Download as pdf or txt
Download as pdf or txt
You are on page 1of 59

THE OPEN UNIVERSITY

PETROLEUM GEOLOGY FOR PRODUCTION STAFF

Manual 3 Reservoir Architecture II: tectonic deformation and structures

Manual 4 Special Topics

©1989 The Open University and Shell Internationale Petroleum Maatschappij BV

i
Produced for the Training Division of Shell Internationale Petroleum Maatschappij by the
Open University Contract Training Unit

Production Team

Andrew Bell (Author) Consultant to the Open University


Brent Cheshire (Author and OU-Shell liaison ) Shell Training
Kevin Gowans (Production Manager) Contract Training Unit, Open University
Frank Jahn (Author and OU-Shell liaison ) Shell Training
Chris Wilson (Author and Project Coordinator) Earth Sciences, Open University

ii
Manual 3 Reservoir Architecture II: tectonic deformation and structures

Acknowledgements

The authors wish to express their gratitude to the following for permission to use
diagrams:

Figures 3.6, 3.7, 3.8, 3.9, 3.10, 3.17, 3.24, 3.25, 3.27, 3.44, 3.45, 3.46 Copyright (c) 1987
by Academic Press (from Ramsay, J. G and Huber, M. I (1987): The Techniques of
Modern Structural Geology, Volume 2: Folds and Fractures)

iii
MANUAL 3

RESERVOIR ARCHITECTURE II: tectonic deformation and structures

Contents

SECTION 1 STRESS AND STRAIN IN ROCKS ............................................................ 1


Study Comment ............................................................................................................. 1
1.1 Theory of stress in rocks.......................................................................................... 1
1.1.1 Force and stress .............................................................................................. 1
1.1.2 Normal and shear stress.................................................................................. 1
1.1.3 Principal stresses ............................................................................................ 2
1.1.4 Mean and deviatoric stress ............................................................................. 2
1.1.5 Graphical representation of stress................................................................... 3
1.1.6 Rock strength and rock failure ....................................................................... 4
1.2 Stress and the production of faults.......................................................................... 6
1.3 Pore fluid pressure and failure ................................................................................ 7
1.4 Theory of strain in rocks......................................................................................... 8
1.5 Relationship between stress and strain ................................................................... 9
1.5.1 Pressure and Temperature ............................................................................ 11
1.5.2 Rock lithology .............................................................................................. 11
1.5.3 Strain rate...................................................................................................... 12
1.6 Summary of stress and strain ................................................................................ 12
SECTION 2 RESERVOIR-FORMING STRUCTURES ................................................ 14
2.1 Faults..................................................................................................................... 15
2.1.1 Description and classification of faults ........................................................ 15
2.1.2 Faults in an extensional setting..................................................................... 19
2.1.3 Growth Faults ............................................................................................... 22
2.1.4 Transfer faults............................................................................................... 24
2.1.5 Faults in a compressional setting.................................................................. 25
2.1.6 Faults in a strike-slip setting......................................................................... 28
2.2 Folds ..................................................................................................................... 31
2.2.1 Description and classification of folds ......................................................... 31
2.2.2 Fold mechanisms .......................................................................................... 34
2.2.3 Folds in an extensional setting ..................................................................... 34
2.2.4 Folds in a compressional setting................................................................... 35
2.2.5 Folds in strike-slip zones .............................................................................. 36
2.3 Diapirism .............................................................................................................. 37
2.4 Examples of structural traps in important hydrocarbon fields.............................. 39
2.4.1 The Brent field - an oilfield with an extensional origin ............................... 39
2.4.2 The Waterton field - a gasfield with a compressional origin........................ 41
2.4.3 The La Paz-Mara field - an oilfield with a strike-slip origin........................ 42
2.4.4 The South Marsh Island field - an oilfield with a diapiric origin ................. 46
2.5 Summary of reservoir-forming structures............................................................. 47

iv
SECTION 3 RESERVOIR-PARTITIONING STRUCTURES....................................... 50
3.1 Fractures ............................................................................................................... 50
3.2 Fault gouges and fault breccias............................................................................. 52
3.3 Clay smearing ....................................................................................................... 52
3.4 Summary of reservoir-partitioning structures ....................................................... 54

Study Comment for Manual 3

We have seen in the overview that hydrocarbons are trapped in a closed,


three-dimensional "container" of reservoir rock. Usually, reservoir rocks have been
deposited in layers that were originally close to horizontal; hydrocarbons are only
trapped in the reservoir if it has been tilted, folded or faulted some way. If the
reservoir has been offset by a fault or affected by close-spaced fractures, we need to
know this before we design a production programme for it. The study of how rocks
deform, by either failure or by bending and buckling, is therefore of great
importance to the petroleum industry.

This section introduces rock deformation in three sections:

⋅ theory of stress and strain


⋅ reservoir-forming structures
⋅ reservoir-partitioning structures

v
SECTION 1

STRESS AND STRAIN IN ROCKS

Study Comment

Engineers are already familiar with the basic concepts of stress and strain related to
engineering applications. Geological engineers need to approach stress and strain in
an unfamiliar way. This is not only because of our poor understanding of the link
between stress and strain in Nature, but also our inability to observe and measure
processes happening at very slow geological strain rates. Nevertheless, all reservoir
rocks show strain in some measure, whether it is directly stress-related like rock
fractures, or indirectly stress-related like folds within the reservoir beds.
Understanding how these folds and fractures occur and where they may be expected
to form is vital to the successful exploitation of the reservoir. The first part of this
section leads us towards why, how and where fractures form within reservoirs. The
second part discusses the natural conditions under which reservoir rocks deform.

1.1 Theory of stress in rocks

1.1.1 Force and stress

Rocks, like other materials, change shape when they are acted on by a force. The force of
gravity affects all rocks and acts at right angles to the Earth's surface. Other forces are
imposed on rocks which change either their amount or direction of action with time. Two
important examples of geological forces we have already discussed are the motion of the
tectonic plates and the changing effects of heat flow from deep within the Earth's crust.

The effect of a force depends on the area over which it acts. For example, you can press a
thumbtack into a wall but you probably can't press a coin into the same place. The applied
force is the same but the area over which it acts is different. To find out how effective a
force is we need to remove the variable of area, which we do by considering forces acting
over a unit area.

The effective force per unit area is called stress. Stress has the units M.L-1.T-2 (N.m-2) and
is measured in pascals. Since this is a small unit for geological purposes, you will
frequently see stress measured in bars or kilobars; 1 bar = 105 pascals.

1.1.2 Normal and shear stress

The force acting on a surface does not necessarily act at right-angles but may be inclined
at some angle to the surface. If the surface is not planar, the direction of action of the
force will change as the surface curves and its effect will vary from place to place.

1
To understand this variability we can resolve the force into two components, one
component acting normal to the surface at any point and the other acting along the surface
at that point (Figure 3.1). The normal component presses down on the surface and is
called the normal stress, σ; the parallel component causes sliding along the surface and is
called the shear stress, τ. Within reservoir rocks both normal and shear stresses vary
continuously from place to place and with time.
FIGURE 3.1
Resolution of stresses
into normal stress, σ
and shear stress, τ.

Increasing shear stress relative to normal stress will eventually produce failure planes in
reservoir rocks.

1.1.3 Principal stresses

Whilst normal and shear stresses vary across any surface, it is clear that the value of the
normal stress will be at a maximum when the force acts directly at right-angles to the
surface. At this point the shear stress will be zero.

If we consider the effects of some force acting on a point within a three-dimensional rock
body, there will in general be three directions perpendicular to each other, which record
zero shear stresses. These directions are known as the principal axes of stress and
conventionally they are labeled σ1, σ2 and σ3 for the greatest, the intermediate and the
least tensile stress respectively. σ3 is therefore the direction in which the rock is trying to
stretch most.

1.1.4 Mean and deviatoric stress

If we know the three principal values of normal stress that are acting at a point, we can
work out the average, or mean stress. This will be given by:

Mean σ = (σ1 + σ2 + σ3)/3

Normally, the mean stress value will not equal the intermediate principal stress value. We
can also calculate how each of the principal stresses differs from the mean stress. This
difference is called the deviatoric stress and is given by:

Deviatoric σ = σ - (σ1 + σ2 + σ3)/3

for any of the principal stress directions.

Within reservoirs, changes in mean stress cause changes in volume and therefore porosity
of the rock. Changes in deviatoric stress control bending, buckling and breaking of the
rock layers.

2
1.1.5 Graphical representation of stress

The principal normal stresses act at right-angles to each other and have different
magnitudes. A convenient way of illustrating this situation is by drawing a stress
ellipsoid with axes oriented in the principal stress directions and axial lengths
proportional to the stress amounts (Figure 3.2). This representation is particularly useful
for visualising mean and deviatoric stresses. Mean stress is always represented by a
sphere in which the length of the radius shows the amount of mean stress. The orientation
of the mean stress axes is not significant. Deviatoric stress is represented by an ellipsoid
in which the long axis is always greater and the short axis always less than the mean stress
sphere radius (Figure 3.3).
FIGURE 3.2 (left)
Stress ellipsoid with
axial lengths σ1, σ2, σ3.

FIGURE 3.3 (right)


Mean and deviatoric stress
(two-dimensional case).

The stress ellipsoid is not an effective way of illustrating total stresses since it cannot
show shear stresses. Consideration of shear stress is critical to our understanding of how
rock fails, so we need to use a more complete form of graph, which can show both normal
stress and shear stress quantitatively. We use a Mohr diagram, which is simply a graph
of shear stress against normal stress (Figure 3.4). Tensile normal stresses are plotted as
positive values and compressive normal stresses as negative values on the horizontal axis.
Shear stresses are plotted on the vertical axis. The mathematical justification for the Mohr
diagram will not be given here, but the equations relating shear to normal stress are of the
mathematical form of a circle.

FIGURE 3.4 A Mohr


diagram (two-
dimensional case).

To examine the stresses at some point, we construct a circle centred on the σ axis, which
just touches the σ1 and σ2 points. The amount of shear stress along planes at any given
orientation to the principal stress planes can then be directly read from the graph. The
angle subtended in the centre of the circle is a "double angle", so the angle between the
plane of interest and the principal stress orientation is half that subtended at the centre of
the Mohr circle.

3
An observation that can be made directly from the Mohr diagram is that the maximum
shear stress occurs along planes lying at forty-five degrees (half the double angle) to the
principal stress orientations.

When stress is considered in three-dimensions, the Mohr diagram is always of the form of
three circles passing through pairs of σ1, σ2 and σ3 (Figure 3.5). The greatest shear stresses
in this system lie at forty-five degrees to the σ1-σ3 plane.

FIGURE 3.5 Mohr


diagram for three-
dimensional case.

On the Mohr diagram, mean stress always plots at a point on the σ-axis, whilst deviatoric
stress plots as a set of circles centred on the σ-axis.

1.1.6 Rock strength and rock failure

We define rock strength as the stress difference (σ1-σ3) the rock can withstand before
breaking. Figure 3.5 shows that the greater this normal stress difference is, the greater will
be the maximum shear stress the rock experiences. Strong rocks can withstand more shear
stress before failing than weak rocks.

