You are on page 1of 42

4

Structural Geology

CONTENTS

1. PURPOSE AND SCOPE

2. STRUCTURAL FEATURES

3. RELATIONSHIP OF STRUCTURES TO
GEOLOGICAL EVENTS

4. PRACTICAL STRUCTURAL GEOLOGY

5. FURTHER LEARNING
LEARNING OBJECTIVES:

At the end of this Chapter the student will be able to:

• Give names for the main types of structural features

• Identify the characteristics of a structural trap

• Describe the main types of faults, and indicate how these structures relate to the
tectonic setting

• Describe typical fold geometries found in the main tectonic settings, and indicate
how these may relate to other structural features

• Understand the importance of fractures and other localised deformation in terms of


their impact on fluid flow characteristics

• Describe the fracture patterns associated with folding

• Identify structural features in core and on dipmeter/image logs

• Identify structural features on cross sections, maps, and interpreted seismic lines

• Describe the causes of, and identify, situations leading to fault compartmentalisation

2
Structural Geology 4
1.PURPOSE AND SCOPE
Structural Geology is a component of Petroleum Geology. It is closely linked to
Geophysics and Rock Mechanics, and there also are strong links to Sedimentation.
Structural Geology has historically been concerned with the study of rock deformation
over geological timescales - sometimes, this activity is also referred to as Tectonics.
Modern Structural Geology has evolved to include an appreciation of Rock Mechanics,
making it possible to produce predictions about the conditions of deformation, and
about the impact of that deformation on rock properties. Rock Mechanics also
includes a consideration of human timescales, and, particularly, how human activity
can influence rock deformation, and vice versa.

In the context of Petroleum Geoscience, Structural Geology is concerned with


creating an interpretation of the deformation history of a zone of interest in the
subsurface in order to understand the trapping potential (exploration) and/or how
production may be affected by the deformation. Traditional structural geologists have
often focused on an analysis of surface outcrops, where the observable scales of
deformation and techniques are usually quite different from those applicable to the
subsurface. Structural geologists working in a petroleum setting have had to learn to
synthesise a wide variety of ideas drawn from many scales and many types of
structures. This review of Structural Geology focuses on those concepts and
techniques that are most-applicable at the reservoir scale, making Structural Geology
relevant to the needs of petroleum engineers.

To meet this goal, this Chapter will address:

(a) The role of Structural Geology in exploration:

• Structural trapping mechanisms - faulting, folding


• Structural history - burial, tectonism and uplift
• Structural features on seismic lines, wireline logs and cores

(b) The role of Structural Geology in reservoir development:

• Fracture production
• Fault properties - compartmentalisation, sub-seismic faulting
• Fault maps - juxtaposition diagrams, shale gouge ratio
• Stress effects

The Chapter begins by describing the geometries of important structures and the
associated terminology. The next section discusses the relationship of structural
features to other geological events. The following section focuses on practical issues
involving applications of the concepts presented previously. A final section suggests
some ways to learn more about this subject, including a set of exercises. Rock
Mechanics concepts are treated in an Appendix.

Because of the close links between Structural Geology and Rock Mechanics, this
Chapter and the Appendix should be studied in parallel.

Department of Petroleum Engineering, Heriot-Watt University 3


2. STRUCTURAL FEATURES

2.1 Faults
A fault is a more-or-less planar surface or zone, across which the rocks on either side
have been moved by shear displacement (i.e. displacement parallel to the fault
surface). Faults can be sharp (infinitesimally-thick) planes, and they can also be wide
zones consisting of an array of complex deformation features. Faults represent a
yielding of the rock mass, and importantly, they indicate that deformation has become
localised (as opposed to distributed) at the scale of observation. Fault geometries, and
fault patterns, are used to infer large-scale deformation states.

The majority of faults are not vertical; instead, most are inclined. The angle “down”
from horizontal is called the dip of the fault plane, and the compass direction of the
horizontal line lying in the fault plane is called the strike (Fig. 1). A vertical fault has
a dip of 90o, and non-vertical faults have dips that range from very shallow (10-30o)

;
y ;
y
to moderate (40-60o) to steep (70-89o). The dip of the fault plane, along with the sense
of motion (see below), is used to categorise the types of faults.

y
; y
; ;
y ;
y
Angle of Dip (=22º) Top of Rock Layer

yyy
;;
y
;
N 300m
75º

y
;
200m

y
;y ;
y
Strike

;
y
100m

; y
0m
Dip

Another Rock Layer

;
Horizontal Line

Left-Hand Rule:
If left thumb points down dip,
then left index finger points in strike direction.

In this example: dip = 22º, strike 105º

This is written as: 22/105 Figure 1


(other conventions exist)
Definition of dip and strike
of a surface

Note: Structural Geology is intimately linked with geometries. The definitions and
explanations of structural features necessarily require drawings of these geometries.
There are three principal methods used to illustrate structural forms (Fig. 2). The 3-
D block diagram is perhaps the most readily understood of the three methods, since
it is “visual”. A cross section can be thought of as being the side of a block diagram
(even if the other side and top are not shown), and a map is simply the top of that
diagram projected onto a 2-D plane (a piece of paper). Although there are various
kinds of maps (see Chapter 7), structural features are usually described using either
geological maps or structure contour maps.

4
Structural Geology 4

;
y y
; ;
y ;
y yy
;; y
;
;;;
yyy
;
y
;;
y;
y
y ;
y
;
y ;
y
;
y
;
y ;
y
;
y ;
y ;;
yy
;;
yy
y
;
;;;
yyy
;
y
;;;
yyy
;
y
;
y ;
y ;
y
;
y
Top of Block

;
y y
;
;
y ;
y
y
; ;
y
;;
yy
;
y ;
y
;
y;
y;
y
;
y ;
y
;
y ;;
yy ;
y
yyyy
;;;;
Geological Map

;
y;y;;
y ;
y
;
y
;
y
;
y
;
y;
y;
y;
y
;
y
;
y;
y
;
y
;;
yy ;
y
yyyy
;;;;
;;;
yyy
y y;y;y;y;y; yyyy
;;;;
;;;
yyy
Figure 2 Block Diagram Side of Block
(Showing a Dome)
Types of illustrations used
to depict structural features Cross Section

Each fault cuts the entire rock mass into two fault-blocks. In the case of non-vertical
faults, the fault-block lying below the fault plane is called the footwall, regardless of
the sense of displacement of the fault, and the block above the fault is called the
hangingwall (Fig. 3). The terms footwall and hangingwall derive from the mining
industry where fault planes are often encountered in underground workings (some
faults in orogenic belts are subject to mineralisation, implying that they were good
fluid conduits). The footwall (imagine a miner’s feet on the footwall) and hangingwall
(imagine a miner hanging from the roof of the mine) cannot be defined for a vertical
fault.

Hangingwall

Footwall
Figure 3 Hangingwall
Definition of the terms
footwall and hangingwall, Footwall
with respect to the dipping
fault plane, but independent
Normal Fault Reverse Fault
of the sense of movement

The types of faults are defined by the sense of movement along the fault plane (Fig.
4). In dip-slip faults (the slip motion is parallel to the dip direction), if the hangingwall
moves down (with respect to the footwall), this is called a normal fault. Normal faults
are associated with extension (lateral increase in dimension). If the hangingwall rises
over the footwall, this is called a reverse fault. Reverse faults (thrusts are reverse
faults whose dip is low - less than 25o) are associated with shortening (lateral decrease
in dimension). In the case of strike-slip faults (where the movement is parallel to
strike), we use the terms left-lateral and right-lateral to indicate the sense of relative
motion as seen looking down on a map.

Department of Petroleum Engineering, Heriot-Watt University 5


NORMAL

REVERSE

Figure 4
contraction
Types of fault defined by
extension
displacement along the
STRIKE SLIP
fault plane A: Normal (dip-
slip) fault; B: Reverse (dip-
slip) fault; C: Strike-slip
Lateral Movement
fault (left-lateral shown)

Faults can be simple or complex zones of displacement (Fig. 5). In the petroleum
industry, faults are often represented as simple, single breaks (this is almost always the
case at the scale of a reservoir map), but they are in reality more likely to be complex
zones. A single fault break may well become a much more complex fault zone as the
scale of observation magnifies. Large, “intact” blocks of rock that are found within
a fault zone, surrounded by sheared and distorted rocks, are called “horses”. Synthetic
(similar dip) and antithetic (opposite dip) faults are names applied to minor faults that
are associated with larger faults.

Figure 5
Types of fault zones. A: a
simple, single surface of
shear; B: a fault zone
composed of a set of shear
surfaces; C: a distributed
(ductile) shear zone

Faults tend to form in groups, or arrays, giving rise to various geometric arrangements
that have the potential for trapping hydrocarbons. For an array of faults that have a
similar strike, it is common to find that some of the faults dip in one direction, while
some dip in the opposite direction (see Rock Mechanics for an explanation). When
seen in a cross-sectional view (Fig. 6), such an arrangement of normal faults produces
blocks that are uplifted or dropped down relative to one another. In this pattern, the
uplifted blocks are called horsts, and the down-dropped blocks are called grabens.
Curiously, deformations that shorten the layers seem to be less likely to produce equal
numbers of left-dipping and right-dipping faults; perhaps this is related to the idea that
such faulting is related to the spreading of a crustal-scale wedge of rocks, a process
which has a preferred direction.

