner margin of forearc basins. To date, the
former have been found to be more petroleum productive. With regard to transforms that cut through
continental crust, such as the San Andreas, strike-slip motion sometimes generates extension and normal
faulting on the cratonal side of related basins. In the southeastern Great Valley of California, such faults were
generated during the episode of most rapid movement along the San Andreas, and have acted to trap
significant amounts of petroleum.
Passive continental margins represent one side of a successful rift separation. The term passive refers to the
general assumption that most orogenic tectonism has ceased along such a margin. They are characterized by
seaward-pro-grading wedges of mostly shallow marine sediments that can reach 14 km in thickness (Bally
and Oldow, 1983). In their overall character these deposits reflect the basic stages of rifting and continental
separation.
Basically, then, the evolution of passive margins can be divided into an early, rift-related phase and a later
drift-related phase. Earlier structures are predominantly associated with basement block faulting. Later phase
deformation is concentrated in the overlying sedimentary wedge and is controlled mostly by gravity (e.g.,
growth faulting, salt and shale structures).
A more detailed look at these two basic structural levels is offered in Figure 9 and Figure 10 , which shows
the current configuration of the passive boundary along northwest Africa.


Figure 9


We note, in this case, how three sedimentary sequences are evident.


Figure 10


In this figure, we see that the early phase lithotectonic sequence is confined to the Triassic-Lower Jurassic
units, while the bulk of the later, continental margin phase is represented by Cretaceous and Tertiary
sediments and that the structures developed within them. With regard to structure, where the Triassic-
Jurassic sequence is characterized by basement faulting, the Cretaceous sequence shows large-scale
detachment faulting, and the Tertiary sequence is, so to speak, faultless. We might notice how the
detachment faulting appears to be broadly related to deeper basement block faults. These, however have not
been reactivated such that they cut up into the Cretaceous sequence.
When we compare the seismic section of offshore western Africa shown in Figure 9 and Figure 10 to that of
the Baltimore Canyon area (offshore U.S.) shown in Figure 11, we immediately see the similarities in the
faulting style.


Figure 11


What we are seeing are mirror images of two sides of a seafloor-spreading center. As in western Africa, the
Baltimore Canyon area is dominated by deep, basement block faulting with younger detached gravity-slide
deformation.
Numerous types of hydrocarbon traps occur in basement-involved fault blocks in rift provinces and passive
margins ( Figure 12 ).


Figure 12


Deep burial may, however, modify some of these, as will salt mobilization. The more important role of this
tectonic setting, vis-a-vis petroleum potential, has been to act as the site for the deposition of great clastic
wedges, especially deltaic, that are characterized by relatively shallow detachment faulting, as well as salt,
shale and gravity sliding structures that have trapped enormous amounts of hydrocarbons.
Di vergent Basins: Growth Faults
By far the greatest number of hydrocarbon traps of the non-basement-involved type have been found along
passive margins in association with growth faulting, so named because of its syndepositional nature. They
are best developed in regions of deltaic or continuous clastic shelf deposition, where sediment accumulates
rapidly and remains semi-consolidated to depths of several kilometers. High pore pressures in shales due to
retarded dewatering is a frequently cited influence on the generation of growth fault structures. Regional
crustal subsidence due to sediment loading provides at least a broad context for extension. However, a great
many growth faults are also directly associated with salt diapirs, which essentially act as large, locally
upthrown blocks. Historically, the Gulf of Mexico has served as the type locality for this structural style.
Figure 1 shows an offshore growth fault, as interpreted on a seismic profile.


Figure 1


Some of its more important characteristic features are:
1. The major fault plane is listric in shape, dips into the basin, and becomes parallel
(flat) to bedding with depth.
2. Displacement along the plane increases with depth.
3. The thickness of individual units increases abruptly in the hanging wall, due to syndepositional
offset.
4. Both secondary synthetic and antithetic faults complicate the basic structure. Many of the former
sole out along the main detachment plane.
5. Back rotation of the downthrown block creates rollover that increases, then decreases and shifts
basinward with depth. Some of the secondary faulting appears to be related to extension in the
crest of this structure.
6. In map view, the fault plane is cuspate, and is usually concave toward the basin center.
Antithetic faults, however often show the reverse geometry.
The flattening of faults with depth has been variously interpreted. Explorationists have
traditionally pointed to the existence of overpressured, ductile shale intervals into which the
faults often seem to die out. Roux (1977), however has attributed this flattening to increasing
compaction-faults originally develop at steep angles that flatten with burial as bed thickness
decreases. Of these two interpretations, the former continues to be preferred.
Shale units in which the effective in situ stress is especially low can be assumed to be preferential zones for
faulting. Pore pressure effects ensure very low shear resistance. This preference also appears to influence
fault geometry. As shown in Figure 2, faults flatten out into shale-dominant prodelta slope facies.


Figure 2


The pattern of such faulting can be directly or indirectly related to preexisting basement structures, as we
have mentioned. The rejuvenation of deep-seated faults by sediment loading can directly lead to structural
adjustment in the overlying cover. Preexisting basement structure can more indirectly establish the context
for growth faulting by influencing the development of depositional hinge areas. This, in turn, can directly
affect the location and trend of detachment structures.
An example of indirect basement influence is given in Figure 3, which shows a complex example of large-
scale, growth-type faults developing above the deepest portion of a rifted margin.


Figure 3


In addition, deep-seated faults can be rejuvenated by sediment loading and this may cause structural
adjustment in the overlying cover. Here, older, fault-generated basement topography has determined
thickness and facies patterns of later sedimentation and, thus, the location of highly ductile limy shale
intervals. Flowage of this lithology under the influence of gravity-induced stress has played a significant part
in the development of growth faulting ( Figure 10 ).


Figure 10


Some researchers consider gravity to be the principal cause of growth faulting in areas such as the Gulf of
Mexico or Niger deltas. With respect to gravity, the maximum principal stress is vertical. The shear
component of this stress, as resolved along the regional depositional slope, leads to instability. In general,
body forces within a volume of sediments deposited over a sloping surface must be considered significant.
Where this volume is very large, as in the case of passive margin clastic wedges, these forces are very likely
to be fully capable of generating large-scale fault systems.
At the same time, a variety of other tectonic influences, related to salt and shale diapirism, possible
basement fault rejuvenation, and so forth, exist in most passive margin settings. Since growth faulting
seems to be a regional response to overburden and slope-related instability or metastability, it may result
from a number of specific causes.
Figure 4


Figure 4


and Figure 5 of the Hibernia anticline show the fallacy of trying to strictly segregate structural styles in every
situation.


Figure 5


The Hibernia discovery was made approximately 250 miles due east of Newfoundland in a narrow, Late
Mesozoic graben in the Jeanne d'Arc Basin that cuts into the northern portion of the Grand Banks platform.
The graben developed as part of the active rifting phase that separated Africa and North America between
the Late Triassic and Cretaceous ages. The structural style, at least in Lower Cretaceous beds ( Figure 5 ), is
growth faulting with large-scale rollover. Compare Figure 1 to Figure 4. The faults, in Figure 4 case,
developed as part of the rifting episode and though syndepositional to some extent, are not all detachment
structures. The more major faults extend deep into the Paleozoic section and most likely continue into
basement (Arthur et al., 1983). Hibernia, therefore, presents a case in which the geometry of features
assigned to a non-basement-involved structural style of growth faulting also appears in the features of a
basement-involved rift style of block faulting.
Figure 6, Figure 7


Figure 7


, Figure 8


Figure 8


,


Figure 6


and Figure 9 shows some of the more common hydrocarbon traps that result from growth faulting.


Figure 9



Di vergent Basins: Salt Movement Faults
Extension above salt structures causes anticlinal warping and normal faulting associated with crestal collapse
that usually occurs with shallow penetration. The latter is sometimes termed keystone, for reasons that
should be apparent in Figure 1.



Figure 1.


With depth, however faults associated with salt structures are more influenced by the upward movement
rather than by downward crestal collapse. As such, normal faulting still dominates, but most often appears
radial in its pattern. As with normal faulting in other structural styles, local complexity around the salt
features can be great. Antithetic faults are very common.
The basic scheme of hydrocarbon traps associated with salt structures is shown in Figure 2.



Figure 2.


In many cases, the largest traps are in reservoirs not pierced by the salt, as in, for example, numbers 1, 2, 8
and 9 in the figure. This, however, is not a hard-and-fast rule.
Di vergent Basins: Shale Movement Faults
The conditions that encourage the flowage of shale within a thick wedge of clastic deposits are primarily
related to retarded dewatering of shale and to the overall sequence of sediments in a regressive
(progradational) shelf.
During rapid burial, sands dewater and compact faster than do shales. Impermeability acts to inhibit water
loss. An exception to this occurs at the contact between a shale interval and an overlying porous sand, but
here the expulsion of pore water into the sand creates an upper seal even more impermeable to continued
water loss from the underlying shale. Trapped, excess pore water keeps the clay mineral grains apart, thus
countering lithostatic pressure, and keeps the rock in a weak, ductile state, susceptible to flow. This is the
classic explanation for the overpressured shales so commonly (and, often, unfortunately) encountered along
passive margins buried by thick, regressive clastic sediments.
Along with Jackson and Galloway (1984), we might think of the entire depositional body as a "spoon" (
Figure 1 ), composed of a proximal, sand-dominated "megafacies" wedge that overlies a more distal, shale-
dominated one.


Figure 1


This setting leads to what is known as differential compaction. As the sands compact and become more
dense, they subside into the less compacted shales.
Experimental and theoretical work have both shown that the stress resulting from differential compaction and
loading is greatest at the front of the load and that the more mobile, overpressured material beneath will
flow laterally to allow continued load subsidence. In geologic terms, this means that, due to the depositional
slope, shales will be forced basinward. This can lead to growth faulting on a local scale and the creation of
shale mounds. Such mounds can subsequently act to localize sand deposition, as progradation continues,
and, thus, the generation of other growth faults ( Figure 2 ,


Figure 2


Figure 3 ,


Figure 3


Figure 4 ,


Figure 4


Figure 5 ,


Figure 5


Figure 6 , Figure 7


Figure 7


, and Figure 8 ).


Figure 6


They may also rise as diapirs into overlying units or they may spread laterally.