The existence of failure planes in all reservoir rocks exposed at the Earth's surface or seen
in core, tells us that natural stress conditions, which exceed the strength of these rocks,
must exist or have once existed. However, measuring stresses deep in the Earth without
altering the stress balance at depth is not generally possible. Therefore, we use
experiments to obtain most of our data about rock strength and the stress conditions under
which rocks fail.

In a typical experiment, cylinders of rock with no existing fractures are placed in a


triaxial test rig, which can produce uniform confining pressures (equivalent to mean
stress) and also impose a known load on one end of the specimen. In modern rigs both
temperature and rate of application of the load can also be altered. As the specimen is
loaded, the stress difference σ1-σ3 increases until at some particular load the rock strength
is exceeded and the specimen breaks. The rig is then dismantled and the orientation of the
failure plane relative to tile load direction is measured. The known values of σ1 and σ3
allow a Mohr diagram to be constructed and, by calculating the failure plane double
angle, a point in σ-τ space can be found which represents the rock strength under the
given test conditions. This point lies on the Mohr circle at the intersection of the circle
with a radius oriented at twice the failure angle from the σ axis.

4
This experiment can be repeated, using fresh samples, for a range of normal stress values
and the complete results plotted on a σ-τ graph. Figure 3.6 shows the results of just such
an experiment for one Jurassic limestone; these tests were performed at ETH Zurich. This
diagram shows five calculated points, P1 to P5, which together define a failure envelope
that separates stresses, which produce rock failure from those which do not. You can see
from Figure 3.6 that the failure envelope rarely coincides exactly with the maximum shear
stress for any set of experimental conditions.
FIGURE 3.6 Mohr
circle and Mohr
envelope.

The simplest envelope is composed of two straight lines tangential to the suite of circles
(Figure 3.7(a)) which is known as the Coulomb failure criterion. The parabolic failure
envelope shown in Figure 3.6 can be predicted from the theoretical considerations of the
propagation of numerous tiny cracks that exist in all natural rocks. These cracks are often
called Griffith cracks after their discoverer, and the envelope is commonly called the
parabolic or Griffith failure criterion (Figure 3.7(b)).

FIGURE 3.7 (a) The


classical Coulomb
criterion for failure.
(b) The parabolic Mohr
envelope predicted from
Griffith crack theory.

5
1.2 Stress and the production of faults

Three observations follow from the nature of the failure envelopes just described that are
very important in petroleum geology.

Firstly, rock fails along planes, which lie at an angle of 45 degrees or less to the maximum
compressive stress. The symmetry of the Mohr diagram tells us that there will be two
failure planes, intersecting along the σ2 direction, and lying at equal angles from σ3.
Within reservoir rocks these conjugate planes may be oriented at any angle within the
crust, depending on the orientation of σ3 at the time of failure.

Reservoir rocks that have been subjected to stress will contain multiple sets of fractures if
the superimposed stress has exceeded the rock strength. Such stress may have been
induced by gravity loading by sediments deposited above them, or deviatoric stresses due
to plate motions stressing the crust. Hardly any reservoirs exist in which these fracture
sets are not present.

Secondly, the surface of the Earth must be a plane of zero shear stress since objects do not
slide about on it of their own accord. The principal normal stresses must therefore be
oriented at right-angles to and along the Earth's surface. There will therefore be three
basic orientations of conjugate failure planes near to the Earth's surface along which the
rock layers can be expected to break. These directions depend on the relative orientations
of the maximum, intermediate and minimum normal stresses and are shown in Figure 3.8,
together with the sense of movement of the failure planes.

FIGURE 3.8 The three


main types of conjugate
faults expected to
develop close to the
surface of the Earth. (a)
Normal faults. (b) Thrust
faults. (c) Strike-slip
faults.

Using this observation we can make a fundamental classification of rock failure planes, or
faults, which will allow the surface of the Earth to stretch, to shorten or to move
sideways.

When the surface of the Earth stretches (σ3 vertical), extensional or normal faults will
form at a high angle to the ground surface. When the surface of the Earth compresses (σ1
vertical), thrust faults that allow beds to overlap will form at a low angle to the ground
surface. When blocks of the crust slide past each other (σ2 vertical), steep strike-slip
faults in diagonal arrays will be formed.

Different stress regimes generate different patterns of failure planes. More than one type
may be present in your reservoir rocks if they have undergone a complex stress history.

6
Thirdly, the angle between the failure planes decreases as σ1 becomes more tensile. This is
clearly shown in Figure 3.9 for three Mohr circles, which have increased values of σ1. If
the crust is in active extension it is likely that the resulting extensional faults will form at
a high angle to the Earth's surface.

FIGURE 3.9
Relationship between the
angle θ and the differing
stress states defining a
curved Mohr envelope.

Mohr

Extensional normal faults are therefore usually steep faults when they form near the
surface. Flat-lying reservoirs found in extensional areas will develop sets of vertical open
fractures.

1.3 Pore fluid pressure and failure

If each of the three principal stresses acting on a point in the rock body is increased by the
same amount, the effect of the deviatoric component will decrease relative to the effect of
the mean component. In the Mohr representation, circles will be displaced to the right of
their original position (Figure 3.10). A consequence of this change is that circles which
did not touch the failure envelope when the effective stress was high may do so when it is
decreased. Increasing each of the principal stresses evenly may therefore lead directly to
failure in rocks.

There is an important application of this feature in production geology. We sometimes


need to increase the fracture density in a reservoir to enable fluid to flow more freely
through it. This can be achieved by isolating the interval downhole and pumping in a fluid
under pressure. If the effective stress is decreased sufficiently the reservoir will fracture. If
the fractures are kept open, the permeability will be increased.

7
FIGURE 3.10 Mohr
circles for a stress
σ1>σ2>σ3, all possible
values of σ and τ for
this state lie in the
stippled area between
the three circles.
By increasing all
stresses by a factor of
σ the stress circles
move towards the
right-hand side of the
diagram so that the
σ1σ3 circle will
eventually touch the
Mohr envelope at P.

1.4 Theory of strain in rocks

When a body of rock is acted on by some force, it must respond by changing either its
shape or its position. The change of place or shape may be temporary so that when the
force is removed it will return back to its original form. In other circumstances the shape
change may be permanent, so that when the force is removed the rock remains displaced
or distorted.

Under stress, the rock must either move without changing its orientation, spin round on
the spot or change its shape. Generally it will have to do all three in some measure, so we
can say the rock has responded to some external force by a combination of translation,
rotation and distortion. This change of shape of a rock body resulting from stress is
called strain.
FIGURE 3.11 Material
deformation.

Figure 3.11 shows six different ways rock may respond to stress. Two distinctly
different styles of deformation are shown. In 2, 4 and 6 the rock specimen has
broken, with some of the broken pieces being rotated or translated. This is termed
brittle deformation. In 1, 3 and 5 the specimen has stretched or squeezed rather
than broken; this is known as ductile or plastic deformation. A third style of

8
deformation exists; the temporary distortion, or elastic definition referred to above.
Elastic deformation of rocks is rarely seen in reservoir rocks, since elastic behaviour in
the past naturally leaves no record within the rocks we see at the present day.

The last point highlights an important difference between the concepts of strain in
geology and engineering. All too often non-elastic strain in engineering situations means
failure, both for the material and for the engineer. In geology, the rock has to fail in either
a brittle or ductile manner before we can become aware of strain.

Strain refers both to changes in length of lines and to changes in angles. An important
definition is the extension, which is the change in length of a line of unit length. It may be
either positive or negative, the response to either an extensional or a compressional set of
forces. Changes in angle and in volume have similar definitions. Strain is a dimensionless
quantity, for example extension is given by (final length - original length)/original length,
and so strain is generally expressed as a percentage, such as % extension or % volume
loss.

1.5 Relationship between stress and strain

The relationship between stress and strain for rocks under a variety of boundary
conditions has been investigated using the experiments described in Section 1.1.6. Very
low differential stresses (less than 5 kbars) with very low confining pressures produce
elastic distortions in which the strain is usually less than 1% and in which there is a linear
relationship between stress and strain (Figure 3.12). Elastic deformation occurs in rocks
when earthquake shock waves pass through them. Release of stress leads to the complete
recovery of the specimen.

FIGURE 3.12 Elastic


stress-strain relationships.

As the differential stress increases, the specimen breaks or distorts abruptly and the linear
relationship between stress and strain is lost (Figure 3.13). The exact nature of the
deformation now depends on the complex interplay of many factors. Figure 3.13 shows
the form of the stress-strain curve when the rock's elastic limit is reached. The rock may
deform in any manner between the extremes shown by the lines C-D or C-E on the
diagram. Failure along discrete planes leads to a sudden drop in the effective stress as the
rock moves along the failure planes. This is brittle deformation and is recorded by the
segment labeled C-D.

9
FIGURE 3.13 Stress-strain
relationships for rocks.

At the other extreme, the rock may distort but not loose its resistance to applied stress.
This is ductile deformation and is recorded by the segment CE. Most rocks behave in a
manner, which is in some way intermediate between these two extremes, in styles which
are shown diagrammatically in Figure 3.14.

FIGURE 3.14 Spectrum


of behaviour illustrating
the transition from
perfectly brittle
behaviour to perfectly
ductile behaviour.
The Figure illustrates
behaviour for typical
compression experiments
and for typical extension
experiments. The shape
of the specimen is
indicated along the
manner in which it
deforms and the shape of
the stress-strain curve.

When we study deformed reservoir rocks we commonly see changes of shape of the rock
near discrete planes of failure. This tells us that most natural rock deformation is a
combination of both brittle and ductile behaviour.

There are four variables, which profoundly affect the response of rocks to stress. These
are pressure, temperature, rock type and strain rate.

10
1.5.1 Pressure and Temperature

These variables are linked through the depth of burial of the rock. The confining pressure
is an effect of the weight of the rock above the reservoir. Average sedimentary rock
densities vary between 2.0 and 2.5 gms/cc, so rocks buried by two or three kilometers of
overburden have a confining pressure of 500 to 800 bars. Increased confining pressure
leads to more ductile deformation in rocks (Figure 3.15).

FIGURE 3.15
Qualitative display of
structural levels in the
crust.
Only 1 and 2 are
important to the
Petroleum Geologist.

Temperature increases with depth, at the rate of about 30°C per kilometer. Rocks in
sedimentary basins buried under two or three kilometers of sediment are almost one
hundred degrees hotter than at the surface. Increased temperature produces ductile
behaviour at lower deviatoric stress levels. In addition, quartz behaves in a more ductile
manner when wet, so that the presence of water can raise the depth at which ductile
behaviour occurs.

1.5.2 Rock lithology

The type of rock plays a major part in controlling rock deformation. Well-cemented
sandstones and carbonates can withstand the greatest effective stress and when they fail
tend to do so in a brittle manner. By contrast shales, mudrocks and, most noticeably,
evaporitic rocks tend to fail under smaller directed stresses and in a more ductile manner.

11
This feature is known as the competency of rocks. Rigid sandstones tend to control
deformation in a variable sequence of lithologies and are termed competent. Weak shales
and salts are often zones of early flow or failure and are therefore called incompetent
strata.