6
Structural Geology 4

Horst Horst Half Grabens

Graben

Synthetic

Antithetic Listric Fault


EXTENSION

Figure 6
Sketch of A: extensional;
and B: shortening
structural regimes CONTRACTION

Thrust Fault (Listric)

Large faults are often not planar. A listric fault is a curved fault that is steeply-dipping
at shallow structural levels, and gently-dipping at deeper levels. Listric normal faults
are important in extensional domains because they provide for rotations of fault blocks
(Fig. 7). The tilting can provide a chance for new sediment deposition, which, in this
case, produces a wedge-shaped rock succession. Such a situation is called a half-
graben (it is a down-dropped block, but only on one side). The shallow portion (crest)
of a tilted fault block is a potential hydrocarbon trap, while the deep portions of large,
tilted fault blocks may subside far enough to become heated and thus turn into a
hydrocarbon kitchen. Listric faults are also important in shortening, where they may
be associated with folds (see below).

shales

sandstones leaks, re-migrated hydrocarbons, re-trapped in shallow layers

erosion

unconformity
late-trapped
sourc late-mature

;;
yy
hydrocarbons e roc d hyd
early-trapped Rese k rocarb
hydrocarbons rvoir ons
Unit
migra
tion
compaction sourc
e roc

yy
;;
;;;
yyy
k

Small hydrocarbon
accumulations

;;;
yyy
early
Figure 7 -ma
ture
d hy
mig droc
arbo
Importance of extensional Res r
ervo ation sou
ns coarse clastics
ir U rce
nit ro from eroded
faulting relative to the ck
fault-block crest
sou
rce
Petroleum System roc
k migration
along fault?

Fault displacements are not constant along a fault surface. Displacement varies
systematically along a fault and reaches a maximum near the centre of the fault

Department of Petroleum Engineering, Heriot-Watt University 7


surface. The edges (or limits) of a fault surface, known as the fault tips, are locations
of zero displacement. Real fault zones are usually made up of numerous fault
segments that partly overlap, with intervening regions of distorted rock (refer to Fig. 8).

;;;
yyy
; yyy
y ;;;
y
; yyyy
y
yy
;; ;
y
;;;;
yy;
;;
;;
yy;;
y ;;;;
yyyy
y
;;yyy
yy
;;;
yyy ;;;
;;; yyy
yyy ;;; Complex Fault Zone
Detail of Distorted Rocks and
Overlapping Fault Segments
Figure 8
Examples of fault zone
complexity

It has been observed that the larger (longer on a map or cross section) faults in a system
have the largest displacements. It is also noted that there tend to be more, smaller
faults, and fewer, large faults. This relationship is often expressed as a power-law
curve (Fig. 9). With such a relationship, the sub-seismic faults (those that are too
small to be identified on seismic data) can be modelled (but see below for cautions
against using this technique blindly).

100
"Sub seismic"
Cumulative fault density (/km2)

10

1
Figure 9
"Seismically
Resolvable" Fault population curve
0.1
exhibiting a power-law
0.01 relationship. Such curves
have been used to
1 10 100 1000 extrapolate fault
Maximum Throw (m) populations below the limit
of seismic resolution

The common fault types are sometimes interpreted to imply the orientation of the
principal stresses that existed when the fault was formed (see Appendix). In this
approach, ideal normal faults would have the maximum compressive pricipal stress
(σ1) vertical, ideal reverse faults would have the minimum principal stress (σ3)
vertical, and ideal strike-slip faults would have the intermediate principal stress (σ2)
vertical (Fig. 10). Although this simple relationship is a useful learning tool, real states
of stress are rarely as uniform and homogeneous as implied by this conceptual model
(this issue is treated further in the Rock Mechanics Appendix), and its over-
application is a common pitfall.

8
Structural Geology 4

σ
1 σ3 σ
2

σ
2
σ
2

30º

Figure 10
Orientation of principal
σ3
stresses for the three ideal σ
1
σ
1
σ
3
types of faults

2.2 Folds
A fold can be defined as the deflection of a marker surface (e.g. from a planar shape
before folding to one that is non-planar after folding). The majority of situations of
interest to Petroleum Engineers involve layered, sedimentary rocks, so the “marker
surface” is usually bedding (but this is not always the case). Even if the rock
succession is a monotonous stack of the same layer on top of the same layer, the
bedding planes represent a significant element of heterogeneity. In the more usual
case where there are lithological variations (e.g. sand/shale/sand, etc), there is an even
greater degree of heterogeneity. These mechanical variations are what make “fold-
ing” a distinctive structural process.

You will probably be surprised if you consult a range of textbooks to get another point
of view concerning folding: none give a definition of folding that is any more definite
than that given in the preceding paragraph! The reason for this “hedging” is that most
geologists wish to classify folds according to their mode of origin: i.e. their genesis.
If it were possible to do this un-ambiguously, then the problem would be solved.
Unfortunately, folding processes are still the subject of research, and the genesis of
any particular fold cannot be determined with certainty, so this strategy cannot be
adopted. We will return to the topic of fold genesis, but first, we need to be sure that
we know the terminology of fold shapes.

Fold shapes
The first point to make concerns the word “surface”. Rock layers, and other-shaped
rock bodies, are bounded by surfaces. A surface is a curvi-planar entity that, in
mathematical jargon, has only two dimensions (a surface has infinitesimal thickness).
A large portion of the fold-shape naming scheme (Fig. 11) is based on the shape of
surfaces. In this context, the “surface” is usually the bedding plane bounding the top
of a rock layer. (But, remember, this bedding plane also bounds the bottom of the next-
higher layer.) The key words are: crest line (the line that represents the locally-highest
elevation), trough line (locally-lowest elevations), inflection line (boundary between
convex-upwards and convex-downwards), culmination (highest point of a crest or
trough line), and depression (lowest point of a crest or trough line). Note that the shape
variations of surfaces are distinctly three-dimensional (it is important that you don’t
forget that the 2-D cross-section drawings of folds that we use are illustrative, and not
very realistic!).

Department of Petroleum Engineering, Heriot-Watt University 9


c

c i

D e pre s s io n
tion i
mina
Cul
i i
e c
ur fac
n ar S i
C u r v i- P la i Figure 11
t
t Terms that describe the
shape of a curvi-planar
Crest Lines (c), Trough Lines (t) and Inflection Lines (i)
surface

Much of the following material is based on how a cross section of a surface appears.
In mathematics, the key concept needed for evaluating the shape of a surface is
curvature. Curvature is actually the inverse of the radius ( 1 / r ) of the circle which
has the same shape as a small segment of the surface. If we determine the curvature
of a folded surface (via equations that are not important for our purposes), there will
usually be some places where the curvature is higher than in nearby regions (i.e. the
folding of the surface is “tighter” in the high-curvature sites). We call such high-
curvature parts of a fold “hinges”. The less-curved (straighter) portions of the fold
are called “limbs”. Both single-hinge and multi-hinge folds are possible. The fold
axis is the line that is formed from the intersection of the axial surface with some layer
boundary. The fold axis and the hinges are usually parallel (Fig. 12).

High-curvature sites
(Hinges)

Less-curved regions
(Limbs)

Interlimb
Angle

Stylised Representation with Figure 12


Straight Limbs and Point Hinges
Fold limbs and hinges

A useful geometric characteristic of a fold shape is its interlimb angle (Fig. 12). This
parameter has to do with the apparent tightness of the fold, or the angularity of the
hinges. The following table gives the descriptive terms that are applied to ranges of
the measured interlimb angle.

10
Structural Geology 4

Descriptor Interlimb Angle (deg)


Gentle 180 - 120
Open 120 - 70
Close 70 - 30
Tight 30 - 0
Isoclinal 0

Other terms are used to describe fold patterns, including fold symmetries. Although
all folds do not occur in the form of wavetrains, the following definitions are most
easily visualised if we draw such a set of repeating fold forms. Enveloping surfaces
delimit the deflections of the surface about some median position. The fold amplitude
is half the distance between the enveloping surfaces. The median surface joins the
primary inflection points. The wavelength is the distance between comparable
inflection points. Note that the wavelength is shorter than is the distance between
inflection points measured along the layer. Symmetric folds have equal-length limbs,
and asymmetric folds have unequal-length limbs (Fig. 13).