Figure 8


Differential compaction apparently leads to the transfer of great masses of shale, sometimes ahead of a
series of growth faults. Bruce (1973) has described shale ridges in south Texas with dimensions of over 20
km in length, 40 km in width, and 3,000 m in height, with a trend that is roughly parallel to regional strike.
Shale diapirs also seem to be related to gravity sliding along master growth faults, which become planes of
detachment at depth.
Shale structures are not usually of great significance to petroleum exploration. The Beaufort Sea represents
an important exception to this observation. Shale structures are, however, intimately associated with growth
faulting, salt diapirism, and the overall structural evolution of thick clastic wedges along passive continental
margins. This makes knowledge of them necessary to both regional and local exploration in such settings.
Di vergent Basins: Gravity-Induced Faulting
Gravity-induced downslope movement has been invoked by structural geologists to explain an almost
dizzying variety of phenomena on almost all geologic scales. It has been used, for example, as an
interpretation of the rumpled-cover nappe structures of the Alps, the de'collement thrusting of the Canadian
Cordillera, the chaotic sedimentary sequences known as melange or olistostromes, the disturbed layering of
single depositional layers, and the generation of growth faulting.
Gravity is now invoked to explain specific small scale structures. In many instances, pore pressure effects are
thought to play a primary role in initiating actual detachment. Tilting due to tectonic uplift will create slopes
along which the force of gravity will be resolved as shear stress. This has been applied to specific
decollement structures (see the classic explanation of the Heart Mountain Fault by Pierce, 1957), and also to
both local and regional detachment features along Gulf Coast-type passive margins characterized by high
rates of sedimentation. Among the latter are rotational slides observed high on upper delta front slopes and
larger-scale, slower-moving rotational blocks that show growth fault-type structures at their head. A third
common setting for apparent gravity-related displacement is in thrust belts, where topography can become
oversteepened due to the continued advance of thrust sheets. Here, local normal faults will relieve the
resulting stresses, which are almost entirely due to the simple weight of the material involved (i.e., the
maximum principal stress is gravity and therefore, vertical).
In relation to normal deposition along passive, divergent margins, so-called thin-skinned gravity sliding is
seen to occur in close association with growth faulting and shale diapirism (Jackson and Galloway, 1984).
'Two proposed versions of the structural evolution which might be responsible for this close interrelationship
are shown in Figure 1 ,


Figure 1


Figure 2 ,


Figure 2


Figure 3 ,


Figure 3


Figure 4, Figure 5


Figure 5


, and Figure 6 .


Figure 4


These have been derived from detailed study of the Mexican Ridges area in the western Gulf of Mexico.


Figure 6


The difference between the two models lies in the steepness of the shelf slope. Zone 4a, which has a slightly
steeper slope, is characterized by gravity sliding, while Zone 4b, which is characterized by growth faulting,
has a slightly flatter slope.
In the view of most researchers, each glide sheet in this type of environment has a growth fault at its head,
a decollement along its principal length, and a thrust or faulted fold at its terminus. To some degree, these
sheets can be likened to giant slumps, with growth "s carps" marking the upper planes of rotation. We should
notice the ponding of sediments between and over the ridges. The loading from this accumulation is thought
to effectively stop gravity gliding and help initiate shale diapirism.
Figure 7 shows an excellent example in seismic section of a productive rollover trap (Slick Ranch Field)
associated with such a glide sheet.


Figure 7


The irregularity of the glide plane, which is partly due to shale diapirism, acts to localize the degree of bed
rotation. Movement along the basal detachment is assumed to be by continuous gravity creep. In the section
we note that the top of the Vicksburg Formation becomes conformable in the direction of gliding.
As with shale structures, these gravity-induced features are important to petroleum exploration mainly for
their relation to growth faulting. The anticlines caused by the type of downslope movement shown in Figure 1
, Figure 2 , Figure 3 , Figure 4 , Figure 5 , and Figure 6 . should be considered potential hydrocarbon traps in
deeper water areas off passive margins. Melanges, on the other hand, are important to onshore exploration,
since they directly indicate the facing direction of a fossil arc, and thus the general orientation of structural
trends to be expected.
Convergent Basins: Overview
There are four basin types that form when plates converge near continental margins. These are the forearc,
backarc, strike-slip (or non-arc) and collision basins. All four convergent margin basin types have a number
of features in common and will be discussed to some extent as a group.
There are two types of basins that are found near subduction zones that have developed island-arcs. Backarc
basins form between an island-arc and a continent. Forearc basins lie between the island-arc and an ocean
trench.
Non-arc basins are formed along convergent margins where the plates move by transcurrent faulting.
Consequently, they are often called strike-slip basins. Strike-slip basins are small basins that form through a
combination of both the transcurrent fault movements and local block faulting. Strike-slip basins include the
Los Angles and Ventura Basins of California, the Vienna Basin, and the Crimea and Baku Basins of the Soviet
Union.
Collision basins, sometimes called median, intermontane or successor basins, are small basins formed within
marginal fold-belts, along sutures where either two continents, or continental coastal mountains and a
trench, have collided. Some explorationists consider this a type of backarc basin, but a well-defined trench-
arc system is usually no longer present.
Convergent Basins: Basement-Involved Faulting
The two principal elements of basement-related structural styles are compressional fault blocks and their
bounding basement thrusts, which may range in inclination from near-vertical to less than 30 degrees. Both
high- and low-angle faults occur along a single thrust front and are now thought to represent the end
members of a continuous spectrum of basement-involved deformation.
Reverse faults involving basement generally occur along convergent plate margins, primarily in foreland
regions, and in the inner trench slope and outer high portions of certain subduction zones. Of these, the
foreland regions are by far the most important to petroleum exploration.
Basement-involved forelands primarily develop in two major settings between a volcanic arc and a craton
(i.e., in a "backarc" setting), and in front of the arc during continental collision. The first of these is
particularly characteristic of the North and South American cordillera, while the latter typifies the great
Alpine/Himalayan orogenic zone.
Compressional basement block faulting is particularly well known and well studied in the Rocky Mountain
foreland region of the western United States. In this area, large, rigid masses of Precambrian crystalline rock
have been forced up and, to some degree, laterally outward many thousands of feet. The larger uplifted
blocks appear to be intimately associated with deep foreland basins. In general, each block has its related
basins, whose timing of subsidence seems directly related to uplift history. Most often, these basins are
strongly asymmetric, with their structural axes running close to the thrust front and their back flanks forming
relatively gentle basement slopes away from the uplift.
These basins have served as the sites for considerable sediment petroleum generation and entrapment.
Structural relief between the present crest of basement uplifts and their associated basins is frequently on
the order of 10 km or more. Thus, the vertical component to faulting is undoubtedly considerable. The
overlying sedimentary cover on these blocks has usually acted in a passive manner, accommodating
basement uplift by drape folding and brittle fracture. Blocks are typically bounded by faults on both sides and
are tilted at various angles.
Figure 1, Figure 2


Figure 2


, Figure 3,



Figure 1





Figure 3


and Figure 4 show four styles of faulting that have been proposed to explain the geometry observed in the
field and in seismic profiles.


Figure 4


To some degree, the differences between these reflect the fact that various investigators have concentrated
their efforts in different parts of this region, along different major fault zones. Such zones display variable
geometries, which range from reverse listric planes (steepening with depth) to low-dipping thrusts of 30
degrees or less ( Figure 6 ).


Figure 6


Single faults apparently show both these geometries along their strike. At present, the proposed styles of
faulting reflect two major schools of thought. Both schools agree that compression is involved to some
degree in deformation, but part company on the question of whether displacement is dominantly vertical
(upthrusting of Figure 1 and Figure 3 ) or horizontal (over-thrusting of Figure 3 and Figure 4 ). The seismic
profiles given in Figure 5 and Figure 6 illustrate how apparently irreconcilable these two schools are.


Figure 5


Many of the principal faults have also apparently suffered a degree of ancillary strike-slip displacement. For
these reasons, we now believe that at least several of the major blocks, such as the Wind Rivet have been
uplifted in a type of scissor-like rotational pattern, with the inclination of the fault plane decreasing and the
amount of lateral movement increasing with distance from a "nodal" zone of almost total vertical uplift. This
has, in turn, been related to the clockwise rotation of the entire Colorado Plateau during Laramide time
(Gries, 1983).
With respect to those blocks bounded by near-vertical faults, Harding and Lowell (1979) describe three basic
structural levels, from basement up into the sedimentary cover:
1. a tilted fault block of basement and immediately overlying units
2. an intermediate level where the principal fault becomes a zone of steep drag folding
3. an uppermost level of gentle drape of monoclinal folding
The structure of fault segments can be complex, involving normal faulting along the crest of
the block, drape fold, tear faulting of the cover and basement, and more minor reverse
faulting in the footwall of the main fault.
Some of our colleagues have pointed to older zones of Precambrian shearing within the basement itself as a
probable control on location and possibly geometry of faults. Certainly, in any region where basement
faulting is a dominant structure style, we should understand (or attempt to understand) the preexisting
structural character of the rocks involved. Metamorphic rocks are, in general, the products of orogenesis and
very often contain important planes of weakness, such as shear zones and major lithologic boundaries, that
will have some amount of influence on all later deformation.
Another possibility involves the inversion of rift-related block faults during later compressional tectonism.
Convergence of older passive plate margins can lead to the rejuvenation of originally normal faults or
thrusts. Such reactivation, since it reverses the sense of displacement, is said to create inverted basins.
Understanding this can be crucial in certain regions, since the sediments normally assumed to characterize
rift provinces will become involved in anticlinal, thrust, and compressional-wrench tectonism. In such cases,
apparent horst-and-graben morphology may be retained; however, the more recent uplift will have
substantially altered the original structural configuration.
Figure 7 indicates some simple hydrocarbon traps associated with basement-involved thrusting.


Figure 7



Convergent Basins: Non-Basement-Involved Faulting
The primary structural style in this category is the decollement foreland thrust-fold belt. Due to the thorough
work of such authors as Price and Mountjoy (1970), Dahlstrom (1970), and Bally, Gordy and Steward
(1966), the Canadian Cordillera has come to be generally treated as a type locality for such deformation.
Figure 1 is the classic cross section by Price and Mountjoy (1970) through the southern portion of this
structural belt and shows most of the principal features we have come to expect in this style.


Figure 1


The majority of hydrocarbon production exists in the foothills region, where faulting is especially complex.
Thrust-faulted anticlines appear to be the most common traps in productive foreland regions throughout the
world. But with regard to the size and specific geometry of traps, it should be emphasized that considerable
variety exists within and between these regions. Over the span of Proterozoic time, differences in the size
and morphology of plates, in their marginal sedimentary character, inherited structural features, and specific
motions have all ensured a high degree of structural variation in foreland orogenic zones.
Like compressive basement faulting, decollement thrusting is most often related to processes occurring along
convergent plate boundaries, both in the mentioned foreland belts and along the leading edges of subduction
zones. Figure 2 displays the interpreted structure on a seismic section across the active subduction complex
offshore of northern California.


Figure 2


This section shows the de'collement faulting in the accretionary prism quite well.
We note the change from the chaotically deformed accretionary prism sediments into the more conformable
forearc basin at the far right in the figure. The deformed complex is broken into thrust slices about 400-500
m thick, whose bounding faults display variable curvature. This is probably due to a degree of stratigraphic
control and ramping of fault development. We can also see that where compressional features dominate the
accretionary prism, extensional faulting characterizes the deep sea sedimentary section west of the trench.
This has been observed in a number of active subduction zones. Often, these faults have developed in the
underlying ocean crust and have been interpreted as being the result of lithospheric bending and consequent
tensional rupture.
Collision-related thrust belts ( Figure 3 ) verge toward the subducting ("oncoming") plate.


Figure 3


A transition to basement-involved thrusting often occurs in these settings behind the thrust front. Figure 4
shows how the involvement is usually interpreted to increase into the mobilized, metamorphic "core" of an
orogen, such as the Canadian Cordillera.