Reservoir rocks are almost always more competent than source rocks or seals. They are
frequently faulted, and almost always contain many minor fractures. Seals, particularly
evaporites and shales, flow rather than break and contain many fewer fractures. The best
seals behave in an almost completely ductile manner and contain no minor fractures at all.

1.5.3 Strain rate

Strain rate is perhaps the single most important factor that affects the way in which rocks
will deform. Natural rock sequences deform over time periods of millions of years. The
Alps represent a tectonic catastrophy in geological terms in which the crust shortened to
about half its original length, and doubled its thickness, in some 10-20 million years. This
strain rate of about 10-14 per second is high for crustal shortening.

The Viking Graben in the North Sea stretched by a factor of 50% in Upper Jurassic and
Lower Cretaceous times, some 20 million years or less. This represents a strain rate of
around 10-13 to 10-14 per second.

These natural strain rates are impossible to duplicate in experiments, which are
necessarily performed at geologically improbable strain rates. The effect of reducing the
strain rate is to produce early failure and more ductile behaviour.

1.6 Summary of stress and strain

Stress

Stress occurs within rocks as a response to depth of burial or to the motion of the Earth's
plates. All reservoir rocks are therefore continuously subjected to stress and stress
variations. If the difference between the maximum and minimum stresses is sufficient to
cause high shear stress, rocks will break and develop both major and minor planes of
failure.

Most reservoirs have been subjected to high stress levels at some stage in their history, so
reservoirs normally contain many failure planes. These fractures contribute to the rock's
permeability.

Reservoir permeability can be increased downhole by exploiting the natural stress


differential and injecting fluid under pressure into an isolated interval. Stress can lead to
overpressure in some formations.

Different major stress regimes lead to distinctive styles of faulting. Crustal stretching
produces extensional basins, which are dominated by high-angle normal faults. The
shortening of crust in mountain fold belts produces many low-angle thrust faults.
Strike-slip zones, which can occur in both extensional basins and mountain ranges are
characterised by steep faults that form anastomosing sets.

12
Strain

Elastic strains do exist in rocks, but they are rarely met in the production environment.
Most rocks within a hydrocarbon field have suffered permanent strain, either by fracturing
or by distorting internally.

The precise way a reservoir or seal rock strains depends on many physical factors,
including pressure and temperature, the rate of deformation and the nature of the rock
type. In general terms, ductile behaviour occurs at greater depths and with slow
deformation. Since most oil and gas fields are found high in the crust, and at relatively
low temperatures, brittle behaviour is more common.

Rock type is the dominant variable in the deformation of sedimentary rocks. Well-
cemented sandstones and carbonates, which form the major reservoir rocks, deform brittly
and contain many faults and fractures. Thus they may have naturally high permeability.
Basement rocks usually behave like this too, and may be permeable, although they rarely
preserve high porosity. Evaporites and shales form good seals because they behave
ductilely at low temperatures and pressures, and flow rather than fracture. This generally
means they have very low permeability.

13
SECTION 2

RESERVOIR-FORMING STRUCTURES

Our theoretical discussion of stress and strain in the preceding section has introduced the
ways in which a body of rock may respond to stress. The actual pattern of deformation for
any field is extremely hard to analyse in terms of stress fields past and present. This is due
partly to the complex relationships between stress and strain in real rocks, and partly to
our relative lack of understanding of natural deformation mechanisms. Within the
sedimentary basins which host the bulk of the world's hydrocarbons reserves, we are
unable at present to explain the structures we see in terms of the Earth's stress field
varying with time.

However, we do see brittle and ductile structures in onshore exposures and in cores and
cuttings recovered during hydrocarbon exploration and production. We can measure,
record and classify the structures we see on the basis of their most obvious features. When
they are grouped together, a systematic pattern of the way in which rocks respond to stress
emerges, even though we have little idea of the nature of the stress field that caused the
deformation.

Rock structures may be divided into two major groups:

(i) structures that are large enough to control the formation of traps, and
(ii) smaller structures that partition the reservoir.

FIGURE 3.16
Structures of different
scales and sizes.

There is only a difference of scale between the two groups and certainly some
structures both form and partition reservoirs, so the distinction is somewhat
artificial. Natural rock structures grade one into another (Figure 3.16) and this

14
twofold discussion is used purely for ease of description. However, the two groupings are
of interest to different parts of the petroleum industry.
Explorationists are largely concerned with the large structures, which define fields,
whereas production geologists need to pay attention also to the problem of smaller-scale
structures within the field, which can affect the production behaviour of a field.

2.1 Faults

2.1.1 Description and classification of faults

Failure planes within a rock sequence, which show some measurable movement of rock
units, are termed faults. Faults are one of the most important rock structures in petroleum
geology as they are large fractures, which control the distribution of rock units within
fields, and provide both migration pathways and seals for fluids.

Faults in the upper levels of the crust of interest to the oil industry are produced by brittle
failure of rock, and as you might expect from your knowledge of the effect of deep burial
on the stress-strain relationship, they are replaced at depth by a ductile form of
displacement, called a shear zone.

Faults are classified on the basis of their geometric effects (Figure 3.17). Faults that
extend the crust are called normal faults. Faults that shorten the crust are called thrust or
reverse faults. Faults that do neither, but that allow blocks of rock to slide laterally past
each other are called wrench or strike-slip faults. In fact, most faults show components
of strike-slip as well as either normal or reverse movement (collectively called dip-slip).
They are hybrids within this classification scheme.
FIGURE 3.17 The
geometric features of
three main types of
faults found near the
Earth's surface.
(a) Normal fault. (b)
Reverse or thrust
fault. (c) Strike-slip.

Faults are rarely exactly vertical, but usually have a slope or dip (hence dip-slip). The
displaced block above the slope of the fault is known as the hangingwall, the block
beneath the fault is called the footwall. These are old mining terms; one wall overhung
the seam, the other was beneath your feet as you worked.

15
Normal and reverse faults can also be defined in terms of the displacement of the
hangingwall relative to the footwall. In normal faults the hangingwall block is displaced
down the fault plane. The term normal fault is an old miners term, which related to the
fact that in British coalfields they "normally" found a coal seam displaced down the slope
of the fault plane. In reverse faults the sense of displacement is the opposite; the
hangingwall is displaced upwards relative to the footwall. Strike-slip faults show no
component of slip up or down the dip of the fault. The relative movement of the blocks is
laterally, along the strike of the fault plane. They do not normally have hangingwalls and
footwalls. Instead, they are assigned a name by the relative direction of displacement of
the two blocks. You imagine yourself standing on one block with the other moving past
you. If the block you are looking at is moving to the left the fault is termed sinistral or
left-lateral. Movement to the right is called dextral or right-lateral. Left-lateral
displacement is shown in the lower diagram in Figure 3.17 and you can see that this sense
of displacement is true whichever block you are standing on.

When we describe the movement direction of faults it is almost always the relative
displacement of one block compared to the other that we are describing. There are very
few situations in which we know whether the hangingwall or the footwall (or both) have
actually moved.

Relative displacement of the hangingwall block up or down the fault plane moves the
layers of rocks so that they are no longer continuous across the fault plane. This
disruption of the layered sequence produces two important effects:

(i) rocks are moved out of stratigraphic order


(ii) some parts of the sequence may be either missing or duplicated.

For normal faults where the hangingwall has slipped down the fault plane (Figure 3.18),
younger rocks higher up in the sequence in the hangingwall are moved into contact with
significantly older rocks in the footwall. A borehole that penetrates this fault will find
some beds missing near the fault plane compared to the full sequence recorded elsewhere.
For reverse faults where the hangingwall moves up the fault plane, older rocks in the
hangingwall are placed in contact with younger rocks in the footwall and parts of the
sequence are duplicated across the fault (Figure 3.18(b)).

FIGURE 3.18

16
This simple feature of the displacement on faults has the most profound effects in
petroleum geology. If unit 2 in Figure 3.18 is the reservoir sandstone, clearly it is missing
in borehole A and the sequence is said to be tectonically thinned. In borehole B the
reservoir sandstone is penetrated twice (along with some other beds). In this case the
sequence is said to be tectonically thickened.

The displacement of any fault can be determined by matching points on either side of the
fault plane that were once in contact with each other, although of course this movement
can only be specified in relative terms. In Figure 3.19 the points A and A' were originally
in contact so we can calculate the slip vector. The slip on this fault has not been exactly
down the direction of dip (dip-slip); there has also been a small component of movement
in the strike direction (strike-slip). Normally we cannot detect this strike-slip component.
We measure the slip as the movement of the fault within the plane, and we call the
displacements seen in sections across the fault the heave and the throw of the fault for
horizontal and vertical movement respectively (Figure 3.19).

FIGURE 3.19

The difference between the throw and the vertical separation is shown in Figure 3.20.

FIGURE 3.20

17
Figure 3.21 shows the methods of observation that can detect fault throws at different
scales.
FIGURE 3.21 Fault
observation at
different scales.

Frequently the down-displaced side of a fault is referred to as the downthrown side,


whether the fault is normal or reverse. The opposite term, upthrow is also in common
use.

You may come across an old-fashioned term, hade, which is a measure of the angle of the
fault. The hade is the angle α in Figure 3.19, and tan-1 α = heave/throw. The modern term
dip (90°-α) is preferred since it is the term also used for the slope of beds and it is given
by tan-1 throw/heave. In the oil industry the term hade is often used in connection with the
direction in which the fault plane dips. For example, an east-hader describes a fault plane
which dips to the east.

Faults are usually not planar, but curved both in plan and in section. They are often
spoon-shaped and such faults are called listric faults. In sedimentary basins they are
usually concave upwards. Listric faults not only change their dip amount down the surface
of the fault, but also their strike direction along the fault (Figure 3.22). Quite commonly
these faults show more than one curved component to their hade, with the fault tracking
almost horizontally between steep sections that cut down through the rock (Figure 3.23).
The steep sections of the fault surface are called ramps and the gently-dipping sections
are called flats. Fault surfaces that have flats and ramps in them will usually be curved in
strike as well. If the dip of bedding planes in the footwall is gentle, then ramps cut steeply
across bedding, whilst flats are parallel to the bedding. Often incompetent shales or
evaporites act as flats, whilst the fault ramps down through competent sandstones and
limestones.

FIGURE 3.22 (left)


Listric fault.

FIGURE 3.23 (right)


Fault plane with flats
and ramps.

Faults are usually not continuous. Individual faults are always of limited extent.
This means that the displacement must die out along the length of the fault and
reach zero at some definite point, called the fault tip (Figure 3.24). Within a

18
relatively constant regional stress field, displacements that die out along the length of one
fault must be taken up somewhere else on a new fault surface. This situation is commonly
met by production geologists. Many oil and gas fields have margins, which are not single
faults, but arrays of fault surfaces slightly offset along a general trend. As the
displacement dies out on one fault, it is taken up on the next one along strike.
FIGURE 4.24 Principal
geometric features of a
fault plane.
With a blind fault the tip
line does not break
through the ground
surface.