Periodic Symmetrical Waves


Enveloping Surface

As A As

i i i i Median Surface, m

W
As W= Wavelength
A = Amplitude
i = Inflection Points
As = Axial Surface
θ = Inclination of axial
surface relative to
enveloping surface

Periodic Asymmetrical Waves


Enveloping Surface

As
A
θ
i i i i m
Figure 13 W
Terms applied to repeating
fold shapes As As

Department of Petroleum Engineering, Heriot-Watt University 11


Of course, our real interest is in the folding of rock layers (which have finite
thicknesses). When a layer is flexed “up”, we call this structure an anticline; when
the layer is flexed “down”, we call this structure a syncline (Fig. 14). Because
petroleum is usually more buoyant than the aqueous fluids that are otherwise present
in the pore spaces of rocks, it tends to migrate upwards. An anticline is an ideal shape
that could serve as a trap for oil and gas (assuming that the migrating hydrocarbons
had entered into the reservoir layer, had moved along it, and that overlying it there was
a seal; see separate Chapter on Petroleum Play).

Anticline: bows upwards, older rocks in


the middle, dips are "Away"

Syncline: bows downwards, younger rocks


in the middle, dips to the centre
Figure 14
Definition of anticline and
syncline

Our real interest is in the (usual) case where there are stacks of layers. Imaginary
surfaces can be created (Fig. 15) to join up the various points defined on each layer
boundary. Axial surfaces bisect the limbs. For the types of folds usually encountered
in petroleum systems (e.g. approximately-constant layer thicknesses; see below), the
axial sx"Vace is essentially the same as the hinge surface, which joins all of the hinge
points. Inflection surfaces join all of the inflection points, and crest and trough
surfaces (not usually drawn) could join all of the high and low points, respectively.
Inflection surfaces define fold domains - within which the changes of fold shapes are
usually regular and predictable.

yyyy
;;;;
;;;;
yyyy
Cross section of multi-layer stack showing
complex array of axial surfaces
;;;;
yyyy
Inflection Surfaces
Define Fold domains
Figure 15
Imaginary surfaces
segmenting a folded
succession

12
Structural Geology 4
A fold may well have a fold axis that is not horizontal. Non-horizontal axes are said
to be plunging (Fig. 16). (A plunging line is specified by its trend - compass direction
of the downward-pointing end of the line, and its plunge angle - measured down from
horizontal). If a fold has a plunging axis, and a non-vertical axial surface, its axial trace
(where the axial surface intersects the ground), and its crestal trace (where the crestal
surface intersects the ground), may not coincide on a map. This is a point that seems
to greatly trouble many students! The fold profile is a section of the fold taken at right
angles to the fold axis. In plunging folds, the profile is not the same as the (vertical)
cross section.

True Shape (Profile) of Plunging Cylinder

Cross Section of Pipe

Map Shape

Horizontal Line

Plunge Angle
Fold A
xis

Figure 16
A plunging fold
Axial Surface

One fold classification scheme is based on variations or constancy in the thickness of


the folded layers (Fig. 17). Parallel folds have constant layer thicknesses (measured
perpendicular to the layer boundaries). Concentric folds are special parallel folds that
have more-or-less constant curvature centered on a point. Similar folds have a
constant distance between layers (measured along a direction that is parallel with the
axial surface).

Department of Petroleum Engineering, Heriot-Watt University 13


t

t t
t

Figure 17
Parallel Fold Concentric (Parallel) Fold Parallel and concentric
Thickness (t) constant along layer Constant layer thickness folds
and constant curvature

Another classification scheme relies on the patterns of dip isogons as viewed in a fold
profile (Fig. 18). Isogons are lines joining points on different surfaces that have the
same dip. There are three classes, although the first class is sub-divided into 1A, 1B,
and 1C sub-classes. In a multi-layer sequence of rocks, and especially if there are
strong mechanical contrasts between the layers, mixed fold classes can occur. Most
of the folds of interest to Petroleum Engineers belong to Class 1B.

1A

1A 1B Parallel 1C
1B Parallel

1C

Class 2

2 Similar

3 Class 3, Divergent Isogons


Figure 18
Fold Shape Dip Isogons
Fold classification based on
(Classes) (lines of constent dip) dip isogons

A suite of terms exists to provide succinct communication about the orientation of a


fold (Fig. 19). These terms are based on: the fold’s tightness (i.e. its interlimb angle;
see above); its symmetry; the dip of the axial surface; and the plunge of the fold hinge.

14
Structural Geology 4

Descriptor Dip of Axial Surface (deg)


Horizontal, Sub-horizontal 0, 1 - 10
Gently Inclined 10 - 30
Moderately Inclined 30 - 60
Steeply Inclined 60 - 80
Sub-vertical, Vertical 80 - 89, 90

Descriptor Plunge Angle (deg)


Horizontal, Sub-horizontal 0, 1 - 10
Gently Inclined 10 - 30
Moderately Inclined 30 - 60
Steeply Inclined 60 - 80
Sub-vertical, Vertical 80 - 89, 90

Each combination of terms suggests a geometric image. A full description links these
terms together: e.g. an asymmetric, tight, inclined, plunging anticline. If the
descriptors are quantified (e.g. strike and dip of the axial surface, trend and plunge
angle of the hinge, etc), then the fold orientation is also communicated.

Horizontal Fold Axis Plunging Fold Axis Vertical Fold Axis

Upright Horizontal Upright Plunging Vertical

Inclined Horizontal Inclined Plunging

Figure 19
Fold orientations
Recumbent Horizontal

Department of Petroleum Engineering, Heriot-Watt University 15


In practice, these terms are not frequently used by petroleum geoscientists or
engineers, but it is useful to know of their existence so that you will be able to
understand fold descriptions that you might read, or that might arise in special
circumstances. Although the terms are not frequently used, it is worth noting that, if
a fold can be described in this way, there is an implication that the fold’s geometry is
somewhat “regular”, and hence, predictable.

Types of flexure
Many geologists make a distinction between buckle folds and folds created by
bending. Buckles are supposed to result from the shortening of a layer, while bends
are created by displacements imposed normal to the layering. However, bending may
(and usually does) produce folds that are shorter (occupy less distance on a map) after
folding than the rocks were before the folding, so this view is inadequate to distinguish
between these types of fold.

A slightly more accurate notion about buckling is the one that states that the rock layer
is “pushed” along its length (Fig. 20). In practice (for natural folds), this statement
means that distant points on the layer are pushed closer together, and the layer deflects
(folds) away from its previous planar shape. The shape of the deflections (primarily
their wavelength) is controlled by a number of factors, including the thickness of the
layer, and its properties relative to the properties of the surrounding materials. This
sort of fold model is thought to produce a wavetrain of folds that have similar
geometric characteristics. Individual folds in the wavetrain grow by increasing their
amplitude, and, because the fold hinges are assumed to be “fixed” in the rock, the
wavelength decreases simultaneously.

Loading Examples

Buckling Fold Train

Block Fault

Figure 20
Bending Differences in loading: (a)
buckle folds and (b)
Diapir Differential
Compaction bending

In map view, buckle folds are associated with a strain ellipse (refer to Rock Mechanics
Appendix) that has its short semi-axis oriented in the “push” direction, and its long
semi-axis oriented along the fold axes. Most real buckle folds are not infinite along
their axis, but instead transfer their shortening to neighbouring folds. The length-to-
width ratio of most buckle folds (as seen on a map) is about 10:1.

It is appropriate here to introduce a notion that applies to all sorts of structures, but

16
Structural Geology 4
which is particularly well illustrated by the image of fold wavetrains. As noted above,
folds tend to be elongated. In a map view, this orientation of this elongation is referred
to as the longitudinal direction, or the trend. These words are also used to indicate the
“long” direction of any structural type. The perpendicular orientation on the map is
called the transverse direction, or, less-precisely (but more commonly), the “dip
direction”. Although cross sections can be constructed along any alignment, it is most
common to draw them in the transverse or “dip” direction, since this orientation
illustrates most-effectively the major changes in shape. For specific purposes, other
cross-section alignments can be used.

A point not (yet) recognized in the literature is that, in a stack of layers, with each
having its own mechanical properties, one layer might buckle whilst the others are
subjected to bending because of the deflection associated with the controlling buckle.
In regions that have been shortened, there are important genetic associations between
thrust faults and bedding-plane faults, and folds (see later in this Chapter).

Regardless of the cause of the folding, flexures in the upper crust of the earth (i.e. those
of interest to Petroleum Geoscience) are highly dependent on the fact that the rocks
are layered. The bedding planes between rock layers are mechanical discontinuities
that are available for slip. When folding occurs, some (but not all) bedding planes do
slip, and this process (called flexural slip folding) dramatically alters the pattern of
deformation - as compared to the folding of a single thick layer that does not have
internal bedding planes. Bedding-plane slip is extremely important in terms of
limiting the magnitude of the strains that are created at any point in the flexed
succession of layers (e.g. fracture intensity), and in controlling the extent of fractures
that may be induced. When flexural slip occurs (this is the “normal" case for the upper-
crustal flexures of interest to Petroleum Engineers), the fractures that are created are
distributed differently than would be predicted by power-law relationships (see
above). Because of the partitioning of strain associated with flexural slip, sequences
of rock layers can be considerably bent without undergoing extreme internal distor-
tion (Fig. 21).