Figure 4


In the case of trench deformation, the situation is less well-understood. 'Two forms of basement-involved
faulting are presumed-that which incorporates "slivers" of ocean crust, known as ophiolites into the
accretionary prism, and that which cuts continental crust. For both trench and foreland thrust zones, overlap
occurs between detached and "connected" structural styles.
Most thrust belts presently exposed at the earth's surface occur as sinuous belts up to thousands of
kilometers long. Their width is not uniform, but is instead characterized by sharply recessed and more gently
extended portions known, respectively, as reentrants and salients. The cause for this type of variation along
tectonic strike is not well-understood, but often thought to be related to preexisting basement features.
In discussing foreland thrust belts, there are a number of general aspects that are of direct importance to
hydrocarbon exploration:
1. Anticlines occur primarily in the hanging wall, and are asymmetric toward the
direction of tectonic transport.
2. The geometry of thrusting is much dependent on the competence of the sedimentary units
involved. The number of thrusts, as well as the intensity of folding, decreases in distinct proportion
to increases in the amount of thick, competent units (e.g., sandstone, carbonates).
3. Sedimentary thicknesses decrease and deposits change character to more shallow-water clastics
in the direction of tectonic transport, i.e., away from the metamorphic core or, in stratigraphic
terms, toward the basin margin. This usually results in an increase in the overall density of faulting
in the same direction.
4. Structures become progressively younger in the direction of tectonic transport.
5. Rocks involved in thrusting also become younger in this direction.
6. Advancing thrust sheets act to load and depress the crust into local foreland basins (we know
these as molasse basins). These fill with coarse marine and nonmarine detritus, which then also
becomes involved in deformation. Such basins are often rich in plant-derived organic matter.
7. In many cases, a regional foreland basin, relatively rich in petroleum, will exist immediately in
front of the thrust belt, presumably created by crustal loading on a more massive scale.
Strike-Slip Basins
Strike-slip, or non-arc, basins are formed along convergent margins where the plates move by transcurrent
faulting. Strike-slip basins are small basins that form through a combination of both the transcurrent fault
movements and local block faulting. Strike-slip basins include the California basins, the Vienna Basin, and the
Crimea and Baku Basins of the Soviet Union, and many others.
Wrench Faults
For most exploration purposes, strike-slip, oblique-slip, and wrench fault systems can be considered as
roughly equivalent, though their specific plate tectonic settings are variable (see Appendix A). In all cases,
they are assumed to involve basement. These fault systems can be pure strike-slip or they may include
components of compression and extension, sometimes called transpression and transtension, respectively.
Major wrench faults are usually associated with transform plate margins. In our general discussion of
faulting, we have looked at several examples of these structures (we refer to Figure 1, Figure 2 , and Figure
3 ), which can occur as single shear planes or systems of parallel faults (Harding, 1974, 1983 and 1985).


Figure 1





Figure 2





Figure 3


For petroleum exploration purposes, we are most interested in the structures associated with wrench faults,
since they form potential structural traps.
To a large extent, the components of compression and extension are determined by the degree of
obliqueness involved in plate convergence. As convergence swings from "head-on" to lower and lower angles
of incidence, the relative amount of strike-slip motion increases. Theoretically, associated compression
should decrease.
Wrench fault structures are extremely varied, as mentioned by Harding (1985), and include many of the
features seen in other styles. Referring once again to Figure 3, we see how wrench fault structures can be
resolved according to the complex stress fields that result from such large-scale shearing. The development
of specific structures and their relative importance is naturally dependent on how much compression or
extension might be involved in plate convergence.
Flower Structures
The term flower structure is often used to describe the upward-branching form that many wrench faults have
been interpreted to show on seismic profiles. The flower itself is denoted positive or negative on the basis of
whether the units within it are arched by reverse components of displacement or dropped by normal
separations, respectively. In Figure 4 we see an interpreted example of a negative flower.


Figure 4


Note that seismic data is relatively chaotic immediately east of the interpreted fault. Letters "A" and "I" refer
to the blocks whose relative motion is away from or toward the viewer.
Despite the relatively wide use of flower structure terminology, questions remain as to its value. Often, the
quality of seismic data in the crucial core areas of supposed flowers is poor, which leads to inconclusive and
alternate interpretations. Some structures that have been called positive flowers (e.g., in the Ardmore Basin
of southern Oklahoma; see Harding and Lowell, 1979) have also been interpreted as thrusts. Those with
apparent negative "bloom" have also been analyzed as rift structures. In regions such as the North Sea or
the Red Sea, the problem is most likely compounded by the fact that components of both strike-slip and
normal displacement are responsible for the fault geometries observed. Figure 6, Figure 7 , Figure 8 , Figure
9, Figure 10, and Figure 11 gives a range of structures from various styles that can appear similar to each
other and to wrench fault structures in cases where data are poor and fault offset cannot be determined.


Figure 6





Figure 7





Figure 8





Figure 9





Figure 10





Figure 11


These examples highlight the potential problems that can arise from identifying any example of highly
complex faulting as wrench-related.
Cratonic Basins
There are two major types of basins that are located in the interior of plates on the stable continental crust
or craton.
The first cratonic basin type forms only in the central areas of continents, so these basins have often been
called interior or intracratonic basins. We also sometimes call them sag basins because they form as simple
sags or depressions on the underlying basement but have no adjacent highlands. In plan view, they are
circular to elliptical in shape. In cross section, they appear saucer-like-symmetrical, shallow and somewhat
flattened.
The origin of interior basins is still poorly understood. Many of them are known to have block-faulted
basement rock and attenuated crust, so the processes involved are mainly extensional. Possibly these basins
may form in areas of aborted rifting, where dense material was injected into the lower crust, reducing crustal
buoyancy (Dickinson, 1976).
Some of the largest sedimentary basins in the world belong to this class. Examples include the Williston,
Illinois, and Michigan Basins of the central United States, and the Paris Basin of France.
A second cratonic basin type, called foreland or craton margin basins, develops more toward the margins of
continents. In a foreland basin, the first depositional cycle is usually interrupted by uplift. Then a second
depositional cycle begins, most commonly from a different direction. A regional unconformity separates these
two cycles. Although foreland basins are usually initiated by extension, the dominant processes in the second
cycle are compressional, related to the uplift and orogeny.
Because they evolve in two stages, foreland basins are sometimes called composite basins. They are usually
elliptical or elongate, vary widely in size, and have asymmetric profiles.
North America oil-producing foreland basins include the Alberta, Anadarko, Permian and Appalachian Basins,
and most of the small Rocky Mountain basins.
Vertical Faults
Hydrocarbons are often trapped by secondary faults and closures within cratonic basins ( Figure 1 ).


Figure 1


The patterns of such smaller-scale structures are completely singular to each basin, but usually display
pronounced trends that appear to be related to large-scale tectonic patterns in the underlying basement that
have been identified from gravity and magnetic data. Generally speaking, then, basement structure is often
an essential key to exploration of the potential in overlying sediments. Drilling, however, rarely penetrates
the entire depth of sedimentary fill, and thus, direct study of basement rocks is usually limited to the
shallower margins of a basin.
Faults are most often high-angle and show either normal or strike-slip displacement. Compressional
structures are usually open folds that may, as in the case of the Michigan Basin ( Figure 2 ), be consistent in
orientation.


Figure 2



Exercise
1/


Figure 1


The seismic section shown in Figure 1 shows the crestal collapse of a salt feature.
Interpret the faults as indicated on this line.
Solution
See the interpreted section. ( Figure 1


Figure 1


)
2/


Figure 1


In the seismic section shown in Figure 1, we can see a reflection off the glide
plane of the fault. Why is the glide plane reflection in this section a continuous
event, rather than a discontinuous series of positive, zero and negative reflection
coefficients?
Solution
The glide plane is along a shale structure that has uniform velocity and density.
Thus, the reflection coefficients across the fault planes vary little, and the
reflection appears to be continuous.
3/





Figure 6


What thrust styles, as defined in Figure 1 ,


Figure 7[a What thrust styles as defined in Figure 1


Figure 2 , Figure 3 , and Figure 4 , are exhibited in Figure 5 , Figure 6 , and
Figure 7 ?
Solution



Figure 1


Figure 1 -- Style c
Figure 2 -- Style a



Figure 2


Figure 3


Figure 3


-- Style b
4/


Figure 1


The synthetic seismic section shown in Figure 1 was generated from a geological
model. Sketch what you think the input model looks like.
Solution



Figure 1


Figure 1 .
5/


Figure 1


What can you say about the timing of the thrust faults present in the seismic
section shown in Figure 1 ?
Solution
The thrusting occurred very recently, probably after the generation and migration
of hydrocarbons. ( Figure 1


Figure 1


)
6/


Figure 1


Is the flower structure depicted in the seismic section shown in Figure 1 positive
or negative?
Solution
Overall, the flower is a negative structure, as defined by the outer boundary
faults, but it also has a positive component, as indicated by the inner fault.
Figure 1


Figure 1


7/


Figure 1


The seismic section shown in Figure 1 is from an offshore cratonic basin (the Sao
Luiz Basin, offshore Brazil). Interpret the faults as indicated in this section.
Solution








Figure 1
Overview
Workstations have given us new ways to manipulate and display seismic data. These new visualizations, in
turn, have given us the ability to incorporate more information and detail into our interpretations, leading to
higher success rates and more production.
For example, we can now join multiple seismic lines at their intersections to form one composite display. This
allows multiple wells to be tied together on one seismic display, even if no one, direct line exists between
them. An example of a composite display tying two wells (even though no data were acquired directly
between them) is shown in Figure 1.



Figure 1.


We notice in this example how the same fault can be observed on two of the lines.
Another example of a workstation graphic is a concertina display. This display is a series of panels on the
screen, which may show multiple parallel seismic lines, or it may be the same line but which different
attributes such as frequency, phase, amplitude, etc. Sometimes this type of display is useful in picking very
small faults.
The input to the two examples addressed above is 2-D data. It is not a requirement to have 3-D data to
conduct an interpretation on a workstation. However, 3-D data is superior to 2-D, particularly in defining
faults and structure.
A 3-D data survey has two substantial advantages over 2-D data. First, the vast amount of data in a 3-D
survey allows us to see much greater detail than was possible using 2-D data. This detail has led to better
definition of structure, direct hydrocarbon indicators and reservoir characteristics. And as we have already
seen, this greater detail has altered our beliefs on the distribution and behavior of faults, as well as structural
contours.
The other advantage is the ability of the workstation to display the 3-D data in many new formats that are
impossible to do with 2-D data. These new displays allow us to view the data in many new sections and
dimensions. Data can now be displayed as a horizontal section, as a three-dimensional cube, in a
combination of vertical and horizontal perspectives, as a random line extracted from the data that does not
follow the orthogonal in line or cross line directions, or as sections of data that follow fault planes. And all of
these graphics can be displayed from any azimuth. The result of these displays is maximum use of the data
and better interpretations.
Of all the new visualizations, the horizontal section is by far the most intriguing and informative. In this
display, the data set is transsected horizontally along a defined time or depth and usually indicated by
variable color density. The horizontal section is also called a slice if it along an isotime, it is called a time
slice; if it is along a particular depth, it is called a depth slice. Figure 2 ,


Figure 2


Figure 3, and Figure 4 show time slices from the Gulf of Mexico, the Gulf of Thailand and onshore Europe
respectively.


Figure 3


Can you see the faults in these displays?


Figure 4


With the capability to create a time slice, several other displays are now possible. A chair or folded section
display is a composite of the conventional vertical section and a horizontal time slice. It is obvious in Figure 5
why this type of graphic is often called a "chair" display.


Figure 5


We notice how the fault is presented in each of the sections and how it "ties" at the corners.
Figure 6 shows another display created with a time slice.


Figure 6


This is called a cube display. Again we notice how the fault is presented in this three-dimensional display and
how we can follow it in three directions.
In all the example displays shown above, we see how the faults can be easily followed around corners, from
line to line, and in a horizontal perspective. Faults can be more accurately interpreted and mapped using
such displays.
Another display that aids in the identification of faults is the dip map. A dip map is a projection of an
interpreted horizon such that the dip magnitude (throw) and/or dip direction (angle) is calculated and
displayed.
Figure 7 is an example of a dip map.