Figure 3.25 shows a set of offset or en-echelon faults, which pass displacement from one
to the next. The overall displacement of the footwall block is the same but the
displacement dies out on any one single fault to be replaced by increasing displacement
on the next. Between two adjacent faults of variable displacement lies a sliver of rock that
is in continuous contact from the footwall into the hangingwall. This is labeled fault
bridge in Figure 3.25.
FIGURE 3.25 (a) Left-
handed en-echelon
group of normal faults
with rotated fault
bridges.
The total displacement
across the three cross-
sections a, b, c is equal,
so there is no different
rotation across the
main blocks on either
side of the en-echelon
zone.
(b) Modification of (a)
The existence of fault bridges is very important in production geology since it shows that where the fault bridges
have been broken by
a reservoir can be in contact across a fault zone, rather than broken and displaced along connecting fault plays.
its entire length.

If fault movement continues, the fault bridges may break and the sliver of rock becomes
an isolated fault block completely surrounded by fault surfaces (Fig 3.25). Alternatively,
the fault array may become interconnected by splays or cross-faults.

It is no longer believed that faults form a continuous plane throughout the rock. They split
and rejoin and isolate small displaced slivers of their footwall. These need to be
recognised by the production geologist because they are not joined to the main reservoir
in a simple way. They are generally avoided during production.

2.1.2 Faults in an extensional setting

Extension is the process by which the crust is stretched, primarily when the crustal
plates move apart from each other. The process is happening at the present-day in the
North Atlantic, where America is moving away from Europe at about 3 cms. per

19
year (about as fast as your fingernails are growing). Some tectonically-active areas of
the present-day crust, like the East African Rift represent an early stage in the process
of continental division prior to separation. The East African Rift is a failed attempt to
split the Horn of Africa from Central Africa and form an ocean between them. It is a
relatively young structure and is characterised by normal faulting on both sides of a
prominent downthrown rift valley or graben.

The same process appears to have happened at different times throughout the Earth's
history. One similar graben, or more correctly a suite of grabens, started to form in
Permian times on the northwestern part of the European continent. This episode of
extension was a precursor to the development of a main centre of plate separation in
the Northern Atlantic. This graben array continued to develop throughout the
Mesozoic and Tertiary to become the site of the present rich North Sea hydrocarbon
province (see Manual 1, Figure 1.1). The Viking Graben in the Northern North Sea
shows many of the typical features of a hydrocarbon-bearing basin formed by
extension. The rift valley itself is defined by parallel sets of normal faults displacing
Jurassic and older sediments downwards towards the centre of the basin. These
faulted rocks, including the important Middle Jurassic reservoirs, were covered by
Cretaceous and Tertiary sediments during a poststretching downwards sag of the
basin floor. These younger sediments are thickest in the centre of the basin. This
region became mature for the generation of hydrocarbons during the later part of the
sag phase as the depth of burial and the temperature increased.

The pattern of faulting followed by sag is a characteristic pattern of many basins. The
initial stretching phase generates normal faults, which increase the overall length of
the pre-existing rocks. During the stretching phase, ductile thinning and stretching in
the lower crust leads to subsidence of the upper layers and elevation of the hot mantle
under the stretched zone. Heat flow increases within the basin during this phase. The
subsequent sag phase that occurs after stretching ceases and is caused by the cooling
of the uplifted mantle. This cooler dense material collapses downwards, allowing
sediments to be deposited in a sag basin above the fault network. The sag basin is
thickest above the centre of the stretched graben, where the lower crust thinned most.
Heatflow decreases as the crust regains thermal equilibrium.

During the stretching phase, structures, which increase the length of pre-existing
rocks, are developed. Normal faults are commonly formed at this time. The same
effect of stretching is achieved whether the faults dip in towards the centre of the
basin or whether they dip in the opposite direction. Both types of structure occur in
all sedimentary basins; those dipping in towards the basin are termed synthetic
faults, the other set dipping in the opposite direction are called antithetic faults.
They often form converging, intersecting sets, which together lower the area between
them, as, can be seen east of the Central Graben in Figure 3.26.

Where antithetics and synthetics diverge, a fault-bounded block is left standing


higher than the surrounding basin. This high block is the structural opposite of a
graben and these features are common within fault-controlled basins. The old name
for this structure is horst, but the modern terms high or ridge are more frequently
seen in reports. The `Central Graben High' in Figure 3.26 is an excellent example.

20
FIGURE 3.26

The fault set forming the Norwegian Basin in Figure 3.26 are shown as listric faults which
join together at depth when they meet a large, laterally-extensive flat or detachment.
Exactly whether faults join these large detachments at depth or whether they keep their
planar nature deep into the crust is the subject of much current debate. Figure 3.27 shows
an alternative model, in which blocks of the rotate bodily as a response to stretching. This
is known as domino or bookshelf faulting.

FIGURE 3.27 (a)


Development of a
half graben structure
on the principle of
the domino- or
bookshelf model.
The model poses a
number of
compatibility
problems indicated
by question marks.
(b) This shows filling
of upper block spaces
by sediments and
volcanic rocks and
the possibility of
magma intrusion in
the lower block
spaces.

Whatever the exact nature of the deep structure of grabens is, the upper parts are
frequently large, planar faults, which separate tilted fault blocks. Many of the biggest oil
and gas fields in the North Sea are trapped in such tilted fault blocks.

At depth, below the zone of brittle failure, the lower crust is hot and under considerable
confining pressure, so the crystalline rocks found at these depths respond to crustal
stretching by ductile flow and distortion rather than by faulting. Materials thin rather than
break when they stretch in a ductile manner. If the lower parts of the crust thin under the
centre of a graben, then the upper surface of the cooler, brittle part of the crust must
respond by subsiding. This means that areas of the crust that are stretching are also areas
of subsidence and therefore likely areas for the deposition of new sediments. Therefore,
sediments will be preferentially deposited in the areas of graben formation (Figure 3.28).

21
FIGURE 3.28

When the lower crust thins, the base of the crust also rises, allowing the relatively hot
mantle to become nearer to the Earth's surface. The result is an increase in heat flow
within the graben together with an associated buoyant effect caused by the reduction in
density of hot rocks. So extending basins are areas of high heat flow.

If the rate of stretching is fast, and the buoyant effect of increase heat flow is significant,
thermal buoyancy may cancel out the effects of subsidence due to crustal thinning. In this
case, the main period of sediment accumulation will significantly post-date the stretching
episode, and the tilted fault blocks may be covered by only a thin layer of
contemporaneous sediment. It is even possible that the upper layers of the crust may rise,
leading to erosion of the highest corners of the fault blocks. Many of the largest of the
North Sea fields show tilted fault blocks with eroded upper corners.

During the sag phase, the upper layers of the hot mantle cool because of their relatively
high position, and the heat flow diminishes. The cool, dense mantle rocks will then sink
under the effects of gravity, leading to a lowering of the top layers of the crust above the
basin. These will now be covered with new sediments in which faulting is not a
significant factor. This process is often called thermal collapse and it is also well
illustrated by the North Sea Viking Graben.

2.1.3 Growth Faults

In the majority of cases, the formation and evolution of a sedimentary basin


presupposes extensional tectonics before and during as well as after sedimentation.
Consequently, we cannot rule out the possibility of faulting and deposition going on

22
at the same time. Indeed, it is inevitable, provided there is a supply of sediment to an
extending basin. Faulting may then have a marked effect on both the thickness of
accumulated sediments and also on the sedimentary facies.

Figure 3.29 shows an excellent example of thickness variations related to a listric fan with
associated roll-over and antithetic faults in a seismic section. Across the fault system all
three units suddenly increase in thickness, suggesting that deposition was more rapid over
what is now the hanging wall of the fault. This indicates that the listric fault system was
active throughout the deposition of units 2, 3, and 4. Unit 3 thins as it passes over the
roll-over and then thickens again in the hanging-wall syncline. This has two possible
implications. One is that deposition was slower over the roll-over than on its flanks,
which is to be expected, as subsidence would have been less over the roll-over. The other
is that the crest of the roll-over was eroded between the deposition of unit 3 and that of
unit 4, so that the common boundary is a minor unconformity. The first is more likely, as
there is no angular discordance and individual reflecting components of unit 3 seem to
thin towards the roll-over. From SE to NW the synthetic faults affect progressively
younger strata, indicating that faulting into the footwall migrated towards the NW with
time. As the earlier faults were cut by a later fault, they were locked, ceased to move and
were carried passively above the younger faults. The antithetic faults seem to be in groups
associated with each synthetic fault and each group affects younger strata to the NW. The
hanging-wall anticline has its crest progressively further to the SE as it is traced down in
the section. Again this implies that listric synthetic faults have migrated to the NW and
have progressively cut back into the footwall of the system.

FIGURE 3.29 Seismic


setion with interpretations.

Faults that were demonstrably active during sedimentation, which affected rates of
sedimentation and which may themselves have been partly controlled by sedimentation
loading are known as growth faults.

23
FIGURE 3.30 Cross-
sections showing the
effect on listric growth
faults of varying rates
of deposition relative
to rates of subsidence.
(a) Deposition is
greater than
subsidence; (b)
deposition =
subsidence; (c)
deposition is less than
subsidence.

The precise effects on cross-sections caused by growth faulting are determined by both
the rate of subsidence during faulting and the supply of sediment, which controls the rate
of deposition. Figure 3.30 summarizes the interplay between fans over a growth-fault
system. Note that the loading effect of accumulated sediments interplays with the position
and sequence of faulting. In Figure 3.30(a) sedimentation is faster than subsidence, and
causes synthetic faults to migrate into the hanging wall. The footwall becomes overloaded
with sediment relative to positions further into the hanging wall. Eventually, the first fault
cannot move because of the loading and a second synthetic fault must develop in the
hanging wall, and so on. Where both factors are the same (Figure 3.30(b)), there is no
differential loading over either wall, and a single synthetic fault can continue to grow,
other factors being constant. Where deposition is slower than subsidence (Figure 3.30(c))
the footwall is unloaded by extension, so that synthetic faults progressively cut back into
it. It is important to remember that the process of faulting itself may accelerate or retard
the supply of sediment, so that the result is a combined interplay between tectonic and
sedimentological controls.

2.1.4 Transfer faults

So far, we have considered extensional faulting in a two-dimensional plane parallel to the


direction of extension. Complex though some of these possibilities are, to approximate
reality we have to consider the dimension normal to the extension direction as well. Fault
lines and their extensions do not remain constant to infinity. If we consider the landslip
analogy for a listric fault, we know that the slipped mass has an approximately
spoon-shaped slip surface. A cross-section parallel to the main displacement looks very
like a listric fault. However, at the flanks of the slip, displacement is not truly normal.
There is a horizontal component of movement, too, giving an oblique-slip sense of
movement. Similarly in extensional tectonics, the faulted masses are often bounded by
such geometrically necessary oblique-slip faults, known as transfer faults. They are part
and parcel of the natural means of avoiding space problems in a non-uniform crust, or in a
set of normal faults with varying slip rates (Figure 3.31).

24
FIGURE 3.31 The
relationship of an
oblique transfer fault to
a listric normal fault.

Transfer faults may separate blocks with very different modes of extensional faulting
(Figure 3.32). This means that different crustal regimes may be juxtaposed across transfer
faults, such as horsts against grabens, or faults with different slip directions. An
interesting feature of transfer faults, which is not always clear in zones of complex
extensional tectonics, is that they are analogous to transform faults. Across a sinistral
transfer fault, related normal faults are stepped to the right.