Bending Strains (expressed as fractures)


Active
Slip
Surfaces
A.

To-be-Activated
Slip Surface
B.
New Bending Strains

Now-Active
Slip
C. Surfaces

Figure 21
Resulting
Deformation in flexural- Superposed
D. Fracture Strains
slip folds

Department of Petroleum Engineering, Heriot-Watt University 17


2.3 Fault / Fold Interactions

Contractional structures
There is a major chicken-and-egg question that arises in the interpretation of folds and
thrusts. It is common in fold/thrust belts (see below) to observe large-scale,
asymmetric folds whose overturned limbs are faulted (Fig. 22). Is the faulting a
consequence of the folding, or is the folding a consequence of the faulting?

100's of m

Reverse fault
cutting overturned
forelimb
?

Figure 22

It has been possible to develop kinematic models of the development of folds that are
related to fault movements. The various types of model include: fault-bend folds,
fault-propagation folds, detachment folds, and break-thrust folds. Now, some twenty
years into this process of creating these kinematic models, they have become very
sophisticated (but not necessarily right!). There are models that address variations in
limb dips, interlimb angles, fault dips, displacement gradients on faults, fixed versus
moving hinges, layer thickness changes, etc, etc. The hope has been that different
modes of formation could be distinguished on the basis of differences that could -
ostensibly - be measured in a given structure. Unfortunately, clever people keep
finding a new parameter that allows them to develop another model to permit them to
interpret their favourite fold as being of type X. Although the initial motivation - to
distinguish the mode of formation - may not be met, these kinematic models are useful
in that they provide us with hypotheses about how these types of structures form and
evolve.

Some important aspects of these kinematic models of contractional structures are (Fig. 23):

• The (oft-assumed) ramp-flat-ramp shape of thrust faults


• The footwall cut-off angle
• The angle between the bedding that is being truncated and the hangingwall flat
• Whether folds remain as the parallel type of whether the layer thicknesses change
• Whether fault displacement is constant
• Strong potential for syn-deformation deposition of sediments, and their immediate
involvement in further deformation

18
Structural Geology 4

h
Fold geometry produced
by motion along fault
tf γ

f
α

Flat

Ram
p

Shape of Fault
Flat
Figure 23

In recent works, various researchers use the kinematic models to calculate an apparent
strain state in the resulting structures. If the inferred strains are thought to relate to
deformation (e.g. fractures), then there is potential to use such models to aid in
reservoir management. However, we do not yet have good forward simulations that
are based on proper rock mechanics, so any such predictions about the mechanical
state should be treated cautiously. Probably the most crucial shortcoming of these
efforts is that none (to date) consider the bending of the layers, or the impact of
flexural-slip folding.

Extensional structures
In contrast to the situation with fold/thrust structures, there has been considerably less
attention paid to fault/fold relationships in extensional settings. Although this
statement is true, a major example of such a relationship has been known for a very
long time: roll-over anticlines (Fig. 24). However, the kinematics of such features
have been studied and understood for only a short time.

Note truncations

(tracing from
seismic image)

Rollover
anticline

listr
ic n
orm
al fa
Figure 24 ult

Department of Petroleum Engineering, Heriot-Watt University 19


In extensional deformation, the orientation of most faults is at high angles to the
layering. This means that any resulting folds are ‘bends’, and certainly not ‘buckles’.
Until fairly recently, it was widely believed that folding in extensional situations did
not occur, and that, instead, the rock layers simply faulted. (It was believed that the
layers would be stretched, and that the resulting state of stress would have very low
stress magnitudes, leading to a situation where the rocks were very weak, and hence
liable to fault. This belief failed to account for the operation of the whole system, and
especially for flowage of some of the rock units.) Any observed flexure of the layered
rocks was dismissed as ‘drag’. Fortunately, good research is now showing that
extension-related folding is an important structural style that can be rationally
explained using similar kinematic approaches as have been used for contractional
settings. Full forward simulations are, however, not yet available for these structures,
either.

Some important aspects of these kinematic models of extensional structures are:

• Angle of layering relative to local fault orientation (many extensional faults are listric)
• Constancy / variability of displacement on the fault
• Presence/absence of ‘weak’ units to permit flow, and related detachment of
overlying layers
• Strong potential for syn-deformation deposition of sediments, and their
immediate involvement in further deformation

2.4 Fractures and Joints


When rocks become deformed, they exhibit one or more of a variety of yield responses
(see Appendix). One of the most common types of response is fracturing, or breakage
of the rocks. Each such break represents a mechanical discontinuity in the rock mass.
These structural discontinuities are amongst the most common of all geological
features: every outcrop and most cores exhibit some sort of fracturing. Fractures and
other discontinuities affect nearly every petroleum reservoir, either by enhancing the
production, or by causing problems for production.

There are different terms that can be used to refer to structural discontinuities —
roughly distinguished by the scale of the feature and the amount of displacement.
Although other authors may propose very specific definitions, we recommend a
pragmatic approach (Fig. 25). We suggest that the terms fracture (preferred) or joint
be used for sharp, localised breaks or discontinuities, and that the term fault be used
to refer to a plane or zone across which there is considerable shear (relative
displacement parallel to the plane/zone). There is inevitably a difficulty in judging
between the use of the terms fracture or fault in some cases. (Many geoscientists wish
to use the term joint for fractures with little discernable movement, but any supposed
distinction between the terms joint and fracture is of little or no practical use.)

20
Structural Geology 4

yyyyyy
;;;;;;;;;
yyyyyy
;;;
Figure 25
Fracture Fault
Displacement distinguishes (Joint)
between fractures and
faults Same, Unspecified Scale

Fractures and/or joints often occur in systematically-aligned groups such that there is
a similar dip and strike to each of the fractures in the group. Such a group is called
a fracture set or joint set (Fig. 26), and sometimes the word “ systematic” is added (e.g.
systematic fracture set). Multiple sets of fractures/joints (each set being characterised
by a different strike and dip) also commonly occur; these groupings of sets are called
fracture assemblages if they are thought to be causally (genetically) related.

;
yy
;
;
y yyyy
;;;; ;
y
;
y;
y
;;;yyy
yyy
;;;
y;
y
;
y yyyy
;;;;
; yyyy;;;
yyy
;
y
;
y;
y
Figure 26
Illustration of fracture sets
and assemblages
;;;yyy
yyy
;;; ;;;;
;;;
yyy A Fracture Set
Fr
actu
re
s

Two Related Fracture Sets


(= Fracture Assemblage)

There is presently a considerable level of interest in fractures. The primary stimulus


is to understand the impact of fracture patterns on reservoir performance. A typical
fracture analysis will attempt to determine:

• the distribution and geometry of the fracture system(s)


• the surface features of the fractures
• the relative timing of the formation of different fractures
• the geometric relationship of fractures to other structures

This knowledge can lead to the creation of a genetic model for the formation of the
fractures, from which predictions can be made concerning the distribution in areas that
cannot be sampled.

The typical spacing between fractures in a set is often seen to be a function of lithology
and bed thickness (Fig. 27).

Department of Petroleum Engineering, Heriot-Watt University 21


1.2
Bed T hickness (m)


Sandstones
0.6

Limestones

Figure 27
Relationships between
1 2
fractures and bed thickness
Spacing Between Fractures (m)
in sandstones and
limestones

Features on the surface of a fracture can provide important information about its origin
(Fig. 28). A set of curvilinear ribs, defining a feather-like (“plumose”) structure, is
evidence of an extensional fracture (Mode I; see Appendix). Mineralised surfaces
with lineations are called “slickensides”, and these are indicative of shear movement
(usually Mode II). Crystal growth on the surface of a fracture is an indication that the
fracture has been a void (open space) at some point in its history, allowing minerals
to grow from circulating fluids. A special case of fracture-surface mineralisation is
the “crack-seal” arrangement, from which we can infer that mineral deposition was
concurrent with fracture opening. In carbonate rocks, fractures can become zones of
dissolution, leading to open fissures. The nature of fracture surfaces is an important
consideration for the performance of the fractures during production of a reservoir.
Partly-mineralised, uneven fracture surfaces are less likely to close up (as fluid is
withdrawn) than may be the case with simple, co-planar extensional fractures.

Fracture Surface
Figure 28
Direction of  Sense Features on fracture
Fracture  of She
ar surfaces: Left) plumose
Propagation
structure indicating brittle,
extensional fracture;
Centre) slickensides
indicating shear; Right)
mineralised surface
Plumose Structure Slickensides Crystal Growth indicating an open fracture

22
Structural Geology 4
Occurrence of fractures
Fractures represent the distortion (e.g. a state of strain) that accumulates in rocks
undergoing cataclastic deformation. Therefore, they can occur anywhere that
conditions favour cataclastic deformation mechanisms. They often occur as subsidi-
ary elements associated with larger structural features, such as faults (Fig. 29). They
can also occur in a distributed fashion throughout a body of rock; this case would
suggest a style of deformation that is not localised.