Figure 7


In this map, colors are used to indicate dip magnitude. Any linear contact that shows an abrupt change in
color, such as we see here, indicates the presence of faults. The more pronounced the color change, the
larger the dip.
Figure 8 and Figure 9 are two more examples of dip maps.


Figure 8


In this case, both maps are of the same area, the Nun River Field in Nigeria.


Figure 9


Figure 8, indicates the dip magnitude, while Figure 9 indicates dip azimuth. Again the faults are identifiable
by their linear appearance and abrupt change in color. We notice the ease of identifying faults in these
displays and the wealth of information that can be obtained. In particular, we notice how the faults appear to
be discontinuous along the same strike line.
Time Slices
Interpreting faults in horizontal time slices is very similar to that in traditional vertical sections. As in seismic
sections, faults in time slices are discernible by a termination and/or displacement of aligned events.
Let us look at the time slice in Figure 1 .


Figure 1


As an exercise, try picking the faults on this time slice. The event terminations indicating the presence of
faults are clearly visible in this particular display. Of course, using vertical sections in conjunction with
horizontal sections makes fault identification and interpretation easier. Figure 2 shows one interpretation of
Figure 1 .


Figure 2


Viewing consecutive time slices in rapid succession on a workstation can aid in the identification of faults.
This type of viewing is called a movie.
In a time slice movie, horizon reflections appear to move across the screen, while faults generally remain
stationary or move only slightly. This visual effect is due to the vertical nature of fault planes relative to the
seismic reflections. The higher the angle of the fault, the more stationary the fault will appear; low-angle
faults will move slowly across the screen. Another characteristic of a time slice movie is the ability to see
horizons move across the screen and then truncate against the faults. Using this technique, we are able to
recognize and locate smaller faults and be more precise with their fault boundaries.
Figure 3 ,


Figure 3


Figure 4 ,


Figure 4


Figure 5 , and Figure 6 show four frames of a time slice movie.


Figure 5


Notice the movement of the horizon reflections versus the stationary appearance of the faults.


Figure 6


Sometimes interpreting faults in time slices is not easy. This is especially true when the trend, or strike, of
the structural features and contours is the same as the strike of the fault. When this occurs, the event
terminations are difficult to see, and therefore, the faults are likewise difficult to identify. The more parallel
the fault strike is to the structural strike, the more difficult it is to see the event terminations. Conversely,
the more the faults are in the dip direction, the easier it is to see the event terminations.
Let us look at the time slice from the Gulf of Thailand in Figure 7 .


Figure 7


Some of the major faults in this time slice are easy to recognize while others are nearly impossible. This is
because the structural and fault strike are generally parallel. Again, as an exercise, try picking the faults on
this time slice.
Let us compare Figure 7 to Figure 8 .


Figure 8


Figure 8 indicates a fault interpretation of the time slice data in Figure 7 . The interpretation is based on both
horizontal and vertical sections. In this example, if the interpretation was based on the time slice alone,
many of the faults might have been overlooked.
Time slices can also be used to generate various fault plane projections. Displaying faults interpreted on
sequential time slices in a 3-D format provides a simple means to do so. Figure 9 shows such an example.


Figure 9


Here, the fault planes are displayed in a cubic projection. Essentially, this type of display is a three-
dimensional view of fault plane maps, but better, because it gives us the ability to validate a fault
interpretation in three dimensions.
Fault Slices
One of the more innovative displays of 3-D data and the workstation is the fault slice. A fault slice ( Figure 1
) is a projection of data along a surface parallel to an interpreted fault plane. It is displayed as a vertical
section.


Figure 1


We use a fault slice for determining fault throw, sealing and leaking properties, juxtaposition of lithologies,
planning of deviated wells, identifying splinter faults, determining other fault related criteria, or simply
mapping detailed structure along the fault plane.
To generate a fault slice, we begin by identifying a fault in the 3-D data set. The fault of interest must be
interpreted at several horizons in order for the computer to calculate the fault plane's projection in x, y, t
space as in Figure 2 .


Figure 2


The more the fault plane deviates from a straight line in the x, y or t directions, the more data points are
required for an accurate projection. Once the fault plane data points are defined, the computer will then fit a
linear projection or slice to the data points by applying a triangulation, least squares or other statistical
method of our choice.
Any data measurement, such as amplitude, frequency, acoustic impedance, or throw, can be plotted in the
fault slice. The type of data displayed is dependent on the purpose of the investigation. For structural detail,
we could display the throw as in Figure 3.


Figure 3.


We could also display amplitudes to define the splinter faults as in Figure 4.


Figure 4.



Case Studies
3-D data have altered our knowledge and understanding of faults. Where we once thought a few long,
continuous faults existed, we now know many short ones to exist instead.
This new understanding is a direct result of the increased data coverage associated with 3-D survey. For
example, previously, a tight 2-D grid was once considered to be 2 km x 2 km (2,000 m x 2,000 m). Now,
with 3-D surveys, a normal grid is 25 m x 25 m! Instead of 4 data points per 4 sq km, we now have 6,400!
This increase in data points has provided us with new detail, particularly in the number of faults, their length
and the shape of structural contours. To exemplify this observation, we will look at three case studies
comparing interpretations based on a 2-D seismic survey to that based on a 3-D survey.
The first case study (Brown, 1991) is of the Gulf of Thailand area and data, shown in Figure 1 and Figure 2 .
Figure 3 shows the structure map derived from a 2-D grid while Figure 4 shows the structure map of the
same area derived from a more recent 3-D survey.


Figure 1





Figure 2





Figure 3





Figure 4


Notice the remarkable change in the fault interpretation. Whereas the 2-D interpretation indicates three fault
systems with a series of intermingling faults, the 3-D interpretation indicates many more separate and
abbreviated faults. Also when we compare the differences in the contour detail, we notice how the contours
in the
3-D interpretation along the faults are very much parallel to the fault strike. Despite their equal contour
intervals, we observe much more detail in the map based on the 3-D survey.
Similar results for several prospects and fields were reported by Nestvold (1992).


Figure 5


Figure 5, which is from that study, offers another comparison of a fault pattern based on 2-D data versus
another of the same area but based on 3-D data. Again, we notice how the faults have increased in number
and have shortened in length.
In the some cases, even the orientation of the faults changes. In another example reported by Nestvold
(1992), a map of the Groningen Gas Field based on 2-D data indicates the principal faults to trend in a
northwest/southeast direction. The more recent interpretation, based on 3-D seismic, indicates that the
dominant faulting actually trends north/south with secondary east/west cross faulting. These maps are
shown in Figure 6 and Figure 7 .


Figure 6





Figure 7


Another case study showing the influence of 3-D data was reported by Bouvier et al. (1989). In this study of
the Nun River Field in Nigeria, the 3-D data not only altered the structural interpretation, but also provided a
means to investigate the sealing and leaking properties of the major faults.
Two vintages of Nun River structure maps are shown in Figure 8 and Figure 9 .


Figure 8





Figure 9


The first map ( Figure 8 ), based on 2-D data, shows a collapsed anticlinal crest dominated by long,
continuous normal faults. The second map ( Figure 9 ), based on 3-D data, shows that the major faults have
the same strike as previously mapped, but the complexity and discontinuity of the faulting have increased.
The major faults in the 3-D interpretation are discontinuous and rapidly vary in throw and hade. In addition,
many minor faults that had been previously unidentified were recognized and mapped.
In this same Nun River Field study by Bouvier et al. (1989), fault slices were used to determine the sealing
and leaking characteristics of a major fault.
In this study, fault slices of acoustic impedance in the depth domain were generated in order to identify
lithologies and clay smears along the fault plane. Figure 10 shows the acoustic impedance fault slice for the
footwall.


Figure 10


The lithologies and fluids along the footwall fault plane were determined by combining this information with
well logs. This procedure was also followed for the hanging wall. By superimposing the interpreted fault slices
from the hanging and footwalls, the juxtaposition of impermeable clays and shales against hydrocarbon-
bearing sandstones was determined, as was a measure of fault sealing and leaking.
REFERENCES