2.1.5 Faults in a compressional setting

In some parts of the world hydrocarbons are found in areas of compressional rather than
extensional tectonics. Mostly these areas were originally sedimentary basins, which
received sediments into fault-controlled basins in the manner described above.
Subsequently the stress field changed in that part of the crust and the basin started to
shorten rather than to extend. Faults that are characteristic of compressive regimes must
therefore shorten the crust; they will be reverse faults or thrusts.

Thrusts are generally low angle structures, which dip at thirty degrees or less. We saw in
Section 1.2 that when the principal compressive stress is almost horizontal, conjugate
faults forming at forty-five degrees or less to this stress direction will develop. Thrusts
therefore have low dips. They also repeat parts of the rock sequence, and stack these on
top of each other. Thrust zones or thrust belts are therefore areas of the Earth's crust
dominated by low angle faults and multiple stacks of rock stratigraphy (Figure 3.32).
FIGURE 3.32 Cross-
section showing the
Lewis Thrust and
Waterton and Pincher
Creek gas fields.

25
Because thrusts thicken the crust on a regional scale, the crust tends to become depressed
by loading when slices of rock of similar composition are placed one on another. This
thickened crust is lighter than normal crust nearby and rises, much as cork when pushed
under water and then released. The thickened crust will continue to rise until all the
"excess" light crust is removed by erosion. Thrust belts are therefore usually mountainous
areas of high relief and active erosion.

Figure 3.33 shows how a thrust fault develops. They are brittle structures so they tend to
originate as low-angle ramps in competent sandstones or carbonates. At this early stage
there is generally little displacement on these ramps, and the incompetent shales within
the sequence undergo ductile rather than brittle strain. With increasing deformation, the
ramps in successive units join together by failure along bedding planes in the less
competent units. The thrust fault becomes continuous, allowing the hangingwall to move
over the footwall.
FIGURE 3.33 Initiation
of a thrust fault.

As the hangingwall block moves over the shaped footwall, the beds contained within it
must distort to avoid gaps developing above the fault plane. Fault-bend folds are
generated by this movement of the hangingwall over ramps and flats (Figure 3.34) and it
is these upfolds in the rock layers which trap hydrocarbons in thrust belts.

26
FIGURE 3.34

As thrusting progresses, the extra loading of the footwall caused by the weight of the
hangingwall will increase the stress within the footwall, which might then itself fail. The
process of thrust-fault initiation will then be repeated within the footwall underneath the
thrust block. A new thrust fault develops and part of the original footwall gets added to
the hangingwall block (Figure 3.35). This process of footwall collapse generates a zone
of thrust-bounded rock slices immediately underneath the original thrust ramp. The zone
is called a duplex (an American two-story house!). Commonly the elements within the
duplex stack on top of each other rather than just lying next to each other. As they
develop these structures become complex upfolds with many repetitions of the rock
sequence (Figure 3.36) and if they contain good reservoirs they can form multiple-stacked
traps and very exciting prospects.

FIGURE 3.35 Stages


in the development
of a thrust duplex by
progressive thrusting
of a footwall.
This leads to a piggy-
back of thrust units
1-4. The folds
associated with
thrusting at ramps
are carried intact.

FIGURE 3.36 Stages


in the development
of a thrust duplex by
progressive thrusting
in the hanging wall.
This leads to an
overlapping
sequence of thrust
units 1-4. Folds
associated with each
thrust are disrupted
by younger, higher
thrusts.

27
Just as antithetic faults develop within grabens, so backthrusts develop within thrust
zones. They can achieve the same degree of shortening as the thrusts discussed above and
are initiated in the same way, as the `companion' failure in a conjugate pair shown on the
top diagram in Figure 3.33.

Thrust belts are generated by compressive stresses wherever they occur, and if these
happen to act on an area, which had once been extensional, the rock sequence might
already contain a number of major normal faults. Early extension structures might then be
reactivated and may then show very complex movement patterns. Alternatively, the early
normal faults might be by-passed by new compressional structures.

The whole process of reversing the style of deformation of a basin is called tectonic
inversion (or simply inversion) and most of the hydrocarbon-bearing basins around the
world show some evidence of inversion. It is not necessary to discuss inversion at length
here, but you can perhaps imagine just how difficult it would be to predict or model the
changing stress pattern through an inverting basin and why this approach has not yet been
successful in petroleum geology.

2.1.6 Faults in a strike-slip setting

Many of the Earth's largest and most obvious faults, such as the San Andreas Fault of
California, involve crustal blocks moving laterally past one another. Zones of strike-slip
deformation can produce extremely complex structures. However, at least one element
lends them a basic simplicity - their dips do not change much with depth- and there are
relatively few examples of listric strike-slip faults.

Figure 3.37 shows the direction of maximum compression and extension associated with
dextral shear and the likely trajectories of different types of minor fault. The way to
visualize the effects of varying shear strain is to sketch the associated strain ellipses. As
shear strain increases, so the ellipse becomes more elongate and more rotated towards
parallelism with the direction of shear strain, so that the associated fault trajectories are
rotated, too.

FIGURE 3.37 A strain


ellipse forming in a zone
of dextral shear.

Huge strike-slip faults, such as those in California (Figure 3.38), propagate down to the
asthenosphere, as they involve the jostling of entire lithospheric plates. As you can see
from Figure 3.38, the blocks between these vast faults are very similar to the situation
shown in Figure 3.37-they are zones of more or less simple shear.

28
In California, the natural situation conforms very well to the model depicted in Figure
3.37. There are thrusts trending roughly ESE-WNW but very few smaller strike-slips
running at right angles across the blocks (Figure 3.38). However, the orientation of these
related structures depends very much on the amount of shear strain and the shape of the
resulting strain ellipsoid. In the San Francisco area shear strain is variable, rising to a
maximum near to the major faults. This results in curved outcrops of thrusts, which
become asymptotic to the major faults.
FIGURE 3.38 Recently
active thrusts and
strike-slip faults in the
San Francisco area.

Secondary faults in sedimentary basins bounded by active strike-slip faults have long
histories. The earlier fault surfaces are rotated because they have been subject to shear
strains for longer. Consequently, they can take an almost spiral or propeller-shaped form,
leading to exceptionally complex cross-sections across the basins. Moreover, they exhibit
progressively less displacement in younger strata. In this respect they are akin to growth
faults in an extensional environment.

It may seem to be a relief to move from multiple strike-slip systems to those with only a
single major fault of this type. Sadly, this is rarely the case, for such faults are almost
never ideal, vertical, planar surfaces. Consider the situation in Figure 3.39 showing a
strike-slip fault with a kink at A.
FIGURE 3.39 (left)
Kinked strike-slip fault.

FIGURE 3.40 (right)


(a) Transtensional
basin; (b)
transpressional uplift
associated with
kinked strike slip
faults.

29
If the movement along the fault is dextral, the motions of the fault blocks are away from
each other, so giving a tensional effect. If the movement is sinistral, motions are opposed
to give local compression across the fault. Combining tension and compression with the
alternative adjective for strike-slip faults-transcurrent- gives transtensional and
transpressional zones respectively for the region near to the kink at A. In the first case,
the tendency is to develop normal faults at the kink to produce a minor basin or
pull-apart. In the second, older rocks are forced to the surface to form an upthrust block
or pop-up (Figure 3.40). Cross-sections across such zones often take on the form of a
system of dip-slip faults curving into the main strike-slip fault to produce what is known
as a flower structure (Figure 3.41).
FIGURE 3.41 Sections
through flowers in (a)
transtensional and (b)
transpresssional
environments

If the strike-slip faults continue to be active throughout the evolution of a sedimentary


basin, the amount of displacement on transtensional and transpressional faults decreases
upwards, in an analogous fashion to growth faults. They may change their sense of
movement as the zone of active strike-slip movement changes position and trajectory.
Moreover, transtensional faults will inevitably control changes in thickness of strata,
depending on the supply of sediment. Transpressional zones may be marked by local
unconformities, where upthrust zones are quickly eroded.

One final point about strike-slip tectonics; so far, we have considered major faults which
break surface and act throughout the evolution of basins. What happens when strike-slip
movements are relatively small, fail to breach the surface and begin during or after
sedimentation? You can easily get an idea by placing a dishcloth over two bread boards
and then sliding one board past the other. The cloth rucks up into folds. In practice, the
folding forms as an en echelon arrangement of doubly plunging anticlines and synclines
(Figure 3.42). Such fold systems can then be used as guides to the existence at depth of
major strike-slip displacement, and the possibility of complex linear systems of potential
hydrocarbon traps associated with flowers.

FIGURE 3.42 En-


echelon folds over a
buried dextral or
right-lateral strike-
slip fault.

30
2.2 Folds

2.2.1 Description and classification of folds

The most widely-developed and obvious evidence of ductile deformation in rocks is the
production of folds in the layers of bedding. Folds are created when rocks are subjected to
stress but do not break to form faults. As we have seen in Section 1.5, rocks will behave
in this manner deeper in the crust where the pressure and temperature are high, or when
deformation conditions allow the strain rate to be low. In addition, certain types of rock,
like shales or evaporites are more likely to "flow" than break.

Under natural conditions rocks fold into an apparently bewildering set of shapes, or fold
styles, and consequently there is a wealth of terms in use to describe folds and folding.
Figure 3.43 shows a fold pair composed of an anticline or upfold of the rocks, together
with its companion syncline or downfold. The curved beds in the crest and trough of the
fold pair are joined by relatively planar dipping beds forming the fold limbs. The distance
between anticlinal crests is known as the fold wavelength and the height separation of a
bed between crest and trough as the fold amplitude. There is a clear similarity between
the terminology of waveforms and these descriptive terms for folded rocks; most of the
terms in general use for waves have been applied to folds.

FIGURE 3.43 Fold


terminology

If several layers of rock are folded together, as is commonly the case, then lines of
greatest curvature of beds within the folds, or hinge lines, can be defined which stack on
top of each other. The stack of hinge lines defines a planar (or more commonly
curviplanar) surface through the rock that is known as the axial plane or axial surface.
Folds are commonly described by reference to the trend and dip of the axial surface, and
the spacing between adjacent axial surfaces.

In naturally deformed rocks, the hinges are not straight lines but are themselves curved.
The fold style will often look rather like that shown in Figure 3.44, where the crests and
troughs rise up and down giving rise to culminations and depressions in the fold hinge.
Upfolds in the hinge zone of an anticline generate domes, downfolds in a synclinal axis
form basins.

Domes are of fundamental importance in the petroleum industry as traps. You can see
that if the folded bed shown in Figure 3.44 were a porous reservoir and the layer above
were an impermeable seal, the dome would be ideal site for trapping hydrocarbons. A
large fold with a gently curving hinge zone might create a giant oil or gas field. The
major oilfields in the eastern Persian Gulf are large, open folds of this style, folding a
fractured limestone reservoir with an evaporite seal.