Synthetic Fractures

Figure 29
Fractures associated with a Antithetic Fractures

normal fault

There are important cautions concerning the interpretation of fractures seen in rock
outcrops. Some of the visible fractures may have been produced by processes that are
active at the earth’s surface (e.g. thermal distortions caused by heating/cooling
cycles). Others may represent the breakage of rocks related to the relief of stress as
the rocks have been brought to the surface (e.g. by erosional removal of overlying
rocks). Still others may represent ‘true’ deformation caused during the uplift (this
might happen in tectonically-active regions). Of course, some of the fractures that are
observed in outcrops were created by tectonic processes, and these are the ones that
are representative of the subsurface distribution that we are interested in.

Fractures are also important in rock successions that have been folded, or flexed (Fig.
30). In folded rocks, fracture density (the inverse of fracture spacing) is sometimes
thought to be related to the curvature of the rock layers. The highest fracture density
seems to occur where there is maximum curvature of the rock layers (such as in fold
hinges), and this relationship is often exploited to assist in targeting fracture produc-
tion in reservoirs.

Figure 30
Fractures associated with
folding. Note the different
fracture patterns associated
with the inner and outer
surfaces of the layer in the
crestal region

2.5 Strain Fabrics


As noted above, fractures are one expression of strain. There are other mechanisms
by which rocks can change their shape (see Appendix), and these produce fabrics in
the rock. ‘Fabric’ refers to planar or linear alignments of minerals or discontinuities

Department of Petroleum Engineering, Heriot-Watt University 23


(so fractures are a ‘fabric’). Non-fracture fabrics of interest in the petroleum context
are usually found in muddy and shaley rocks. Because these rock types can undergo
substantial flow, there is potential for the platy minerals to rotate and become aligned.
Such fabrics can give information concerning the structural history, and especially
about the physical conditions that may have existed (pressures and temperatures).
These fabrics can also impact drilling - primarily because they are not ‘expected’ - by
deflecting the path of the bit in the same way that normal bedding lithological contrasts do.

2.6 Diapirs
The Earth’s gravitational attraction is one of the most important factors in creating the
loads that cause deformation. The mass of the rocks that comprise basins, when acted
on by the gravitational acceleration, produces large forces. Variations in density can
lead to significant force anomalies, and these, in turn, can produce deformational
responses.

An important type of structural feature is the diapir (Fig. 31). A diapir is a body of
‘flowable’ rock that migrates upwards due to its lower density (compared to the
surrounding rocks). Rock salt commonly forms diapirs, but under-consolidated
mudstones also can become diapiric. (Large granitic intrusions that rise into the
middle crust are also diapiric - while molten, their density is less than that of the
country rock around them.)

Bending of Layers Above

Truncation of
Layers at Figure 31
Depth Diapir that pierces lower
layers and flexes upper
layers
Diapir

Salt and shale diapirs are commonly treated as ‘closed systems’, meaning that we
assume that there is no loss or gain of material. Thus, when the diapir rises, it leaves
a virtual void somewhere below. Such voids do not, of course, actually exist. Instead,
the surrounding rocks subside into the space vacated by the upward movement of the
diapir. Because such a process can be long-lived (perhaps a hundred million years),
and because sedimentation occurs during this time period, and because the process of
sagging and movement can cause serious distortions of the rock succession (including
both normal and reverse faulting), diapirs can be structurally very complex. However,
diapirs are often good for trapping hydrocarbons. This is because the diapir itself often
serves as a seal, so that rock layers that are depositionally-truncated against the diapir,
or structurally-truncated by its movement, can serve as reservoirs.

24
Structural Geology 4
3. RELATIONSHIP OF STRUCTURES TO GEOLOGICAL EVENTS

3.1 Tectonics and Sedimentation


Some faults can be active during the deposition of sediments. A common situation
occurs at the margins of rapidly-subsiding basins: the young rocks are weak, and
“slide” towards the deeper portion of the basin (Fig. 32). As this motion occurs, the
rocks are stretched, and large, listric normal faults are produced. Because one side of
the fault subsides more rapidly than the other side, the rock layers (each representing
a time increment) that are deposited are of different thicknesses in the footwall and
hangingwall blocks. Such a fault is called a growth fault (the sediment “grows” in
thickness across it).

A
A'

B
A B'
Figure 32
Syn-depositional faulting A'
can lead to different rock
thicknesses in the footwall
(eg, layer A) and
hangingwall (layer A’)
blocks

In basins (i.e. where sediments are depositing and accumulating), structures are often
very important in terms of localising where deposition takes place. For example, a
fault at depth might move, producing a sea-floor bathymetric expression (e.g. a low-
to-high change in sea-floor elevation). Something similar might happen in a river
valley in an area of active tectonism. Currents that are carrying sediment would
preferentially flow into the lower places, and their sediment load would tend to be
deposited there. Such a sequence of events can be responsible for causing a
lithological change in the resulting rock sequences: e.g. a lateral change from
sandstone to shale.

The tops of tilted fault blocks, or the crests of folds that are caused by faults beneath,
can also have an impact on the distribution of the sediment that is deposited. In a basin,
such crests can actually extend above the water, and be subject to erosion. The debris
that is produced during erosion may be deposited locally, leading to anomalous coarse
sediments amongst finer-grained materials. If the tops of the structures are not
exposed, but instead are at an elevation to cause them to be covered by only shallow
waters, there may be little or no deposition on their crests. If the waters are clear and
warm, reefs might form on the tops of such blocks. Such spatio-temporal variations
in lithology are quite common in sedimentary rocks deposited in tectonically-active
basins.

Department of Petroleum Engineering, Heriot-Watt University 25


In a very broad sense, tectonism (formation of large-scale structures, such as mountain
ranges and adjacent oceans) usually produces places of uplift and places of subsid-
ence. If the uplifted region becomes eroded, the resulting debris tends to be
transported to the subsiding area. Thus there is an empirical relationship between
large-scale deformation and large-scale sedimentary bodies. As an example, the
North Sea basin has an abundance of sand-rich sediments that characterise the Triassic
and Jurassic rocks. This clastic material is directly related to the rift-tectonics that
initiated this basin. Later, mud-rich sediments characterise the younger, post-rift
deposition within this basin (but there are additional, limited pulses of sand related to
younger tectonic events).

3.2 Unconformities
In the preceding section, we noted that tectonism can cause rocks to be uplifted and
therefore subject to erosion. If allowed to continue, erosion produces horizontal
surfaces (mountains become plains). This process will truncate rock layers that are tilted.

Later in time (perhaps very much later), the horizontal surface, and the rocks that lie
beneath it, may once again subside and become a site of deposition. The new rock
layers will be essentially horizontal when deposited, and thus there will be an angular
difference between the younger rock layers and the older, truncated layers that lie
below the old erosion surface (Fig. 33). We use the word unconformity to refer to such
a surface. An unconformity (which is just a boundary) represents missing geological
time. The time is “missing” because there is no depositional record of what happened.

Create Structure

Erosion

Angular
Unconformity
Deposit New Layer(s)

Figure 33
Development of an
unconformity

Because there are other ways that unconformities can occur, we actually call the
situation in the previous paragraph an angular unconformity. Another type of
unconformity is a heterolithic (“different rocks”) unconformity. When an orogenic
belt is uplifted an eroded, the intrusive rocks and metamorphic rocks that characterise
the middle crust become the uppermost rocks in that area. If later deposition occurs,
the sedimentary rocks above the old erosion are very different from the crystalline
rocks below. This situation is found within most basins, where sediments lie above
“basement”.

26
Structural Geology 4
3.3 Intrusions (Dykes / Sills), Hydrofracture, Veins
Molten rock can be injected into pre-existing rocks, forming an intrusion. Here, we
are primarily concerned with intrusions that occur in thin sheets (Fig. 34). These are

yyyyy
;;;;;
called dykes (when they cut across the layering; because layering is usually more-or-
less horizontal, dykes are sub-vertical) or sills (which are more-or-less parallel with

y
;
the layering; sills usually have shallow dips). Dykes and sills may have thicknesses

;;;;;
yyyyy
ranging from a few 10s of cm to kilometres. Their areal extent can be anywhere from
a few square metres to many thousands of square kilometres.

;;;;;
yyyyy
;;;;;
yyyyy
;;;;;
yyyyy
;;;;;
yyyyy
Layers

;;;;;
yyyyy
yyyy
;;;;
Lifted Up

Figure 34
Sill

;;;;;
yyyyy
A dyke crosses the layers. A
sill is intruded parallel to
the layering (but has a
feeder dyke)
Dyke Feeder Dyke

These features are of interest because of their mode of formation. They do not fill a
pre-existing void; instead, they make space for themselves. The pressure associated
with the molten rock overcomes the least compressive principal stress (σ3), and the
sharp tip of the intrusion serves to concentrate stresses. In combination, these two
effects cause a crack (Mode I) to open in the host rocks, and the molten rock flows
forward into the space. Dykes and sills that are very thick are the product of continued
injection, and the subsequent widening of the space. Because these injection features
represent new “external” material, they result in an increase in the bulk dimension of
the rock mass in the direction that is perpendicular to the intrusion. Sometimes, there
are “swarms” of dykes - many, many individual dykes that have similar strike - that
are associated with major crustal extension.