Allen, C. R. 1962. Circum-Pacific faulting in the Philippines-Taiwan region. Jour.
Geophys. Res. 67: 4795812.
Anstey, N. A. 1992. GP204 The Geological Message in the Seismic Trace:
Boston: IHRDC.
Anstey, N. A. 1990. GP203 The reflection process. Boston: IHRDC.
Anstey, N. A. 1987. GP201 Wiggles: A graphical introduction to signal theory.
Boston: IHRDC.
Anstey, N. A. 1983. GP102 The seismic method. Boston: IHRDC.
Anstey, N. A. 1983. GP101 The exploration task. Boston: IHRDC.
Anstey, N. A. and M. E. Badley. 1984. GP501 Basic interpretation. Boston:
IHRDC.
Anstey, N. A. and M. E. Badley. 1988. GP502 Seismic contouring. Boston:
IHRDC.
Arthur, K. R., D. R. Cole, G. L. Henderson and D. W. Kushnir. 1983. Geology
of the Hibernia discovery. AAPG Memoir no.32:181-197.
Atwater, T. 1970. Implications of plate tectonics for the Cenozoic tectonic
evolution of western North America. GSA Bull. 81: 3513-3536.
Badley, M. E. 1985. Practical seismic interpretation. Boston: IHRDC.
Bally, A. W., ed. 1983. Seismic expression of structural styles: A picture and
work atlas. AAPG Studies in Geology Series, no. 15 (3 vols.). Tulsa, OK: AAPG.
Bally, A. W., P. L. Gordy and G. A. Steward. 1966. Structure, seismic data,
and orogenic evolution of southern Canadian Rocky Mountains. Can. Petrol. Geol.
Bull. 14: 337-381.
Bally, A. W. and J. S. Oldow. 1983. Plate tectonics, structural styles, and the
evolution of sedimentary basins. Houston Geological Society, short course
manual.
Barbat, W. F. 1958. The Los Angeles Basin area, California. Habitat of oil, L. G.
Weeks, ed.: 62-77. Tulsa, OK: AAPG.
Biddle, K. T. and D. R. Seely. 1983. Structure of a subduction complex.
Seismic expression of structural styles, A. W. Bally, ed. AAPG Studies in Geology
Series no. 15, v. 3: 3.4.2-129-135.
Billings, M. P. 1972. Structural geology (third edition). Englewood Cliffs, NJ:
Prentice-Hall.
Bishop, R. S. 1978. Mechanism for emplacement of piercement diapirs. MPG
Bull. 62:1561-1583.
Bloom, A. L. 1978. Geomorphology. Englewood Cliffs, NJ: Prentice-Hall.
Bobbit, J. E. and J. 0. Gallagher. 1978. The petroleum geology of the Gulf of
Suez. 1978 Offshore Tech. Conf, Houston, TX. 37-382.
Bouvier, J. D., C. H. Kaars-Sijpesteijn, D. K Kluesner, C. C. Onyejekwe and
R. C. Van Der Pal. 1989. Three-dimensional seismic interpretation and fault
sealing investigations, Nun River Field, Nigeria. MPG Bull. 73: 1397-1413.
Brink, A. H. 1974. Petroleum geology of the Gabon Basin. MPG Bull. 58: 216-
235.
Brown, A. R. 1991. Interpretation of three-dimensional seismic data (third
edition).MPG Memoir 42. Tulsa, OK: AAPG.
Brown, L. F., Jr. and W. L. Fisher. 1977. Seismic-stratigraphic interpretation
of depositional systems: examples from Brazilian rift and pull-apart basins.
Seismic stratigraphy-Applications to hydrocarbon exploration. AAPG Memoir no.
26, C. E. Payton, ed. 213-248.
Bruce, C. H. 1973. Pressure shale and related sediment deformation-
Mechanisms for development of regional contemporaneous faults. MPG Bull. 57:
878-886.
Bruce, C. H. 1983. Shale tectonics, Texas coastal area growth faults. Seismic
Expression of structural styles, A. W. Bally, ed. AAPG Studies in Geology Series
no. 15, v. 2: 3.1-1-6.
Bullard, T. 1973. Eugene Island Block 126 Field, offshore St. Mary Parish,
Louisiana. Offshore Louisiana oil and gas fields. Lafayette Geol. Soc.: 111.
Casella, C. J. 1964. Geologic evolution of the Beartooth Mountains, Montana and
Wyoming-Part 4: Relationship between Precambrian and Laramide structures in
the Line Creek area. GSA Bull. 75: 969-986.
Chapin, C. E. and S. M. Cather. 1983. Eocene tectonics and sedimentation-
Colorado Plateau, Rocky Mountain area. Rocky Mountain foreland basins and
uplifts, J. D. Lowell, ed. RMAG: 33-57.
Christensen, C. K 1983. An example of a major syndepositional listric fault.
Seismic expression of structural styles, A. W. Bally, ed. AAPG Studies in Geology
Series no. 15, v. 2: 2.3.1-30.
Cloos, E. 1947. Oolite deformation in the South Mountain Fold, Maryland. GSA
Bull. 58: 843-918.
Cloos, H. 1939. Conversations with the Earth: Munich: Piper & Co.
Cook, F. 1982. The COCORP seismic reflection traverse across the southern
Appalachians. AAPG Studies in Geology Series no. 14: 61p.
Cook, F. et al. 1979. Thin-skinned tectonics in the crystalline southern
Appalachians; COCORP seismic reflection profiling of the Blue Ridge and
Piedmont. Geology 7: 563-567.
Crowell, J. C. 1974. Origin of late Cenozoic basins in southern California.
Tectonics and Sedimentation. SEPM Special Publications no. 22: 190-204.
Currelle, R. and R. Marco. 1983. Reflection profiles across the Aquitaine Basin
(salt tectonics). Seismic expression of structural styles, A. W. Bally, ed. AAPG
Studies in Geology Series no. 15, v. 2: 2.3.2-11-18.
Dahlstrom, C. D. A. 1970. Structural geology in the eastern margin of the
Canadian Rocky Mountains. CSPG. Bull. 18: 33206.
Dailly, G. C. 1976. A possible mechanism relating progradation, growth faulting,
clay diapirism, and over-thrusting in a regressive sequence of sediments. CSPG
Bull. 24: 92-116.
Davies, P. A. and S. A. Runcorn. 1980. Mechanics of continental drift and plate
tectonics. London: Academic Press.
Davis, G. H. 1984. Structural geology of rocks and regions. New York: John
Wiley & Sons.
De Jong, K. A. and R. Scholten, ed. 1973. Gravity and tectonics. New York:
John Wiley & Sons.
Dennis, J. G. 1972. Structural geology. New York: John Wiley & Sons.
De Sitter, L. U. 1964. Structural geology. New York: McGraw-Hill.
Dickinson, W. R. 1976. Plate tectonic evolution of sedimentary basins. Plate
tectonics and hydrocarbon accumulation. AAPG Course Note Series no.1, 1-56.
Dieterich, J, H. 1969. Origin of cleavage in folded rocks. Am. Jour. Sci. 267:
155-165.
Dillion, W. P., C. K. Paull, R. T. Buffler and J. P. Fail. 1979. Structure and
development of the southeast Georgia embayment and northern Blake plateau:
Preliminary analysis. Geological and geophysical investigations of continental
margins. AAPG Memoir no. 29, J. S. Watkins, L. Montadert and P. W. Dickerson,
ed. 273.
Douglas, R. J. W. 1950. Callum Creek, Langford Creek, and Gap map-areas,
Alberta. Geol. Survey of Canada Memoir no.255.
Elliot, D. 1976. The motion of thrust sheets. Jour Geophys. Res. 81: 949-963.
Erxleben, A. W., and G. Carnahan. 1983. Slick Ranch area, Starr County,
Texas. Seismic expression of structural styles, A. W. Bally, ed. AAPG Studies in
Geology Series no. 15, v. 2: 2.3.1-22-27.
Evans, R. 1978. Origin and significance of evaporites in basins around the
Atlantic margin. AAPG Bull. 62: 223-234.
Fischer, A. G. 1969. Geologic time-distance rates: the Bubnoff unit. GSA Bull.
80: 549-551.
Fischer, A. G. and S. Judson, ed. 1975. Petroleum and global tectonics.
Princeton, NJ: Princeton Univ. Press.
Flint, R. F. 1971. Glacial and quaternary geology. New York: John Wiley & Sons.
Fox, K G. 1959. Structure and accumulation of hydrocarbons in southern
foothills, Alberta, Canada. AAPG Bull. 43: 992-1023.
Gansser, A. 1964. Geology of the Himalayas: New York: Interscience.
Gibbs, A. D. 1983. Balanced cross-section construction from seismic sections in
areas of extension tectonics. Jour. Struct. Geol. 5 (2)153.
Gibson, I. L. and G. P. L. Walker. 1964. Some composite rhyolite-basalt lavas
and related composite dykes in eastern Iceland. Proc. Geol. Assoc. London 74:
301-318.
Gries, R. 1983. North-south compression of Rocky Mountain foreland structures.
Rocky Mountain Foreland Basins and Uplifts (J. D. Lowell, ed.): RMAG: 9-33.
Griggs, D. T., and J. Handin. 1960. Observations on fracture and a hypothesis
of earthquakes. Rock Deformation - A Symposium, GSA Memoir no. 79 (D. T.
Griggs and J. Handin, ed.): 347-373.
Grow, J. A., R. E. Mattick, and J. S. Schlee. 1979. Multichannel seismic depth
sections and interval velocities over outer continental shelf and upper continental
slope between Cape Hatteras and Cape Cod.
Halbouty, M. T. 1979. Salt domes, Gulf Region, United States and Mexico, 2nd
ed.: Houston: Gulf Publ. Co.
Handin, J. W., R. V. Hager, M. Friedman, and J. M. Feather. 1963.
Experimental deformation of sedimentary rocks under confining pressure: pore
pressure tests. AAPG Bull. 47: 717-755.
Harding, T. P. 1985. Seismic characteristics and identification of negative flower
structures, positive flower structures, and positive structural inversion. AAPG Bull.
69: 582-601.
Harding, T. P. 1984. Graben hydrocarbon occurrences and structural style.
AAPG Bull. 68: 333-362.
Harding, T. P. 1983a. Divergent wrench fault and negative flower structure,
Andaman Sea. Seismic expression of structural styles, A. W. Bally, ed. AAPG
Studies in Geology Series no. 15, v. 3: 4.2-1-9.
Harding, T. P. 1983b. Structural inversion at Rambutan Oil Field, South
Sumatra Basin. Seismic expression of structural styles, A. W. Bally, ed. AAPG
Studies in Geology Series no. 15, v. 3: 3.3-13-18.
Harding, T. P. 1976. Tectonic significance and hydrocarbon trapping
consequences of sequential folding synchronous with San Andreas faulting, San
Joaquin Valley, California. AAPG Bull. 60: 356-378.
Harding, T. P. 1974. Petroleum traps associated with wrench faults. AAPG Bull.
58: 1290-1304.
Harding, T. P. and J. D. Lowell. 1979. Structural styles, their plate-tectonic
habitats and hydrocarbon traps in petroleum provinces. AAPG Bull. 63: 1016-
1069.
Hatcher, R. D., Jr. 1972. Development model of the southern Appalachians.
GSA Bull. 83:2735.
Heard, H. C. 1960. Transition from brittle to ductile flow in Solenhofen limestone
as a function of temperature, confining pressure, and interstitial fluid pressure.
Rock deformation-A symposium. GSA Memoir no. 79, D. T. Griggs and J. Handin,
ed. 193-226.
Heard, H. C. and C. B. Raleigh. 1972. Steady-state flow in marble at 500 C to
800 C. GSA Bull. 83: 935-956.
Hills, E. S. 1971. Elements of structural geology (second edition). London:
Methuen and Co., Ltd.
Hobbs, B. E., W. D. Means and P. E. Williams. 1976. An outline of structural
geology. New York: John Wiley & Sons.