31
FIGURE 3.44
Structural features
arising from changes in
the hinge line plunges
of folds.

When layers of rocks are folded (a multilayer) it is rare that two adjacent beds have
exactly the same curved shape. Usually the upper surface of one bed is either more curved
or less curved than the bed below it. A useful classification of folds has been developed
based on this difference. In Figure 3.45, five different fold profiles are shown which
include the whole range of possible variations of two folded surfaces. A set of lines have
been drawn onto each profile which are lines joining places on the two beds having the
same inclination, or dip. The constructed lines are known therefore as dip isogons, and
they provided a ready means of discussing the shape variation within the folded layers. If
the isogons converge downwards then the curvature of the inner arc must exceed that of
the outer arc, and we can call these Class 1 folds. If the isogons diverge downwards, the
opposite must be true and the outer arc curvature exceeds that of the inner. These are
Class 3 folds. There is a special case between these two where the curvatures of the two
arcs are the same. These are Class 2 folds and are also called similar folds because of the
similar (in fact identical!) shape of the two adjacent folded layers.
FIGURE 3.45
Classification of
folds based on dip
isogons.
Dip isogons have
been drawn at 10°
intervals from the
lower to the upper
surfaces X and Y.

The true value of this classification scheme becomes apparent when we consider the
thickness of beds around the folded multilayer. We can measure the thickness of a bed
contained between two folded surfaces; it is convenient to measure the orthogonal
thickness, which is the true thickness at right-angles to the bedding surfaces. Since in
nature different beds will be of different thicknesses before folding, we can make an
accurate comparison of the effects of folding on thickness by normalising the thickness,
assuming the bed to be of unit thickness in the fold hinge. Figure 3.46 shows a graph of
this standardised orthogonal thickness against limb dip for the three classes of fold.

32
FIGURE 3.46
Graphical plot of
standardized
orthogonal thickness
t'α plotted against
angle of dip α and
the main types of
fold classes.

Class 1 includes both folds in which the beds thicken into the limb areas and folds in
which the beds thin into the limbs. The diagram highlights another special case within
Class 1 in which orthogonal thickness is maintained around the fold. Such folds are called
parallel folds since the folded layers remain parallel throughout the fold. These parallel
folds have convergent dip isogons but the bed stays at its original thickness during
folding. Class 2 and 3 folds both show considerable thinning in the limbs.

Examination of Figure 3.45 from a point of view of bed thickness shows a further
property of similar folds. Similar folds thin into the limbs when measured orthogonally,
but maintain a constant bed separation measured parallel to the axial plane of the fold. If
the folds are upright the actual thickness of rocks that have been involved in the folding
process will therefore be maintained.

Within our classification scheme we have now highlighted two important types of fold.
Parallel folds keep the true thickness of the beds constant, but the vertical thickness of the
folded multilayer changes from crest to limb. Similar folds maintain constant vertical
thickness but individual beds thicken and thin from crest to limb.

In naturally folded rocks, Mother Nature seems to be striving to both keep the crust at a
constant thickness during folding (by generating similar folds) but also to maintain the
original thickness of beds (by generating parallel folds). The result is often a multilayer in
which the competent sandstones and limestones show Class 1C, nearly parallel folds
whilst the incompetent shales and evaporites show Class 3, nearly similar folds. The
resultant stack approaches conservation of vertical thickness.

The production geologist can expect to find a predictable fold style in fields
throughout folded terrains. Competent rocks which usually form reservoirs will
form parallel folds in which the orthogonal thickness (isopach) is maintained
whilst the vertical thickness (isochore) increases from crest to limb. If the limb dips
are small this may be only a small variation also. Good evaporite or shale seals are

33
incompetent and these horizons will thin in the limbs and thicken in the crests to
compensate for thickness changes within the reservoir. Shale seals will show similar folds
which are thicker in the crests than they are in the limbs. Note that the isochore thickness
of these horizons could be maintained from crest to limb.

2.2.2 Fold mechanisms

Since folds are in some sense a ductile equivalent of faults, we need to consider whether
the basic fault classification scheme has an equivalent in fold style. In section 2.1, faults
were divided into normal, reverse or strike-slip depending on whether the local stress
regime was extensional, compressional or translational. We can examine ductile behavior
within extensional, compressional and translational stress regimes to see what folds are
likely to be produced in each.

2.2.3 Folds in an extensional setting

When a body of rock is stretched it is able to extend either by breaking or by thinning


(Figure 3.47(a)). We have already examined the crustal-scale effects of thinning in
Section 1.4. and seen that rocks deforming in the deep, ductile parts of the crust thin to
produce necks within the more competent layers. In the upper parts of the crust of interest
to the petroleum industry, breaks are more likely to develop in most rocks, but clearly
gaps like that shown in Figure 3.47(a) will not develop; gravity would cause the
overhanging portions of rock to collapse into the newly-formed hole. If the rock maintains
its continuity above a break (usually a fault plane), the layers in the hangingwall must fold
in order to keep the hangingwall in continuous contact with the footwall (Figure 3.47(b)).
These folds are distinctive since they are formed by bending of the hangingwall layers in
response to the shape of the fault. If the major fault has several steps (flats and ramps)
then a series of anticlines and synclines will be formed having shapes that are functions of
the major fault shape (Figure 3.48). Such folds are known as fault-bend folds. Within
extensional regimes, fault-bend folds are the most common ductile structure.
FIGURE 3.47 (left)
Ductile effects of
extension in rocks.

FIGURE 3.48 (right)


Idealised response in
the structure of the
hanging wall of a
normal fault to ramps
and flats in the
footwall.

Anticlines formed in the hangingwalls of major faults are obvious hydrocarbon


traps. They are important in hydrocarbon exploration and production, and have
become known as rollovers since the hangingwall rocks seem to roll over into the
fault plane. Figure 3.49 shows a seismic profile through a major listric fault. The

34
fault plane is clearly imaged and a broad, open anticline can be seen in the hangingwall
above the curved fault. In areas where these structures close they are capable of hosting
major hydrocarbon accumulations.

Figure 3.49 (a)


Seismic profile of a
simple listric fault.
The anticline or
rollover on the
hanging wall can be
seen clearly. The
zone of confused
reflections on the
hanging wall is due
to other faults in the
section.
(b) Line drawing of
the major structural
components in 3.49
(a)

2.2.4 Folds in a compressional setting

Fault-bend folds are also common in areas where thrust faults are the dominant
deformation mechanism. We saw in Figure 3.35 how fault-bend folds can develop in the
hangingwall of a compressional duplex, and of course these folds are important traps in
hydrocarbons-bearing thrust belts, such as the Canadian Rockies. They make better traps
in some respects than extensional fault-bend folds since they are naturally closed on both
the footwall and hangingwall sides (see Figure 3.35) and because they can repeat the
reservoir section, which enables the same reservoir potentially to be hydrocarbon-bearing
at different levels.

35
There is an alternative way of crustal thickening, which does not involve faulting. Ductile
rocks can fold by buckling, in response to pressure applied to the ends of the layer (Figure
3.50). Buckle folds are commonly seen in the crust in areas where major crustal
compression has been active, since the crust has been thickened and heated and ductile
deformation is favoured. If buckle folds are closed and associated with source rocks that
are not already overmature they can form important traps.

FIGURE 3.50 Ductile


effects of compression
in rocks.

There is a maximum limit to the amount of strain that can be accommodated by buckling.
When folds form they usually have gentle limb dips, but with increased amounts of
buckling the fold limbs will rotate. If they rotate so much they become normal to the
compression direction, no further shortening can be achieved by folding. This occurs after
about 36% shortening. If the buckling strain exceeds this amount the rock will deform
internally. Mineral grains within the rock rotate and recrystallise to form a planar structure
known as cleavage.

Generally speaking when the rocks have started to form a cleavage they are no longer
likely to bear hydrocarbons, both because the reservoir porosity will be significantly
decreased and because any associated source rocks will have become overmature. For
this reason the interiors of major mountain belts do not generally contain significant oil
and gas fields, whereas the less deformed but folded margins can be major hydrocarbon
provinces.

2.2.5 Folds in strike-slip zones

These are by far the most complex structural settings as we saw in Section 2.1.6. Both
fault-bend folds and buckle folds occur in these zones, controlled by the compression and
extension directions in the strain ellipsoid (Figure 3.37). Anticlinal buckle folds will form
normal to the maximum compression within the zone, labeled "fold axis" on Figure 3.37.
Fault-bend folds may form either parallel to this direction or at right-angles to it, and be
basically of compressional or extensional origin. Major fold trends will not be aligned
along the shear direction but at angles of up to 45 degrees to it.

Folds form along a belt defined by the underlying shear zone within the basement but they
have their long axes not parallel to the basement shear but at an angle to it. Such folds are
called en echelon folds (see Figure 3.42). In right-lateral or dextral zones the major shear
zone lies to the right or clockwise from the fold axes; in left-lateral zones the major shear
lies to the left or anticlockwise from the fold axes.

36
Figure 3.51 is a depth contour map in feet of the Newport-Inglewood dextral strike- slip
zone in California. This fold zone is associated with the fault structures shown in Figure
3.38. The en echelon folds and their associated thrust and normal faults give rise to a
complex structural geometry, but nevertheless a highly prospective hydrocarbons trend. In
general the large number of trapping structures in strike- slip zones makes them exciting
prospects, but they are difficult to understand and interpret, and relatively poorly
understood at present.
FIGURE 3.51 The
Newport-Inglewood
wrench fault zone,
California.

2.3 Diapirism

There is one further geological process that creates traps, both folds and faults, but which
does not strictly fall into the category of tectonics. It is a process that happens both in
areas that are active tectonically as well as those that are tectonically inactive, and is
almost exclusively associated either with evaporate salts or with overpressured clays.

Salt beds have some unusual properties. They are both weak and light, and despite being
highly soluble in groundwaters, they have a very low natural porosity. Whilst other
sediments compact with increased weight of rock above them by expelling water from
their pore spaces, salt does not. With time, the contrast in density between the light salt
horizons and the more dense sediments above and below becomes accentuated.
Undercompacted clays will have a large amount of water trapped between their grains and
are similarly very light and mobile.

Both evaporates and overpressured clays are incompetent and behave in a highly
ductile manner even at the low temperatures that exist high in the crust. If a
substantial thickness of evaporate or clay occurs within the sedimentary sequence it

37
will eventually become gravitationally unstable because it is light rock overlain by
heavier sediment of "normal" density. Evaporates also flow plastically under the smallest
of stress differences, so the salt sequence will flow upward into any irregularities in its
upper surface. Concentrations of the light lithology will then balloon upwards through the
rock cover, piercing through each bed and eventually rising towards the land surface. A
body of rock that pierces through shallower rocks in the sequence like this is called a
diapir, and the process is known as diapirism.

Figure 3.52 shows the variety of shapes caused by large diapiric structures and the names
given to them. Whilst these structures are collectively known as diapirs, strictly not all of
them are, since not all of them pierce the overlying rock sequence. Pillows are domes that
merely bulge the lower surface of the succeeding beds upwards, creating anticlines in the
cover sequence that are frequently important traps.
FIGURE 3.52 The
main types of large
salt structures.
Structure contours
are in arbitary units.