The mechanical process of intrusion is essentially identical with a fracturing process


called hydrofracture. Natural hydrofractures may exist, but the term is usually
applied to features deliberately created to stimulate a well. In this situation, high-
pressure fluids are pumped into a rock layer, with the fluid pressure being sufficiently
high to cause fracture of the rocks. Usually, the fluid is pumped along with solids that
serve as proppants (glass beads or grains of sand). This is necessary because, unlike
a dyke, where the intruding fluid continues to occupy the crack, the aqueous fluids of
a hydrofracture will dissipate into the rock, allowing the newly-formed crack to close.
In order to make the new crack a suitable conduit for increasing the flow of
hydrocarbons, the proppant is a critical component.

Department of Petroleum Engineering, Heriot-Watt University 27


Elevated fluid pressure can create an effective state of stress that is capable of causing
rocks to fracture - even though the framework (“solid”) stress is below the level
necessary to cause deformation. In addition, fluid pressure affects the stability of pre-
existing discontinuities - such as fractures that are already present in the rock mass.
(see Appendix) This knowledge is exploited in drilling, where the leak-off test is used
to estimate the magnitude of σ3. Data obtained from such tests can be used to establish
a “fracture gradient”, which is the plot of wellbore fluid pressures at which fluid loss
will occur due to opening of the void spaces (Fig. 35). As a VERY rough rule of thumb,
the fracture gradient is approximately 0.8 of the lithostatic gradient (which is taken to
be 1.00 psi/ft).

Pressure

Lithostatic Pressure

Leak-Off Pressure

"Seal"
Depth

Fluid Pressure
(showing overpressure
below seal)

Figure 35
Hydrostatic and lithostatic
profiles with an
Hydrostatic Pressure overpressure zone and leak-
off data.

Tectonic fractures (frequently, the extension fractures, or joints) may be filled or


partly-filled with minerals (e.g., calcite or quartz). Such features are called veins.
Veins are significant because they indicate that the rock mass was dilated (its volume
was expanded).

There are two ways that veins can form. In the first, a fracture is opened slightly.
Fluids then circulate through the open crack, leaving behind minerals that were in
solution. In the second type, there is very little opening at any time. Circulating fluids
repeatedly deposit minerals that fill the small opening, and then these minerals are
cracked to re-form another small opening, with continued fluid movement. The
minerals that are deposited form fibres that are elongated in the direction of crack
opening. The process is known as “crack-seal”. It is more common in high-pressure/
high-temperature environments (metamorphism). The crack-seal process gives a
record of actual motion, which is usually not known in the more common types of veins.

3.4 Inversion
In basins, structures are important in terms of creating space within which new
sedimentation can occur. In many basins, listric normal faults, and associated growth
faulting, are very common. If at a later time, the basin is the location of a different type
of tectonic event (such as shortening), it is possible that the previous growth faults may

28
Structural Geology 4
become re-activated as thrust faults. There would be a thickened rock succession in
the hangingwall of the new thrusts, as well as, potentially, a change from reverse to
normal motion on the fault surface (depending on the magnitude of the displacements).
This process is called “inversion”. It is not certain if the tectonics works in exactly
this fashion, or if there is a different “true” interpretation that merely appears to be the
one described here. There is considerable debate about this notion, but it is important
to know what is being implied by the discussions that may occur around you.

4. PRACTICAL STRUCTURAL GEOLOGY

4.1 Tilting
Most of the processes that form structures result in tilting of the rocks. (As noted in
the Appendix, rigid-body rotation is a component of deformation.) Surprisingly, this
aspect of deformation is often not emphasised in descriptions of Structural Geology.
The surprise factor is because the tilting of rocks is almost essential for migrating
hydrocarbons from their source region to any available traps.

Because hydrocarbons are buoyant (relative to aqueous porefluids; refer to the


Chapter on Petroleum Play), there is potential energy available to cause them to rise.
What is needed is a pathway for them to follow. A tilted carrier bed (nominally, a rock
layer that has characteristics similar to reservoir rocks) can provide such a pathway.
Tilting is also important for a special type of trap. Stratigraphic traps occur if
hydrocarbons accumulate in a reservoir rock unit at the point where that rock unit
interfingers with other, non-reservoir rocks. The general view is that the hydrocar-
bons gained access to the carrier/reservoir rock layer, moved upwards along it (the tilt
component), and became trapped due to the termination of that particular lithofacies.

4.2 Trap Shapes


The primary rationale for the study of Structural Geology is that structures form traps
for hydrocarbons. Each trap is unique, with its own combination of rock types, the
stratigraphic arrangement, the geometry of the rock layers, the reservoir properties,
and the timing of structural events in relation to the migration of hydrocarbons.

The most important point to extract from this section is that structures produce
differences in elevation (Fig. 36). This difference can be produced by flexure of the
layers (into an anticline), or by faulting. Rotation (tilting) is frequently observed in
traps, but traps can be formed where the rock layers remain horizontal. Flexure can
be produced by faulting, by buckling, by diapirism, and by differential compaction of
underlying rocks (e.g. sand compacts less than shale, so a sand “pod” will produce a
bump in the overlying rocks following compaction).

Department of Petroleum Engineering, Heriot-Watt University 29


Elevation Differences Produce Trapping Opportunities

Flexure Faulting

Figure 36
A range of trap shapes
sand
produced by structures

Not all trap-shapes result in petroleum reservoirs. If hydrocarbons have not migrated
to the trap, or if the timing of formation of the structure is later than migration, no
reservoir is created. In other cases, the failure to accumulate hydrocarbons is a
consequence of the lack of a seal. For example, a seal may be removed by erosion of
the crest of a fault block. In other cases, deformation may impair the integrity of the
seal.

4.3 Fault-Zone Properties


Faults are often surrounded by a zone of deformed rocks. These damage zones may
be a few 10s of cm wide, or they may involve a hundred metres or more of the
hangingwall and footwall rocks. Fault zones can be characterised by permeability
impairment, or enhancement, or both. Often fault zones are sites for fluid flow in the
subsurface, or they may perhaps be the boundaries of regions of fluid flow. As a result,
over geological timescales, there are many opportunities for the permeability of faults
to become further enhanced or impaired through diagenetic dissolution or precipita-
tion events.

As noted in the Appendix, many rocks yield by dilatant fracturing. The fractures
produced this way tend to be conduits for fluid flow. If a fault damage zone consists
of fractures of this type, there can be a permeability enhancement - generally aligned
with the orientation of the fault. Other rocks, or perhaps the same rocks that are
deformed under different conditions, can yield by compactant modes of failure. For
example, in porous sandstones, shear fractures and small faults are characterised by
arrays of small-scale (up to about a mm in width) slip surfaces that have offsets of a
few millimetres to perhaps a few centimetres. These features can be observed in core
as well as outcrop, and they have been produced in the laboratory. They are called
cataclastic slip bands (CSBs), or granulation seams, or microfaults. In the usual case,
CSBs are characterised by breakage of the original detrital grains (and the limited
cement that is present in the rock), so that there is a grain-size reduction within the zone
leading to a permeability decrease (Fig. 37).

30
4

yy
;;
Structural Geology

;y;;
yy ;y
yyyy
;;;;;;
yy
;;;
yyy
;;;;
yyyy
;
y ;;
yy
;
y
yy
;; ;;
yy 0.1m 0.002 m

yy
;;
;yyy
yy
;;;
yyy
;;;;
yyyy;y;
;; ;;
yy
0.1m

;;;
yyy
;;;;
yyyy
y;yy
y;
;; ;;
yy
Figure 37
Granulation seams, or
cataclastic slip bands, in
clean sandstones Outcrop Core Thin Section

The granulation process can result in a reduction of permeability by an order of


magnitude (or more). Preferential cementation may occur at the granulation seam
(because of the reactivity of the fine-scale particles, and because of differences in
fluid-flow characteristics), reducing the permeability even further (by several orders
of magnitude). It is important to note that these structural phenomena occur most often
in clean, porous sandstones (nominally, excellent reservoir rocks) where they can
have a detrimental effect on the reservoir quality. In two-phase flow, the fine pore
throats associated with these networks can lead to high residual oil saturations because
of capillary trapping.