Hoeppener, R. 1964. Zur physikalischen tektonik. Felsmechanik und
Ingenjeurgeologie 2: 224.
Holland, D. S., C. E. Sutley, R. E. Berlitz and J. A. Gilreath. 1976. East
Cameron Block 270: A Pleistocene field. North American oil and gas fields. AAPG
Memoir no. 24, J. Braunstein, ed. 205-228.
Huff, K. F. 1980. Frontiers of world exploration. Facts and principles of world
petroleum occurrence. CSPG Memoir no. 6, A. D. Miall, ed. 343-362.
Huff, K. F. 1978. Frontiers of world exploration. Oil and Gas Jour. 76:214-220.
Inoue, E. 1960. Land deformation in Japan. Geograph. Sun;. Inst (Japan) Bull.
6: 73-134.
Jackson, J. and D. P. McKenzie. 1983. The geometric evolution of normal fault
systems. Jour. of Structural Geology 5: 47182.
Jackson, M. P. and W. E. Galloway. 1984. Structural and depositional styles of
Gulf Coast tertiary continental margins: Application to hydrocarbon exploration.
AAPG Continuing Education Course Notes Series, no. 25.
Jenyon, M. K. and A. A. Fitch. 198 5. Seismic reflection interpretation. Berlin:
Gebruder Bomtraeger.
Kennedy, W. G. 1946. The Great Glen Fault. Quart. Jour. Geol. Soc. London
102: 41-72.
Klemme, H. D. 1976. World oil and gas reserves from analysis of giant oil fields
and petroleum basins (provinces). The future supply of nature-made petroleum
and gas, R. K Meyer, ed. New York: Pergamon Press: 217-260.
Kleyn, A. H. 1983. Seismic reflection interpretation. London: Applied Science
Publishers.
Lamerson, P. R. 1982. The fossil basin area and its relationship to the Absaroka
thrust fault system. Geologic studies of the Cordilleran thrust belt, R. B. Powers,
ed. RMAG: 279-340.
Lehner, P., H. Doust, G. Bakker, P. Allenbach and J. Gueneau. 1983. Active
margins-Caribbean margin of South America, profiles C-1422, C- 1412, and C-
1413. Seismic expression of structural styles, A. W. Bally, ed. AAPG Studies in
Geology Series no. 15, v. 3: 3.4.2-111-129.
Ligget, M. A. and H. E. Ehrenspeck. 1974. Pahranagat shear system, Lincoln
County, Nevada. Argus Explor. Co., Rept. of Inv., NASA CR 1356388, E74-10206.
Lowell, J. D. 1985. Structural styles and petroleum exploration. Tulsa, OK:
OGCI Pub.
Lowell, J. D., G. J. Genik, T. H. Nelson and P. M. Tucker. 1975. Petroleum
and plate tectonics of the southern Red Sea. Petroleum and global tectonics, A. G.
Fisher and S. Judson, ed. Princeton, NJ: Princeton Univ. Press. 129-153.
Maher, C. E. 1980. Piper Oil Field. Giant oil and gas fields of the decade 1968-
1978. AAPG Memoir no. 30, M. Halbouty, ed. 131-173.
Manspeizer, W. 1981. Early Mesozoic basins of the central Atlantic passive
margins. AAPG Continuing Education Course Notes Series, no. 19.
Martin, R. G. and J. E. Case. 1975. Geophysical studies in the Gulf of Mexico.
The Oceans Basins and Margins, V. 3, A. E. Nairn and F. G. Stelhi, ed. New York:
Plenum Press.
Mescherikov, Y. A. 1968. Neotectonics. Encyclopedia of geomorphology. R. W.
Fairbridge, ed. New York: Van Nostrand. 768-773.
Mitchum, R. M., P. R. Vail and S. Thompson. 1977. The depositional
sequence as a basic unit for stratigraphic analysis. Seismic stratigraphy-
Applications to hydrocarbon exploration. AAPG Memoir no. 26, C. E. Payton, ed.
53-62.
Montgomery, S. L. 1984. Michigan Basin: Expanding the deep frontier.
Petroleum Frontiers 1 (Spring). Denver: Petroleum Information Corp.
Montgomery, S. L 1987. GP601 Structural geology. Boston: IHRDC.
Morgan, L. and W. Dowdall. 1983. The Atlantic continental margin. Seismic
expression of structural styles, A. W. Bally, ed. AAPG Studies in Geology Series
no. 15, v. 2: 2.2.3-30-36.
Nadai, A. 1950. Theory of flow and fracture of solids. New York: McGraw-Hill.
Nettleton, L. L. 1934. Fluid mechanics of salt domes. AAPG Bull. 18:1175-1204.
Nestvold, E. 0. 1992. 3-D seismic: Is the promise fulfilled? The Leading Edge
(SEG) 11(6): 12-19.
New Orleans Geological Society. 1983. Oil & gas fields of southeast Louisiana,
v. 3. New Orleans, LA.
Park, R. G. 1983. Foundations of structural geology. London: Chapman & Hall.
Pew, E. 1983. Seismic structural analysis of deformation in the southern
Mexican Ridges. Master's thesis: University of Texas, Austin, TX.
Pierce, W. G. 1957. Heart Mountain and South Fork detachment thrusts of
Wyoming. AAPG Bull. 41: 591-626.
Pieri, M. 1983. Three seismic profiles through the Po Plain. Seismic expression
of structural styles, A. W. Bally, ed. AAPG Studies in Geology Series no. 15, v. 3:
3.4.1-8-26.
Plafker, G. 1965. Tectonic deformation associated with the 1964 Alaskan earth-
quake. Science 148:1675-1687.
Price, R. A. and E. W. Mountjoy. 1970. Geologic structure of the Canadian
Rocky Mountains between Bow and Athabasca Rivers-A progress report. Geol.
Assoc. Canada Spec. Paper 6. 25 p.
Ragan, D. M. 1984. Structural geology: An introduction to geometrical
techniques (third edition). New York: John Wiley & Sons.
Ramsay, J. G. 1967. Folding and fracturing of rocks. New York: McGraw-Hill.
Ramsay, J. G. and M. I. Huber. 1985. The techniques of modern structural
geology. London: Academic Press.
Reading, H. G. 1980. Characteristics and recognition of strike-slip fault systems.
Sedimentation in oblique-slip mobile zones. Int. Assoc. of Sed. Spec. Pub. 4, P. F.
Ballance and H. G. Reading, ed. 7-26.
Richter, C. K 1952. Elementary seismology. New York: W. H. Freeman.
Robson, D. A. 1971. The structure of the Gulf of Suez rift, with special reference
to the eastern side. Jour. Geol. Soc. 127: 247-276.
Roeder, D. H. and 0. E. Gilbert. 1977. Structure, kinematics, and hydrocarbon
prospects of thrust and fold belts. AAPG Structural Geology School Course Notes.
Rodgers, J. 1953. The folds and faults of the Appalachian Valley and Ridge
province. Kentucky Geol. Survey Spec. Pub. 1:150-166.
Roux, W. K, Jr. 197 7. The development of growth fault structures. AAPG
Structural Geology School Course Notes.
Royse, K, Jr., M. A. Warner and D. L. Reese. 1975. Thrust belt structural
geometry and related stratigraphic problems, Wyoming-Idaho-northern Utah.
Deep drilling frontiers in the central Rocky Mountains, D. W. Boylyard, ed. RMAG
Symposium: 415.
Sacrison, W. R. 1978. Seismic interpretation of basement block faults and
associated deformation. Laramide folding associated with basement block faulting
in the western United States, V. Mathews, ed. GSA Memoir no. 26. 165-184.
Sbar, M. L. and L. R. Sykes. 1973. Contemporary compressive stress and
seismicity in eastern North America: An example of intra-plate tectonics. GSA
Bull. 84:1861-1881.
Seely, D. R., and W. R. Dickinson. 1977. Structure and stratigraphy of forearc
regions. AAPG Continuing Education Course Notes Series, no. 5.
Selley, R. C. and D. C. Morrill. 1983. GL107 The habitat of hydrocarbons in
sedimentary basins. Boston: IHRDC.
Seni, S. J. and M. P. Jackson. 1984. Evolution of salt structures, east Texas
diapir province, Parts 1 and 2. AAPG Bull. 67:1245-1274.
Seyfert, C. K. and L. A. Sirkin. 1973. Earth history and plate tectonics. New
York: Harper & Row.
Sodbinow, E. S. 1984. GP401 Basic processing. Boston: IHRDC.
Spicher, A. 1980. Tectonics profile through Switzerland. Geology of Switzerland,
Swiss Geol. Comm.
Spillers, J. P. 1965. Distribution of hydrocarbons in South Louisiana by type of
traps and trends-Frio and younger sediments. Trans. Gulf Coast Assn. Geol Soc
15:37-39.
Sprague, E. L. 1983. Geology of the Tepee Flats-Bulldog Fields, Natrona County,
Wyoming. Rocky Mountain foreland basins and uplifts, J. D. Lowell, ed. RMAG:
339-345.
Stone, D. S. 1983a. Seismic profile-South Elk Basin. Seismic expression of
structural styles, A. W. Bally, ed. AAPG Studies in Geology Series no. 15, v. 3:
3.4.1-8-26.
Stone, D. S. 1983b. The Greybull sandstone pool (lower Cretaceous) on the Elk
Basin thrust-fold complex, Wyoming and Montana. Rocky Mountain foreland
basins and uplifts, J. D. Lowell, ed. RMAG: 345-357.
Stude, G. R. 1978. Depositional environments of the Gulf of Mexico, South
Timbealer Block 54: Salt dome and salt-dome growth models. Trans Gulf Coast
Assn of Geol. Soc. 28: 627-644.
Sunwall, N. T., K. A. McQuillan and C. J. Nick. 1983. Salt diapir-Gulf of
Mexico. Seismic expression of structural styles, A. W. Bally, ed. AAPG Studies in
Geology Series no. 15, v. 2: 2.3.2-32-38.
Swann, D. H. and A. H. Bell. 1958. Habitat of oil in the Illinois Basin. Habitat of
oil, L. G. Weeks, ed. Tulsa, OK: AAPG. 44772.
Tearpock, D. J. and R. E. Bischke. 19 91. Applied subsurface geological
mapping. Englewood Cliffs, NJ: Prentice-Hall.
Trusheim, F. 196 0. Mechanism of salt migration in northern Germany. MPG
Bull. 44:1519-1541.
Turcotte, D. and G. Schubert. 1982. Geodynamics. New York: John Wiley &
Sons.
Vail, P. K., R. N. Mitchum, R. G. Todd, J. N. Widmier, S. Thompson, J. N.
Bubb and W. G. Hatfield. 1977. Seismic stratigraphy and global changes in sea
level, Parts 1 and 2. AAPG Memoir no. 26: 49-213.
Vine, K J. 1966. Spreading of the ocean floor: New evidence. Science 154:
1405-1415.
Wallace, R. E. 1970. Earthquake recurrence intervals on the San Andreas Fault.
GSA Bull. 81: 2875-2890.
Watson, J. M. and C. A. Swanson. 197 5. North Sea-Major petroleum province.
AAPG Bull. 59:1098-1112.
Wennekers, J. H. L. 1958. South Sumatra basinal area. Habitat of oil, L. G.
Weeks, ed. Tulsa, OK: AAPG. 1347-1358.
Wilson, J. T. 1965. A new class of faults and their bearing on continental drift.
Nature 207: 343-347.
Withjack, N. 0. and C. Scheiner. 1982. Fault patterns associated with domes-
An experimental and analytical study. MPG Bull. 66: 302-316.
Wood, D. S. 1974. Current views of the development of slaty cleavage. Ann.
Rev. Earth Sci. 2:1-35.
Woodbury, H. 0., I. B. Murray, Jr. and R. E. Osborne. 1980. Diapirs and their
relation to hydrocarbon accumulation. Facts and principles of world petroleum
occurrence. CSPG Memoir no.6, A. D. Miall, ed. 119-143.
Ziegler, P. A. 1980. Hydrocarbon provinces of the northwest European basins.
Facts and principles of world petroleum occurrence. CSPG Memoir no.6, A. D.
Miall, ed. 653-705