As pillows grow upwards into diapirs, they generally form a bulbous, overhanging top
above a slender stem. We know from boreholes through the bulb tops, as well as from salt
mining, that these overhangs can have complex, turbulent internal structures that have
been produced during salt flow (Figure 3.53). When the salt flows upwards it exerts a
force on the adjacent sediment layers, bending them upwards and perhaps even
overturning the beds. These beds now dip away from the diapiric dome but they are still
in contact with it at their upper end. The diapir is impermeable, however, so that the salt
or clay contact acts as a seal. If the appropriate reservoir and source facies exist down-dip
in the sediments that surround the diapir (but which originally overlay it), then an ideal
trapping structure is both generated and sealed by the salt movement.

FIGURE 3.53
Section of a diapir
showing overhang.

38
Salt domes are located by irregularities in the upper surface of the salt, or more strictly in
the lower surface of the bed immediately succeeding the salt. If this is a sedimentological
irregularity it may be a point feature, so the succeeding salt dome or diapir will be
focussed on that point and become circular in plan. These domes usually develop a
characteristic radial pattern of faults in the overlying beds, sometimes intersecting circular
faults associated with the upward push of the salt (Figure 3.54). Such accommodation
structures increase the permeability in the beds immediately overlying the dome, and in
certain cases (e.g. North Sea Ekofisk) turn an otherwise impermeable rock into a reservoir
of good permeability.

FIGURE 3.54 Contour


map of the Wilcox sand,
just above the Clay Creek
dome.

In other cases, the irregularity in the bed originally overlying the salt unit is caused by
some tectonic process such as faulting. Fault-generated salt structures generally form
elongate salt walls or rollers aligned along the strike of the fault. These walls and rollers
may subsequently evolve into diapiric walls because of their gravitational instability.
These zones made up of both salt movement and normal, reverse or strike-slip faulting
produce structural nightmares, but often contain many different trapping styles which may
be salt sealed. Many of the gas-bearing structures in the Southern North Sea Gas Basin are
tectonically triggered salt walls.

2.4 Examples of structural traps in important hydrocarbon fields

2.4.1 The Brent field - an oilfield with an extensional origin

The Brent field is a giant North Sea oil field with recoverable oil reserves of 1,750 billion
barrels (280 x 109 cubic metres) situated on the west side of the Viking graben in UK
block 211/29. The structure is a major rotated fault block on one of a series of extensional
faults (Figure 3.55(a)). The fault strikes north-south and downthrows to the east and is
part of one of two prominent structural lineaments, which are both the locus of important
oil and gas fields. The western fault trend includes the Dunlin, Hutton, Murchison and
Ninian fields whilst the eastern trend includes the Brent and Statfjord fields.

39
FIGURE 3.55 (a)
Isometric block
diagram of the
northern North Sea.
(b) Map of Brent
field. (c) Cross-
section of map.

40
Brent has major hydrocarbon reserves in both the Jurassic Statfjord Formation and Brent
Group. The two reservoirs are separated by the impermeable Dunlin Group (Figure
3.55(c)). The structurally high corner of the block has been eroded array by late Jurassic
erosion that accompanied extension, so both reservoirs have been sealed by Cretaceous
muds and marls which lie unconformably over the rotated block (Figure 3.55(b)). The
Brent Group delta deposits contain the most productive reservoirs in the North Sea. The
best reservoirs found in the massive well-sorted sands of the Etive Formation (middle
Brent), which has been interpreted as a barrier sand bar and the Tarbert Formation (upper
Brent) which is a sand unit marking the transgression which finally drowned the Brent
delta.

The Upper Jurassic Kimmeridge Clay Formation is the principal source rock which lies at
oil generation depths over much of the area. Peak oil generation occurred in early Tertiary
times.

Prior to production the Statfjord Formation had a gas cap of 140m and an oil column of
130m. Good reservoir sands have porosities of 10% - 25% and permeabilities of up to
5500mD. The reservoirs are capable of 6500 bbl (1040 m3) per day of 38° API oil with a
GOR of 1550 scf/bbl.

2.4.2 The Waterton field - a gasfield with a compressional origin

The Waterton gas field is one of Canada's largest and is situated on the eastern flanks of
the Canadian Rockies in the Foothills of Alberta. The structure is a series of anticlines
which have formed in the hangingwalls of thrusts that meet in a roof fault to form a
duplex (Figure 3.32). The duplex was generated by the collapse of the foot wall of the
Lewis Thrust during the formation of the Rockies thrust-fold belt in early Tertiary times.

The reservoir rock is porous and permeable Carboniferous carbonates which have been
folded into fault-bend anticlines. The reservoir is sealed by impervious Mesozoic shales,
which are also folded since they were deposited prior to the onset of thrusting. The gas is
sourced from the Palaeozoic carbonates that were deeply buried prior to folding and
faulting by the emplacement of the overlying thrusts.

The Foothills contain a series of major thrusts with folded carbonates in their
hangingwalls, which splay out eastwards into more minor thrusts through the Cretaceous
sequence (Figures 3.32 and 3.56). Each of the major anticlinal folds forms a
hydrocarbon-bearing structure, although as is the case at Waterton there is often no
surface expression of these deep folds.

FIGURE 3.56
Structural cross-section
of Plains, Foothills,
front Ranges and main
ranges.
The Plains-Front
Ranges section is
roughly along the Bow
River Valley, and the
Front Range to Main
Range section is
through Castle
Mountain.

41
2.4.3 The La Paz-Mara field - an oilfield with a strike-slip origin

The La Paz-Mara complex is a giant oilfield situated in the northern part of the Maracaibo
basin in Venezuela. Ultimate recoveries from both the La Paz and Mara structures
together are expected to be in order of 1.8 billion barrels (300 x 106 m3) of oil. Production
is from fractured Cretaceous carbonates and crystalline basement rocks.

FIGURE 3.57 Major


tectonic elements.

The Maracaibo basin is dissected by long faults that are known to be wrench faults
associated with basement dislocations (Figure 3.57). The La Paz-Mara fault zone itself
cuts through the centre of the La Paz and Mara anticlines, and has a left-lateral sense of
displacement of the order of 2.8 kms. The zone is sinuous in outline and consists of two
principal faults bordering a highly complex arrangement of rotated, pushed-up and
slipped-down slices of rock, bounded by faults branching off from or between the
principal faults (Figure 3.58).

42
FIGURE 3.58
Simplified contour
map Top La Paz
Shale Fm.

Sequential horizontal sections (Figures 3.59 and 3.60) suggest that the lateral
displacement of the wrench fault is constant with depth and post-dates Lower Eocene
sedimentation. The structural development can be reconstructed to show the movements
of the main fault zone throughout time (Figure 3.61). The fault zone was initiated as an
extensional fault and reactivated as a strike-slip fault during Lower Eocene times. Figure
3.61 highlights high angle thrust faults coexisting with high angle extension faults, a
feature that is characteristic of wrench zones.

FIGURE 3.59 Mara


Field-horizontal
section at 3000'.

43
FIGURE 3.60 Mara
Field-horizontal
section at 5000'.

FIGURE 3.61
Tectonic history of
the La Paz-Mara
Fields.

The nearby Icotea fault zone (Figure 3.57 and Figures 3.62 and 3.63) shows further
typical features of wrench zones. The accumulations on the west flank of the main
fault in Lake Block 1 occur in a series of en echelon anticlinal flexures in Lower
Eocene clastic reservoirs. The Main Icotea fault itself is almost perfectly straight
and dips steeply at about seventy degrees down to 10000 feet. Below this depth
seismic evidence indicates that the fault plane becomes nearly vertical. Throw

44
reversals along strike-slip faults are common; along the Icotea fault the apparent vertical
throw of 3000ft westwards in Lake Block 1 reverses to become an easterly throw in the
concession of Sun Oil Co. to the south.

FIGURE 3.62

45
FIHURE 3.63

2.4.4 The South Marsh Island field - an oilfield with a diapiric origin

The South Marsh Island 130 field is located 90 miles offshore from Louisiana, USA in
215 feet of water.

The structure, shown in Figures 3.64 and 3.65, is a highly faulted piercement salt dome.
The salt dome itself has been emplaced by the effects of gravity into the footwall of two
major extension faults. Fault F1 on Figure 3.64 throws down to the northwest and west
and is itself cut by fault F-30 downthrowing to the southeast. The major faults, together
with the many smaller ones shown on both the map and section, allow the rock sequence
above the salt dome to stretch to accommodate upward salt emplacement.

FIGURE 3.64

46
FIGURE 3.65 Line of
section - South Marsh
130 Field.

The sediments have been deposited in an area of active growth faulting, as can be seen
both from the increased displacement of lower sandstones, (compare the displacement of
E4 with that of I2 above the salt dome) and the additional horizons present in the
hangingwall of the major easterly-dipping fault (E2, E10, F, H and others).

High geopressures occur as shallow as 3000ft. subsea near the crest of the structure, but
most of the 300 gas and oil accumulations that have been penetrated occur in
normally-pressured Pleistocene sands. These range in depth from 580ft. to 7000ft. subsea.

The reservoir sands are interpreted to be of deltaic origin and typically have porosity
values in excess of 30% and water saturations of less than 15%. Typical dip of bedding
ranges from 10° to 30° at the productive horizons. Many of the objective sands occur as
distinct stratigraphic events separated by relatively thick shale intervals.

Oil columns in the range 300 - 900 ft. are common in the field. The largest oil
accumulations occur on the downthrown side of major faults. Sealing faults that have
throws ranging from 40 ft. to 2000 ft. are the predominant trapping mechanism. Most oil
accumulations have associated gas caps.

2.5 Summary of reservoir-forming structures

Faults

Faults are brittle failure planes within rocks, which show measurable displacements from
footwall to hangingwall block. Normal faults allow the crust to stretch and are associated
with stratigraphic omission whereas thrust faults allow the crust to shorten and are
associated with stratigraphic repetitions. Wrench or strike-slip faults allow blocks of the
crust to move past each other.
47
Faults are rarely planar, but more usually curved in plan and cross-section, and track
through the rock sequence in a series of ramps and flats. They are rarely continuous,
more usually they often form an overlapping set in which displacement dies out on
one to be picked up on its neighbour.

Faults are important in production geology both as margins of fields and also as
either migration pathways for hydrocarbons or up-dip seals of reservoirs. Faults
within the reservoir have important effects in controlling fluid flow through the
reservoir.

Folds form as anticlinal upfolds and synclinal downfolds in response to the ductile
deformation of rocks. Rises and falls of their axes form basins and domes in map
view. Significant rises along the length of anticlinal axes form structurally closed
domes which are probably the most common trapping structures.

Folds are produced by two processes, either as fault-bend folds in the hangingwalls
of stepped normal or reverse faults or as buckle folds by regional compression along
the layers. Both types of structure can trap hydrocarbons. Buckle folds, however,
more usually characterise extensive compressional deformation that may also have
destroyed reservoir porosity and overheated source rocks.

Extensional structures

Areas where the crust has stretched regionally are characterised by normal faults,
crustal subsidence with associated sediment deposition, and high heat flow.

Normal faults may be planar or listric; sets of these define graben lows and block
highs. Sedimentation is preferentially concentrated in grabens and halfgrabens.
Crustal extension leads to rotation of rigid blocks, which form important traps if they
are sealed up-dip. Folds in these areas are usually low amplitude, long wavelength
structures associated with shaped faults or compaction of sediment.