4.4 Fault Sealing


Because faults do not have constant displacements, there is a progressive variation in
the footwall and hangingwall rocks that meet at the fault plane. At the edges of a fault
surface (its tips), a rock layer is effectively continuous, because fault displacement is
zero. However, as the fault displacement increases (towards the middle of a fault
surface), rocks from different levels, higher and lower in the stratigraphic succession,
are juxtaposed due to the fault movement. If some of these rocks are capable of sealing
hydrocarbons, while others are reservoirs, it becomes a non-trivial exercise to predict
if trapping will occur along the fault. A graphical method to assist this problem is
known as the fault juxtaposition diagram (Fig. 38). This diagram projects both
footwall and hangingwall rocks onto a 2D view of the fault plane. Visual inspection
allows one to estimate the extent of sealing and communication across the fault
surface.

Shaded = Reservoir, White = Seal (Shale)


Grey-on-White Overlaps = Sand / Shale,
Grey-on-Grey Overlaps = Sand / Sand
Fault Throw = Offset of Layers

Figure 38 Fault tip

Example of fault
juxtaposition diagram Distance Along Fault

Department of Petroleum Engineering, Heriot-Watt University 31


A related topic concerns shale smearing along faults. The principal idea is that rock
successions with higher percentages of shale are likely to produce clay smears along
the fault plane as a consequence of fault motion. This is because shales are often very
weak and are easily deformed (Fig. 39). Techniques now exist to calculate a parameter
called the shale gouge ratio (SGR). This number represents the proportion of shale that
has moved past any point, expressed as a percentage of the total displacement on the
fault. Rock sequences with high shale fractions, and larger fault displacement,
produce higher shale gouge ratios. The numerical value is locally calibrated against
experience: some reservoir rocks contain hydrocarbons, while others have only water,
and the value of SGR that separates these cases is used to predict other trapping

yyyyy
;;;;;
situations in the local area.

;;;;;
yyyyy
;;;;;
yyyyy
;;;;;
yyyyy
Figure 39
Shale smear in a sand-
shale sequence

4.5 Overpressure
A condition of overpressure is said to occur when the fluid pressure at depth is higher
than the “expected”, normal hydrostatic pressure (e.g. Fig. 35). There are a number
of mechanisms that can cause this situation to occur:

• Dis-equilibrium compaction — porewater is unable to escape from low


permeability (usually shale) rocks during subsidence due to burial that is too fast
• Source rocks generating hydrocarbons
• Uplift of sealed units that have been buried (and which acquired high pressures there)
• Oil or gas columns
• High pressure generated elsewhere is communicated to the site

For rocks that are undergoing compaction, overpressure stops the compaction
(deformation ceases, or slows). For consolidated rocks, overpressure can create
fractures, or cause existing fractures to open, or perhaps to slip. Thus, there are
significant impacts on rock mechanics whenever overpressure occurs. Overpressure
is a dynamic phenomenon that can be found in nearly every basin.

High pore pressures cause difficulties in drilling, and if they are encountered without
proper precautions, a blowout can occur.

4.6 Stress-Sensitive Reservoir Behaviour


The existing (in situ) state of stress in a reservoir may very well be different from the
stress state(s) that existed when the reservoir was created, and when it acquired its
fracture suite. Knowledge about the in situ stress state can be derived in several ways.
During drilling, the state of stress causes part of the borehole wall to be unstable, and

32
Structural Geology 4
hence to be lost during drilling operations. The resulting oval borehole shape can be
measured with four-arm caliper tools, and from the direction of the long dimension,
the horizontal maximum stress direction can be inferred. Hydrofracturing, and packer
tests, can provide additional information.

There is an empirical observation that the in situ stress state affects the production of
a reservoir. Fractures, and perhaps pores, are subject to distortion due to the in situ
stress. Fractures that are parallel to the σ1 - σ2 plane are most favoured for being open
(because their opening is resisted by the smallest stress, σ3). This directional character
of flow can be used to infer the stress orientation. Similarly, it is possible to develop
reservoir management plans that exploit this phenomenon (Fig. 40). The anisotropy
of seismic shear waves may also be used to detect the preferential opening of fractures.

Figure 40 N
Example of a fractured Well
reservoir in which the FMS Fracture
direction of the “open” Orientations

fractures, as evidenced by Directional Permeability


from Interference Tests
interference tests, is used to
infer the σ1 orientation 0 0.5 1.0 mi C.I. = 200 ft

Another type of stress-sensitive behaviour concerns rocks that are “under-com-


pacted”. Loosely-consolidated sandstones are sometimes observed to have very high
porosities. When the reservoir is produced, the pore pressure falls, and the under-
compacted rocks then proceed to compact. During this process, loose grains of sand
can migrate to the wellbore, where they are produced along with the fluids. Chalk
reservoirs can also exhibit such production-induced compaction. The compaction can
be expressed at the earth’s surface by subsidence, which can be a serious problem -
for surface production facilities, including platforms, and for any other human
activities.

4.7 Fractured reservoirs


A fractured reservoir is one in which naturally-occurring, dilatant, conducting
fractures (not CSBs) have a significant effect on the producing characteristics of the
reservoir — usually through increased permeability. Nelson (1992) gives four types
of fractured reservoirs (Fig. 41):

Department of Petroleum Engineering, Heriot-Watt University 33


• TYPE I: Fractures provide the essential porosity and permeability to the
reservoir. These reservoirs often deplete rapidly and are not very economic.

• TYPE II: Fractures provide essential permeability. Matrix porosity supports


the fracture flow to maintain performance and provide sufficient reserves.

• TYPE III: Fractures add to the permeability. Fractures enhance the reservoir
performance, significantly improving the otherwise poor-quality reservoir.

• TYPE IV: These are normal matrix reservoirs where fractures may introduce
some anisotropy or compartmentalisation. Fractures of some sort are to be
expected in all reservoirs.

All
Fractures
100 %
Fractures

II
% of Total Permeability

III

Figure 41
IV
Crossplot showing the
100 %
Matrix relative contributions of
All 100 % 100 %
Matrix Matrix % of Total Porosity Fractures matrix and fractures in of
fractured reservoirs (after
Nelson, 1992)

Effective flow properties in a realistic fractured rock mass depend on the geometry and
intersections of fractures belonging to multiple sets. Fracture porosity is usually <1%
in reservoirs. Fracture permeability can range from a few mD to several Darcies.
There are two useful expressions for estimating fracture permeability and porosity, as
a function of fracture aperture and spacing, for sets of parallel fractures:

a3 .
kf = 8.35.109
d
a .
ϕf = 100
a+d
where
kf = fracture permeability (mD)
ϕf = fracture porosity
a = fracture aperture (cm)
d = fracture spacing (cm)

34
Structural Geology 4
4.8 Structural Observations and Interpretations from Typical Data
Seismic reflection data (see Chapter 5) represent the primary information source on
the structural forms that are present in petroleum basins. After the seismic reflections
are correlated with specific rock layers, it is possible to determine the strikes and dips
of those layers (in a grid of 2D seismic lines, this process can be error-prone!). Where
the orientations change, we can expect the rocks to be damaged. Where layers are not
continuous, we can interpret faults. Where seismic attributes change, we can
(possibly) infer a change in fracture intensity, etc. Because faults are often steeply-
dipping, they are usually not well imaged on seismic data. Instead, faults are
interpreted as the junction between regions of the reflection section where layers are
observed to be intact. Very good seismic data can be subjected to fault-plane mapping,
which identifies faults ‘directly’.

Geological maps and cross sections are also used to identify structural features. These
data forms are particularly valuable for determining the structural style of an area.
Subsurface maps and cross sections (see Chapter 7) are also used to interpret the style.
However, a caution is in order. Many maps and sections are made to address a specific
need. They may not be appropriate for another need. For example, a regional-scale
structural-contour map produced from widely-spaced 2D seismic lines may reveal
major anticlines (exploration stage), but that same map is probably unsuitable for
determining curvatures, and hence fracture intensities (production stage). Wherever
possible, the history of a map or cross section should be investigated before its
information is used.

Natural fractures can be identified in cores by visual inspection (they have to be


distinguished from drilling-induced fractures), on logs (image logs are invaluable), by
loss of circulation during drilling, and by anomalous flow rates during production or
drill-stem tests (production logs can identify flowing fracture zones). Fracture zones
can sometimes be identified on seismic records (because fractures can change the
physical characteristics of the rock). However, evaluating fractured reservoirs
requires the integration of sparse, often non-quantitative data from many disciplines.

Faults can be identified on wireline logs by constructing correlation panels, and from
these, determining missing or repeated portions of the stratigraphic succession. This
effort is made considerably more difficult if the rock layers are dipping, or if the
wellbore is highly deviated. This is because the true thicknesses are stretched on the
wireline logs. Dipmeter logs can be used to produce True Stratigraphic Thickness
corrections to the log suite, so enabling an improved correlation and identification of
faults.

4.9 Balancing
The principle of structural balance is merely a geologically-phrased version of the
classic law of physics: the conservation of mass. To paraphrase: structural balance
means that the materials that existed before deformation are still present afterwards,
although they are re-arranged (and possibly may have left the local area!). In practice,
balancing is used to assist in judging the validity of structural interpretations of
deformed rocks. Since we always lack complete exposure, alternative ways of
interpreting the missing (unseen) parts of structures can be compared through this
approach. This is done by attempting to restore the deformation, and then assessing

Department of Petroleum Engineering, Heriot-Watt University 35


if the pre-deformation geometry looks plausible. A key point in balancing is the
identification of a pin line: that is, a place where there has been no movement.