ADDITIONAL REFERENCES

Alam, A. and Caragounis, P. (1994). "Advances in 3D seismic fault
interpretation." Abstract 56 th Annual Meeting. Houten: European Association of
Geoscientists & Engineers.
Anderson, J., Thomas, C., Kat, H. and Lynn, W. (1994). "Velocity and
structural interpretation with 3D pre-stack depth migration." Abstract 56 th
Annual Meeting. Houten: European Association of Geoscientists & Engineers.
Anfiloff, V. (1989). "Discussion on 'Structural interpretation of the Rukwa Rift,
Tanzania', by J. W. Peirce and L. Lipkov." Geophysics, 54, 1499-1500. Tulsa:
Society of Exploration Geophysicists.
Bally, A. W. (2001). "On the Role of Plate Tectonics in Hydrocarbon
Exploration." Annual AAPG Meeting Expanded Abstracts, 11. Tulsa: American
Association of Petroleum.
Bally, A. W. (1991)."What Else to Do with Reflection Seismology." Abstract 61
st Annual International Meeting, 722 . Tulsa: Society of Exploration
Geophysicists.
Barton, C. A., Moos, D. and Zoback, M. D. (1997). "In-situ Stress
Measurements Can Help Define Local Variations in Fracture Hydraulic Conductivity
at Shallow Depth." The Leading Edge, 16, 11, 1653-1656 . Tulsa: Society of
Exploration Geophysicists.
Brown, A. R. (1983). "Structural Interpretation from Horizontal Seismic
Sections." Geophysics, 48, 1179-1194. Tulsa: Society of Exploration
Geophysicists.
Brown, A. (1982). "Structural Interpretation from Horizontal Seismic Sections."
Expand Abstract 52 nd Annual International Meeting. Tulsa: Society of Exploration
Geophysicists.
Bruno, M. S. and Winterstein, D. F. (1994). "Some Influences of Stratigraphy
and Structure on Reservoir Stress Orientation." Geophysics, 59, 924-962. Tulsa:
Society of Exploration Geophysicists.
Chopra, S. and Sudhakar, V. (2000)." Fault Interpretation; the Coherence
Cube and Beyond." Oil and Gas Journal,98, 31, 71-74. Tulsa: PennWell.
Cloke, I. R., Craig, J. and Blundell, O. (1998). "Variations in Stress Regime
through Time - their Effect on the Formation and Modification of the Kutai Basin,
Kalimantan." Abstract 60 th Annual Meeting. Houten: European Association of
Geoscientists & Engineers.
Colletta, B., Piron, E., Marchal, D. and Moretti, I. (2002). "Gravitational
Folds and Compressional Structures Related with Normal Faults in the Suez Rift."
Abstract 64 th Annual Meeting. Houten: European Association of Geoscientists &
Engineers.
Csato, I. and Foldes, T. (1999). "3D Visualisation of Architectural Patterns in
Salt-rift Basins." Abstract 61 st Annual Meeting. Houten: European Association of
Geoscientists & Engineers.
Dancer, P.N. (2002). "The Corrib Field - Changes in Structural Interpretation
through 2D, 3D and PreSDM Seismic Data." Abstract 64 th Annual Meeting.
Houten: European Association of Geoscientists & Engineers.
Dancer, P.N. (2001). "The Corrib Field - Impact of 2D, 3D and PreSDM Seismic
Data on the Structural Interpretation." Abstract 63 rd Annual Meeting. Houten:
European Association of Geoscientists & Engineers.
Davies, R. K., Crawford, M., Dula, W. F., Jr., Cole, M. J. and Dorn, G. A.
(1997). "Outcrop Interpretation of Seismic-scale Normal Faults in Southern
Oregon: Description of Structural Styles and Evaluation of Subsurface
Intepretation Methods" The Leading Edge, 16, 08, 1135-1141. Tulsa: Society of
Exploration Geophysicists.
Destro, N., Szatmari, P., Alkmim, F. and Magnavita, L. P. (2000). "Release
Faults, Associated Structures, and their Control on Petroleum Trends in the
Reconcavo Rift, Northeast Brazil." AAPG Bulletin, 87, 7, 1123-1144. Tulsa:
American Asotiation Petroleum Geologists.
Deville, E. and Collaku, A. (1999). "The Front of the Central Albanian Thrust
Belt: Structural Interpretation and Sub-thrust Plays." Abstract 61 st Annual
Meeting. Houten: European Association of Geoscientists & Engineers.
Driggs, A. F. (1997). "The Use of 3-D fault Visualization and Reservoir
Juxtaposition Analysis in an Exploration Setting, U. S. Gulf of Mexico." Abstract
American Association of Petroleum Geologists 1997 Annual Convention, 6, 70.
Tulsa: American Association of Petroleum Geologists.
Edwards, A. (1989). "Basin Margin Structural Styles, Jeanne d'Arc Basin
Offshore Newfoundland, Canada." Abstract 59 th Annual International Meeting,
538 . Tulsa: Society of Exploration Geophysicists.
Fehmers, G. C. and Hocker, C. (2003). "Fast Structural Interpretation with
Structure-oriented Filtering." Geophysics, 68, 1286-1293. Tulsa: Society of
Exploration Geophysicists.
Gawthorpe, R., Jackson, C., Young M., Sharp, I., Moustafa, A. and
Leppard, C. (2003). "Normal Fault Growth, Displacement Localization and the
Evolution of Normal Fault Populations: the Hammam Faraun Fault Block, Suez
Rift, Egypt." Journal of Structural Geology, 25, 6, 883-885. London: Reed
Elsevier.
Geraghty, D., Shannon, P. M and Feely, M. (1999). "Sedimentary Provenance
of Rift Sediments of the Porcupine Basin, West of Ireland." Abstract 61 st Annual
Meeting. Houten: European Association of Geoscientists & Engineers
Gibson, R. I. (1994). "Structural Styles of the West Siberian Rift System and
Their Impact on Hydrocarbon Accumulation." Abstract 64 th Annual International
Meeting, 1626 . Tulsa: Society of Exploration Geophysicists.
Goulty, N. R. (1998). "Relationships Between Porosity and Effective Stress in
Shales." First Break, 16, 2, 413-419. Houten: European Association of
Geoscientists & Engineers.
Grandi, S., Rao, R., Huang, X. and Toksoz, N. (2003). "In Situ Stress
Modeling at a Borehole - A Case Study." Abstract 73 rd Annual International
Meeting, 297-300 . Tulsa: Society of Exploration Geophysicists.
Gran t, J. and Kattenhorn, S. (2004). "Evolution of Vertical Faults at an
Extensional Plate Boundary, Southwest Iceland." Journal of Structural Geology,
26, 3, 537-557. London: Reed Elsevier.
Grechka, V., Bakulin, A. and Tsvankin, I. (2001). "Seismic Characterization
of Vertical Fractures Described as General Linear-slip Interfaces." Abstract 71 st
Annual International Meeting, 57-60 . Tulsa: Society of Exploration Geophysicists.
Hall, D. H. (1976). "Crustal studies: Their Implications for the Explorationist."
Exploration Geophysics, 07, 53-59. Sydney: Australian Society of Exploration
Geophysicists.
Hedlund, C. A., Couzens-Schultz, B. A., Maler, M. O., Garmezy, L. and
Naruk, S. J. (1999). "The State-of-the-art in Interpretation of Complex
Structure from 2D and 3D Seismic Data." Abstracts and Programs, Geological
Society of America Meeting, 31, 7, 127. Boulder: Geological Society of America.
Hesthammer, J. (1999). "Improving Seismic Data for Detailed Structural
Interpretation." The Leading Edge, 18, 02, 226-247 . Tulsa: Society of
Exploration Geophysicists.
Hesthammer, J., (1998). "Evaluation of the Timedip, Correlation and
Coherence Maps for Structural Interpretation of Seismic Data." First Break, 16, 5,
151-167. Houten: European Association of Geoscientists & Engineers.
Hesthammer, J. and Fossen, H. (1997). "The Influence of Seismic Noise in
Structural Interpretation of Seismic Attribute Maps." First Break, 15, 6, 209-219.
Houten: European Association of Geoscientists & Engineers.
Hockers, C. and Fehmers, G. (2002). "Fast Structural Interpretation with
Structure-oriented Filtering." The Leading Edge, 21, 03, 238-243 . Tulsa: Society
of Exploration Geophysicists.
Huang, X., Sinha, B., Burns, D. and Toksoz, M. (1999). "Formation Stress
Estimation Using Standard Acoustic Logging." Abstract 69 rd Annual International
Meeting, 53-56 . Tulsa: Society of Exploration Geophysicists.
Huber, W. (1988). "Thrust Fault, Gulf of Mexico (Ewing Bank area)." Bulletin -
Houston Geological Society, 30, 7, 17-18. Houston: Houston Geological Society.
Iran, B. (2003). "Structural Styles in the Zagros Simple Folded Zone." Journal of
the Geological Society of London, 160, 3, 401-412. London: Geological Society of
London.
Ivanova, N., Verba, M., Sakoulinna, T. and Matveev, J. (2000). "Rift
Structures of the Barents Plate from Regional Geophysical Investigation." Abstract
70 th Annual International Meeting, 709-712 . Tulsa: Society of Exploration
Geophysicists.
Jacobs, J. A. C., Jardin, A., Delprat-Jannaud, F., Versteeg, R. and Lailly, P.
(1991). "Prestack Structural Interpretation of Seismic Data." Expand Abstract 61
st Annual International Meeting, 188-191 . Tulsa: Society of Exploration
Geophysicists.
Jackson, M. P. A., Schultz-Ela, D. D., Hudec, M. R., Watson, I. A., and
Porter, M. L. (1998). "Structure and Evolution of Upheaval Dome: A Pinched-Off
Salt Diaper." G eological Society of America Bulletin, 110, 1547-1573. Boulder:
Geological Society of America
Jones, G. and Knipe, R. J. (1996). "Seismic Attribute Maps: Application to
Structural Interpretation and Fault Seal." First Break, 14, 12. Houten: European
Association of Geoscientists & Engineers.
Jones, P. B. (1988). "Seismic Interpretation 10 - Balanced Cross-sections - An
Aid to Structural Interpretation." The Leading Edge, 07, 08, 29-31 . Tulsa:
Society of Exploration Geophysicists.
Jones, R. M. (2003). "An Integrated, Quantitative Approach to Assessing Fault-
seal Risk." AAPG Bulleting, 87, 3, 507-524. Tulsa: American Association of
Petroleum Geologists.
Kattenhorn, S. A. and Pollard, D. (2001). " Integrating 3-D Seismic Data,
Field Analogs, and Mechanical Models in the Analysis of Segmented Normal Faults
in the Wytch Farm Oil Field, Southern England, United Kingdom." A.A.P.G.
Bulletin, 85, 1183-1210. Tulsa: American Association of Petroleum Geologists.
Kligfield, R., Rowan, M. G. and Ratliff, R. (1989). "Applications of Computer-
aided Section Construction, Restoration, and Balancing Models to the Petroleum
Industry." Abstract 59 th Annual International Meeting, 589 . Tulsa: Society of
Exploration Geophysicists.
Koledoye, A.B., Aydin, A. and May, E. (2000). "Three-dimensional
Visualization of Normal Fault Segmentation and its Implication for Fault Growth."
The Leading Edge, 19, 07, 692-701. Tulsa: Society of Exploration Geophysicists.
Konischev, V. (1998). "Rift Halokinesis." Abstract 60 th Annual Meeting.
Houten: European Association of Geoscientists & Engineers.
Konyukhov, A. I., Sokolov, B. A. and Yandarbiev, N. S. (1998). "Diapirizm,
Mud Volcanizm and Folding in the Sedimentary Basins of Northern Peritethys."
Abstract 60 th Annual Meeting. Houten: European Association of Geoscientists &
Engineers.
Le Turdu, C., Keskes, N., Jeanjean, F., Kristvik, I., Ringenbach, J. and
Straw, P. (1999). "New Tools for 3D Fault Detection by SISMAGE (super TM);
Implications to Interpretation of Complex Faulted Zones of the North Sea."
A.A.P.G. Bulletin, 83, 8, 1324. Tulsa: American Association of Petroleum
Geologists.
Looff, K. M. (2001). "Recent Salt Related Uplift and Subsidence at Sour Lake
Salt Dome, Hardin County, Texas." Transactions, 51, 187-194. New Orleans: Gulf
Coast Association of Geological Societies.
McCabe, P. J. (1984). "Influence of Strike-slip Movement on Terrestrial
Sedimentation in Upper Carboniferous of Nova Scotia and New Brunswick."
Abstract 54 rd Annual International Meeting, 57-60 . Tulsa: Society of Exploration
Geophysicists.
Mehillka, L., Dhima, S. and Basha, J. (1999). "A New Option on Structural
Styles and Independent Petroleum Systems in Outer Albanides." Abstract 61 st
Annual Meeting. Houten: European Association of Geoscientists & Engineers.
Mitra, S . (1990). "Fault-Propagation Folding: Geometry, Kinematic Evolution
and Hydrocarbon Traps." A.A.P.G. Bulletin, 74, 921-945. Tulsa: American
Association of Petroleum Geologists.
Mitra, S . (1988). "Effects of Deformation Mechanisms on Reservoir Potential in
the Central Appalachians." A.A.P.G. Bulletin, 72, 536-554. Tulsa: American
Association of Petroleum Geologists.
Mitra, S . (1988). "Three-Dimensional Geometry and Kinematic Evolution of the
Pine Mountains Thrust System, Southern Appalachians." Geological Society of
America Bulletin, 100, 72-95. Boulder: Geological Society of America.
Nilsen, T. H. and Sylvester, A. G. (1999). "Strike-slip Basins: Part 1." The
Leading Edge, 18, 10, 1146-1152 . Tulsa: Society of Exploration Geophysicists.
Nilsen, T. H. and Sylvester, A. G. (1999). "Strike-slip Basins: Part 2." The
Leading Edge, 18, 11, 1258-1267 . Tulsa: Society of Exploration Geophysicists.
Moretti, I., Delhomme, J. P., Cornet, F., Bernard, P., Schmidt-
Hattenberger, C. and Borm, G. (2002). "The Corinth Rift Laboratory:
monitoring of active faults." First Break, 20, 20, 91-97. Tulsa: Society of
Exploration Geophysicists.
Murtuzaev, I. and Bayramov, A. (1998). "Strike-Slip Faults of the South
Caspian Basin - Azerbaijan and Adjacent Regions." Abstract 60 th Annual Meeting.
Houten: European Association of Geoscientists & Engineers.
Ott, V. D., and Eaves, G. P. (1998). "3-D seismic Examples of Structures from
the Permian Basin." Abstract American Association of Petroleum Geologists 1998
Annual Meeting. Tulsa: American Association of Petroleum Geologists.
Park, J. O., Tsuru, T., Kaneda, Y., Kinoshita, H., Kodaira, S., Kono, Y. and
Takahashi, N. (1999). "New Seismic Reflection Images of the Nankai
Subduction Zone Off Southwestern Japan from JAMSTEC Cruise Seamount
Subduction and Seismic Thrust Fault." Expand Abstract 69 nd Annual
International Meeting, 1009-1012 . Tulsa: Society of Exploration Geophysicists.
Parkin, E. L., Lockman, D. F. and Schwalbach, J. R. (1998). "Fault Detection
and Analysis Using Borehole Image Logs at Pescado Field, Santa Ynez Unit,
Offshore California." Abstract A.A.P.G. Pacific Section Meeting, 82, 5, 885. Tulsa:
American Association of Petroleum Geologists.
Pawlowski, R. (1999). "Megaregional Rift-drift Structural Controls on
Hydrocarbon Accumulations Offshore West Africa." The Leading Edge, 18, 5, 600-
603. Tulsa: Society of Exploration Geophysicists.
Raditchev, R., Dimovski, S., Stavrev, P. and Gerovska, D. (1999). "Various
Density Models of Rift Zones." Abstract 61 st Annual Meeting. Houten: European
Association of Geoscientists & Engineers.
Ratliff, R., Rowan, M., Forester, R. and White, G. (1989). "Section
Restoration and Balancing as an Aid yo Seismic Interpretation: GSI Line TWTS-
83-13 (Regional Well-Tie Line, Offshore Galveston, Texas)." Abstract 59 th Annual
International Meeting, 435 . Tulsa: Society of Exploration Geophysicists.
Roberts, M.J. and Reilly, S.G. (2002). "Frontier Exploration on the Abyssal
Plain of the Deepwater Gulf of Mexico - Syn-Rift through Passive Stage Mesozoic
Objectives Identified." Abstract 64 th Annual Meeting. Houten: European
Association of Geoscientists & Engineers.
Rodnikova, R. D. (1998). "HC Resources of the Sea of Okhotsk Rift Basins in
the Light of Fuel Demand in NE Asia." Abstract 60 th Annual Meeting. Houten:
European Association of Geoscientists & Engineers.
Rouby, D. and Suppe, J. (1996). "Restoration in 3D of Folded and Faulted
Surfaces."
Abstract Geological Society of America, 28th Annual Meeting, 28, 7, 241. Boulder:
Geological Society of America (GSA) : Boulder, CO, United States
Sabat, F., Muos, J. A., Poblet, J., Roca, E. and Verges J. (1997). "Bed-by-
bed Fold Growth by Kink-band Migration: Sant Lloren de Morunys, Eastern
Pyrenees." Journal of Structural Geology, 19, 443-461. London: Reed Elsevier.
Sayers, C. M. (2002). "Stress-dependent Elastic Anisotropy of Sandstones."
Geophysical Prospecting, 50, 85-95 . Houten: European Association of
Geoscientists & Engineers.
Schneider, F., Bouteca, M. and Vasseur, G. (1994). "Validity of the
Porosity/effective-stress Concept in Sedimentary Basin Modeling." First Break, 16,
2, 413-419. Houten: European Association of Geoscientists & Engineers.
Scheidhauer, M., Beres, M., Dupuy, D. and Marillier, F. (2001). "A High-
Resolution 2D/3D Seismic Study of a Thrust Fault Zone in Lake Geneva,
Switzerland." Abstract 61 rd Annual Meeting. Houten: European Association of
Geoscientists & Engineers.
Schultz-Ela, D. D. (2003). "Origin of Drag Folds Bordering Salt Diapers." AAPG
Bulletin, 87, 5, 757-780. Tulsa: American Asotiation Petroleum Geologists.
Tang, X., Cheng, A. and Cheng, N. (1999). "Formation Stress Determination
from Borehole Acoustic Logging: A Theoretical Foundation." Abstract 69 th Annual
International Meeting, 57-60 . Tulsa: Society of Exploration Geophysicists.
Toomey, A. (2003). "Seismic Monitoring of Stress on Fractures/Faults." Abstract
73 rd Annual International Meeting, 2263-2266 . Tulsa: Society of Exploration
Geophysicists.
Velaj, T. (1999)."Structural Styles and Exploration Opportunities from
Thrustbelt to Platform in Western Albania." Abstract 61 st Annual Meeting.
Houten: European Association of Geoscientists & Engineers.
Viola, G, Odonne, F. and Mancktelow N. S. (2004). " Analogue Modelling of
Reverse Fault Reactivation in Strike-slip and Transpressive Regimes: Application
to the Giudicarie Fault System, Italian Eastern Alps." Journal of Structural
Geology, 26, 3, 401-418. London: Reed Elsevier.
Wickham, J. and Moeckel, G. (2002). " Sequential Restoration and Unstraining
of Structural Cross Sections; Applications to Extensional Terranes; Discussion."
AAPG Bulletin, 86, 1, 183-184. Tulsa: American Association of Petroleum
Geologists.
Withjack, M. O., Meisling, K. E. and Reinke, W. J. (1987). "Seismic
Expression of Structural Styles." Abstract 57 th Annual International Meeting, 57-
60 . Tulsa: Society of Exploration Geophysicists.
Woodcock, N. H. and Rickards B. (2003). " Transpressive Duplex and Flower
Structure: Dent Fault System, NW England." Journal of Structural Geology, 25,
12, 1981-1992. London: Reed Elsevier.
Wride, V. C. (1995). "Structural Features and Structural Styles from the Five
Countries Area of the North Sea Central Graben." First Break, 13, 10, 395-407.
Houten: European Association of Geoscientists & Engineers.
Zalan, P. V. (1987). "Identification of Strike-slip Faults in Seismic Sections."
Abstract 57 th Annual Meeting. Houten: European Association of Geoscientists &
Engineers.