Sedimentation during tectonic activity produces growth faults which form major
hydrocarbon plays world-wide. Extension fault blocks are bounded both regionally
and locally by strike-slip faults which form important partitions between and within
fields.

Subsidence after stretching produces a thick, undeformed cover sequence and leads
to thermal maturity of source rocks.

Compressional structures

Areas where the crust has shortened regionally are characterised by thrust faults,
folded beds, uplift and erosion associated with high mountain chains, and low heat
flow.

Thrust faults are low angle brittle structures, which often show large displacements.
There are usually complex arrays of thrusts and duplexes, particularly in the
footwalls of major thrusts. Both fault bend and buckle folds are intimately associated
with thrusts, and anticlinal domes form the most important trapping structures within
thrust-fold belts. Folds may be tight and have overturned limbs, in contrast to the
open folds of extensional terrains.

48
Thrust complexes are also bounded both regionally and locally by strike-slip faults which
form by crustal flexure.

Uplift during and after compression leads to a major period of erosion. Source rock
maturity is usually caused by burial when major thrust sheets are emplaced.

Strike-slip structures

Areas where blocks of crust are moving past each other on a regional scale are
characterized by steep strike-slip faults, en echelon folds and faults, and rapid alternations
of sedimentation or erosion.

Wrench zones are always areas of highly complex structures in which both folds and
faults are common. Faults can be either normal or reverse and frequently change from one
to the other along their length. Thrusts and normal faults tend to form at high angles to the
main zone, whilst strike-slip faults tend to parallel the zone. The main faults are
characteristically steep to vertical at depth and splay outwards at shallower angles near the
surface. En echelon fold arrays lie along the zone but individual fold axes usually make
angles of twenty to forty degrees to the main structural trend.

Patterns of sedimentation and source rock maturity vary rapidly both along the zone and
with time at any point.

Diapirism

Diapirs are generated by the gravitational instability caused when light rocks lie beneath
heavier ones. The light rocks are most commonly salt but overpressured shales can also
form diapiric structures. The upward movement of the light rocks bulges the overlying
sediment upwards into an anticlinal dome, and this is the most common trapping structure
associated with diapirs. Bending of the cover rocks above diapirs and salt pillows
increases their permeability.

Diapirs can also pierce through the sequence above them, tilting and sealing the sediments
and forming important traps.

49
SECTION 3

RESERVOIR-PARTITIONING STRUCTURES

The structures we have described so far, in particular faults and folds, occur in
rocks everywhere throughout the world where tectonic conditions are appropriate
for their formation. We have discussed them under the heading of reservoir-
forming structures because they are the major rock structures, and in general all
structural traps are either fault blocks or folds, or perhaps some combination of the
two.

It is important that we recognise that structures exist on all scales (look back at
Figure 3.16). If we consider faults, then displacements vary from almost nothing up
to many kilometers in major strike-slip zones. Usually, faults with a small
displacement are of small areal extent, whilst those with a large displacement are
large fault planes, often forming the boundary to several oil fields. Folds too can be
major structures with wavelengths of many kilometers, as in the Zagros belt of Iran,
through to the smallest of folds seen in some metamorphic rocks.

So, to divide structures into "reservoir-forming structures" and


"reservoir-partitioning structures" leads to a somewhat arbitrary subdivision. Many
large fault blocks which are important fields have smaller faults passing through
them which do not form the structure but which partition it. Similarly, minor folds
on the crest of a major anticline are not considered to be part of the field structure,
but rather an annoying extra which is likely to lead to production difficulties at
some later stage.

For the production of a field, small-scale or minor structures are of great


significance, because they may displace the reservoir sufficiently to form a fluid
barrier, or they may produce separate pools of hydrocarbons which will require a
carefully designed drilling programme to exploit fully.

In geological terms, though, these structures are no different from their larger
counterparts that we have already studied. The general comments about, say,
normal faults apply to small-scale normal faults as well as larger, field-forming
ones. In addition to the large and small scale faults and folds, there are two types of
structure which occur in all fields and which always have profound effects on the
production strategy, since they may or may not form sealed zones within the
reservoir. These two structures are fractures and fault gouges.

3.1 Fractures

Fractures are brittle failure planes which are not faults because they appear to show
no displacement of beds whatever from one side to the other. Fractures occur
everywhere on surfaces that are exposed at the present day, where they are called
joints, but they become less frequent at depth. They characteristically affect
competent rocks rather than incompetent ones, and in sections of mixed
competency, say limestones and shales, it is common to see vertical fracture planes
through the limestones whilst the shales are apparently unfractured.

50
Fractures generally occur as parallel or almost parallel arrays of planes called fracture
sets. Normally, there are at least two fracture sets in outcrops of competent rocks, and in
many instances there are three or more. One of the sets, at least, is usually at right-angles
to the bedding surfaces. If the bedding is folded the fracture set usually changes both dip
and strike to maintain its bedding-normal attitude around the fold.

Fracture sets may intersect at right-angles in which case they are known as orthogonal
fracture sets, or they may intersect at a much lower angle, when they are called diagonal
fracture sets. Another common orientation is parallel to bedding, giving rise to bedding
fractures.

Fractures have an established relationship to folds, frequently occurring in a set parallel to


the fold axis and a set normal to it (J2 and J1 in Figure 3.66). Commonly a conjugate set
of diagonal fractures will form, with the fold axis bisecting the acute angle between the
sets. The similarity of these orientations with shear planes and extension planes related to
the three principal stresses (refer back to Section 1.2) suggests that diagonal sets form in
response to shear stress and that at least one set of the orthogonal fractures forms in
response to extension. Since fractures by definition show no displacement across them it
is impossible to verify this suggestion.
FIGURE 3.66 The
"classical" view of the
relationships of joints to
fold geometry.
Joint surfaces stippled,
bedding planes
unstippled. J1 cross
joints, J2 longitudinal
joints, J3 and J4 diagonal
joints.

Fractures provide pathways for fluid movement both into and through reservoirs.
Conversely, fractures that are closed actually inhibit movement of fluids and frequently
provide an impermeable barrier within the reservoir that can cause a pressure drop during
production. Fractures that started life as open sets may become closed by precipitation of
minerals from solutions passing through them, or by a change of stress regime changing
them from extension-related to compression or shear related. The reverse may also be
true; once closed fracture sets may become open at a later stage in their geological history.

Fracture sets become closer spaced nearer to major faults, and close to the fault plane
itself the hangingwall block in particular can become highly fractured. This can lead to
the fault zone being an important pathway for fluids and a zone of relatively high
permeability. Should deposition take place along the open fractures, however, the fault
zone can steadily seal up, eventually becoming a barrier to fluid flow.

51
3.2 Fault gouges and fault breccias

Fractures increase in frequency near fault planes, and the act of faulting breaks
through the rock. These two observations suggest that there is a local stress
concentration immediately prior to the production of a fault. When studied in detail,
propagating faults seem to change their locus of action from hangingwall into
footwall and back, depending on the exact nature of the micro-stress field.

Pore fluid pressure within both the footwall and the hangingwall will also control the
nature of propagation of the fault, with increased pressure leading to multiple
fracturing of the fault zone.

These factors together lead to the formation of finely broken fragments of rock
within the fault zone along the fault planes themselves. In its early stages, this forms
a special type of broken rock known as a fault breccia which like fractures may
either enhance or impede fluid flow, or change from one to the other with time.

Continued fine comminution of the wallrock reduces the breccia first to sand-size
and eventually to clay-size fragments. High volumes of fluid flow and localised work
heating along the fault plane favour the precipitation of clay minerals. The end
product of repeated stick-slip movement of the fault is a fine clay smear along the
fault itself, known as a fault gouge. Further deformation of the gouge can cause it to
become lined, a feature often seen on the rock surfaces forming the fault. Such
slickensides are diagnostic features of faulting; lined and grooved clay gouges are
sometimes recovered in well cuttings or core.

When faults contain a gouge it usually inhibits fluid flow in a major way. Even small
faults that may be undetectable on seismic sections can generate impermeable
barriers locally within the reservoir.

3.3 Clay smearing

If a fault passes through a sedimentary sequence containing clays and shales, these
ductile lithologies may be smeared along the fault plane due to the movement of the
hangingwall block over the footwall. This process especially takes place during
movement of synsedimentary faults through unconsolidated sand/clay sequences, so
growth faults within deltaic sequences are particularly prone to clay smearing.

The process has been described from some Niger delta oilfields and from the
Louisiana Gulf Coast, and it has also been observed in outcrops in West German
lignite pits. Clay smearing can lead to a dramatic reduction of permeability across the
fault.

The length over which clay is smeared is dependant on the amount of fault
displacement, the nature and history of fault motion, the thickness of individual clay
layers and the material properties of the clay. Figure 3.67 illustrates diagramatically
the way in which a clay horizon may be smeared out along a normal fault, showing
the patterns of shear fractures that will develop near the fault plane and the irregular
thickness of the clay smear itself. The clay smear may completely seal the permeable
sandy units, as shown in Figure 3.67. Alternatively, the seal may be partial if the
throw of the fault is large or if the clay units are thin.

52
FIGURE 3.67 Clay
seal in a normal fault.

Computer models can be effective in correlating the smear potential of the various clay
horizons with the fault throw to predict the sealing potential of the clay smear. In the
Nembe Creek field, a synsedimentary collapse structure in the Southern Niger Delta,
thirty-six reservoirs in a series of adjacent footwall closures were analysed. In one-third of
these, the proven OWCs could be shown to coincide closely with the prediction of the
spill point level based on the clay smear potential. Most other reservoirs could be
classified as probably underfilled or non-diagnostic cases. There were no examples of
OWCs below the predicted spill points.

Fault seal quality is strongly dependant on differential pressures across the fault, and this
is independent of any potential the clay units within the sequence have for smearing.

53
3.4 Summary of reservoir-partitioning structures

Fractures

Fractures are planes of brittle failure, which generally show no appreciable movement
across them. They are very common in competent lithologies in all tectonic settings, so
they are frequently recorded in reservoir rather than seal lithologies. They related to
extension and shear stresses and most commonly occur in sets normal to bedding, often at
right-angles to each other. They provide important fluid pathways when open and inhibit
fluid flow when closed. Fracture sets may be open initially but close with time, or vice
versa.

Fault breccias and gouges

Fault breccias form by repeated build-up and release of stress along a fault. At an early
stage of breccia development, the broken rock fragments will probably be quite large and
will provide a fluid pathway within the fault zone. With continued action the rock
fragments become extremely fine-grained and will probably cause the fault to act as a
seal. There may be preferential pressure solution of grains or cementation, which will
change the permeability of the fault zone with time.

Advanced brecciation leads to both mechanical and chemical production of clay gouges,
which usually seal faults and prevent fluid flow across them.

Clay smearing

Appropriate clay lithologies may be smeared along a fault plane during fault movement,
particularly in the case of synsedimentary growth faults. The clay smear will be an
effective seal where it is present, but this in turn depend on the amount and type of fault
movement and the thickness and physical nature of the clay units.

54

You might also like