ADDITIONAL READING

Allmendinger, R. W., T. A. Hauge, E. C. Hauser, C. J. Potter, S. L.
Kemperer, K. D. Nelson, P. Knuepfer and J. Oliver. 1987. Overview of the
COCORP 40N transect, western U.S.: The fabric of an orogenic belt. GSA Bull.
98: 308.
Badley, N. E., T. Egeberg and C;. Nipen. 1984. Development of rift basins
illustrated by the structural evolution of the Oseberg feature, Block 30/6, offshore
Norway. Jour. Geol. Soc. London 141: 639.
Brown, A. (1999). Interpretation of Three-dimensional Seismic Data. Tulsa:
Society of Exploration Geophysicists.
Chapman, T. J. and G. D. Williams. 1984. Displacement-distance methods in
the analysis of fold-thrust structures and linked-fault systems. Jour. Geol. Soc.
London 141:121.
Cliffs, E. (1985). Principles of Structural Geology. New York: Prentice.
Ellis, P. G. and K. R. McClay. 1988. Listric extensional fault systems-Results of
analogue model experiments. Basin Res. 1(1): 55.
Fangin, S. W. (1991). Seismic modeling of Geologic Structures. Tulsa: Society
of Exploration Geophysicists.
Fox, J. 1987. Seismic interpretation in salt-controlled basins. The Leading Edge
(SEG) 6 (3): 10.
Fulton, T. K. 1985. Some interesting seismic noise. 1985 Offshore Tech. Conf
paper no. 4929, Houston, TX.
Gibbs, A. D. 1984. Structural evolution of extensional basin margins. Jour. Geol.
Soc. London 141: 609.
Jarvis, G. T. 1984. An extensional model of graben subsidence-The first stage of
basin evolution. Sed. Geol. 40:13.
Johnson, J. D. 1992. Structural imaging in the real world. The Leading Edge
(SEG)11 (1): 32.
Jones, P. B. Balanced cross sections-An aid to structural interpretation. The
Leading Edge (SEG) 7 (8): 29.
King, G., and J. Nabelek. 1985. Role of fault bends in the initiation and
termination of earthquake rupture. Science 228: 984.
Mitra, S. and Fisher, G.W (1992). Structural Geology of Fold and Thrust Belts,
Baltimore: Johns Hopkins University Press.
Proffett, J. N., Jr. 1977. Cenozoic geology of the Yerington District, Nevada,
and implications for the nature and origin of Basin-and-Range faulting. GSA Bull.
88: 247.
Roberts, A. N., G. Yielding and B. Freeman, eds. 1991. The Geometry of
Normal Faults. Special Pub. no. 56: Bath, England: Geol. Soc. London Pub.
House.
Schlische, R. W. 1991. Half-graben basin-filling models: New constraints on
continental extensional basin development. Basin Res. 3 (3): 123.
Schumaker, R. C. 1981. Anatomy of a thrust belt as viewed by a petroleum
geologist. Appalachian Petrol. Geol. Symposium, West Virginia Univ.
Sclater, J. G. and M. D. Shoney. 1988. Mid-Jurassic through mid-Cretaceous
extension in the Central Graben of the North Sea-Part 2: Estimates from faulting
observed on a seismic line. Basin Res. 1(4): 201.
Sclater, J. G. and B. Celerier. 1988. Errors in extension measurements from
planar faults observed on seismic reflection lines. Basin Res. 1(4): 217.
Tearpock, D. J. and Bischke, R. (1991). Applied Subsurface Gological
Mapping. New York: Prentice Hall.
Tearpock, D. J. and Bischke, R. (2002). Applied Subsurface Gological Mapping
with Structural Methods. New York: Prentice Hall.
Tucker, P. M. (1973). Pitfalls in Seismic Interpretation. Tulsa: Society of
Exploration Geophysicists.
Watts, A. B. 1982. Tectonic subsidence, flexure and global changes of sea level.
Nature 297: 469.
Woodward, N. B. and Boyer, S. E. (1985). An Outline of Balanced Cross-
Sections. Knoxville: University of Tennessee Department of Geology.

Anda mungkin juga menyukai