Although structural balance sounds like a panacea, in practice, there are many
potential pitfalls. These are related to the simplifications of the concept that are
routinely adopted (such as: line-length balancing, area balancing, etc). For our
purpose, we can accept that balancing has the potential for assisting in developing
geometric interpretations of structural features, and in deriving interpretations of the
development history of those structures. However, the actual techniques require
specialist training, and are not presented here; Petroleum Engineers should consult the
geological staff to assist in any balancing work.

4.10 Structural families


The study of Structural Geology is merely one aspect of efforts to understand the
complete geological history of an area. In the context of Petroleum Engineering, that
history includes the operation of the Petroleum System, including the deposition of
suitable source rocks, seals, and reservoir units, and their deformation into trap
geometries at a suitable time to catch the migration of hydrocarbons. This section
focuses on the idea that the geometries and kinematics of deformation often produce
characteristic patterns. In other words, there are common associations of structural
types. After a given style is identified (in terms of associations that are “typical”),
reasonable predictions may be made of structural forms in locations where information
is sparse.

The idea of repeated associations is sometimes referred to as structural families


(sometimes also called structural styles). Although everyone accepts the notion of the
existence of structural families, there is no universally-accepted list of them. The
reason for this is that the definition of a style is often linked with ideas concerning the
process of formation of structures. For a decade or so, most types or styles are
“explained” via some plausible model, but new research seems to uncover difficulties
with these ideas, and new models of formation are then created. Therefore, there is
always a degree of controversy concerning the origin of some structures (depending
on the point in this idea/revision cycle), and it is difficult to tell students “the” answer.
The succeeding sections give (some of) the characteristics of the primary groups of
structural families.

Fold/thrust belts
Fold and thrust belts are long (100’s to 1000’s of km), curvi-linear zones of
deformation within which there is substantial shortening (typically, the belt is only
half of the original width). These belts usually occur in zones where major plates
converge. In a major orogenic system, there may be two sub-parallel thrust belts: one
on each ‘margin’ of the orogenic belt. Some key characteristics of fold/thrust belts are
(Fig. 42):

• Asymmetric folds, truncation- and buckle-folds

• Décollement (detachment-style) thrust/reverse faults

• Good strike continuity of structures, moderate dip and depth continuity

36
Structural Geology 4

• Generally forms an arcuate belt

• Often a narrow(ish) zone of deformation located at a continental margin (or a


former margin)

• Progressive intensification of structural complexity and metamorphism from


the externides to the internides (foreland to hinterland)

Fold and Thrust Belt

Cover

Basement

Horizontal reference

Internal 100's of Km External


(Hinterland) (Foreland)
Faulted, and even re-mobilised, basement Fault-associated folding
Re-folded metamorphosed cover rocks Buckle folding

Long-travelled thrust sheets

Figure 42
Structures typical of fold/
thrust belts
;yy;y;y;y;y;y;y;y;y; Nappes

Wrench provinces
Wrench provinces in continental terrains can be broad (1000’s of km in each direction)
regions produced by the lateral movements of major plates. In oceanic regions,
wrench zones typically are more narrow (100’s of km), but possibly very long.
Because the wrench faults are usually not ‘perfect’, but instead are irregular (jogs,
changes in strike, etc), there is a considerable degree of associated deformation
necessary to allow rock masses to move past one another. Some key characteristics
of wrench provinces are (Fig. 43):

Department of Petroleum Engineering, Heriot-Watt University 37


• Intermixing of folding and faulting

• Fault sense (normal to reverse) and fold shapes change rapidly

• Variety of trends (refer to “shear zones” in Appendix)

• Often a strong localization of deformation, with intervening regions of little or


no damage

• Vertical and horizontal predictability is poor (even though there may be a


strong preferred orientation of fault trends)

• Sometimes adopted as a style whenever there is great complexity, even though


definitive evidence of wrench movements may be lacking

• Regional context is very important in identifying this style correctly

• Contains examples of structural forms of all types

Releasing bend Restraining bend Dilational jog Restraining jog

Wrench Structures
Map expression of wrench zone
Mountains
Sediment supply
Fault continues
Do aul
f
wn ting

monocline
dr or

Depositional
op fle
pi xu

area
ng re

Slide block
via

Normal faults Volcanics

Oblique - slip faults


Away Toward

Fault continues Basement


Sediment supply
Coarse debris
Cross section of flower structure along wrench fault
Map of local depocentre (releasing bend or dilational jog)
Fault
Anticline often on echelon continuing

Reverse fault possibly oblique slip

Fault tip

Fault tip
Erosion =
sediment supply Figure 43
Block diagram of local uplift (Restraining bend or jog)
Structures typical of
wrench terranes

Rifting and extension


Rifting and extension represent the lateral stretching of the crust, caused by its own
self-weight, when the lateral constraints are removed (sometimes by stretching of the
mantle underneath caused by a rising plume). Rifting and extension often occur at the
sites of previous orogenic shortening (where the crust was thickened in the shortening
phase) as these later spread sideways and return to normal crustal thickness. Some key
characteristics of rifting and extension are (Fig. 44):

38
Structural Geology 4
• Tilted fault blocks, often with considerable asymmetry

• Syn-tectonic deposition, and possibility of major and/or rapid changes in


depositional setting (non-marine to deep marine, and possibly back to non-marine)

• Listric faults, and abundant smaller-scale faults

• Moderate strike continuity, but abrupt terminations (transfer zones)

• Moderate dip continuity, but variable depth predictability (location of


décollements)

• Volcanics

Half-Graben
Half-Graben tilted right
tilted left

Rifting Structures
Transfer fault

Detachment fault

Rollover anticline "Fan" of bedding dips in


Transfer fault syn-tectonic sediments

Antithetic
fault
Faulting/folding of
List ric normal fault shallow layers above
fault-block edges

Headwall
Core complex Extreme tilting

Metamorphic rocks rise


Figure 44 and warp detachment fault
Extreme Rifting Leading to Detachment Faults and Metamorphic Core Complex
Structures typical of rifting

Diapirs and growth structures


Diapirs and growth structures often occur together. They are typically found in
passive-margin settings (a passive margin is produced when a rift is successful at
separating a plate). The evaporites that produce the major diapirs are deposited in the
early rift-stage. They become mobilised and affect deposition and deformation
thereafter. Major growth faults are often located on the basin-ward flank of major
diapirs. Modern interpretations of the Gulf of Mexico margin suggest that salt diapirs
can move laterally through portions of the basin, leaving behind complex ‘welds’
where sedimentary sequences of differing ages are juxtaposed across the former site
of the salt tongues. Some key characteristics of this style are (Fig. 45):

Department of Petroleum Engineering, Heriot-Watt University 39


• Typically present in passive margins, but can occur elsewhere

• Depositional thickness changes, rollover structures

• Synthetic and antithetic fault sets

• Predictability is proving less good than was thought only a short time ago

• Growth faults often on flanks of diapirs

• Diapirs of both salt and mud activated by density differences

• Toes of major sheets may show contractional structures that are contemporaneous
with headwall extension

Wedge of passive-margin sediments


which contain these other structures Salt tongue

Oceanic crust
Continental crust
Crustal-scale view
0 50 km "weld"

Growth fault system

Toe of slide sheet


50-100
km
Thrust faults
Overpressed shale and buckle folds

Figure 45
Structures typical of
Salt
Salt “passive margins”

“Drape” folds / Block faulting


An interesting class of structures that is often not separately considered is that of
“drape folds” and block faults. In these features, a deep-seated, high-angle fault (in
basement rocks, or in deeper portions of the basin) causes bending of the overlying
layered rocks at shallow structural levels. The kinematics of this structural style have
led to heated arguments. This is because there is a discrepancy between the line-
lengths of the layered rocks that are folded, and the line-lengths of the faulted rocks.

40
Structural Geology 4
In other words, these structures do not follow the rule of ‘structural balance’. (What
is actually going on is that there are significant volume strains that affect line-lengths
and areas of cross sections. Such strains are not, typically, included in balancing
methods. In addition, there is a considerable lateral motion associated with flow of
ductile rocks.) The transition from fault to fold is accomplished via the ‘flowable’ unit
within the sedimentary succession.

Structures of this type are actually quite common. They characterise the style of intra-
cratonic basins and many shallow-shelf and platform regions. There is a growing
appreciation of this style in rift environments. This style probably occurs in many
other places, but the structures may have been interpreted to be some other structural
type. Some key characteristics of this style are (Fig. 46):

• Wide range of fault orientations

• Depending on cause of faulting in basement, lateral predictability varies as per


above styles

• Depth predictability is generally good

• Key element is unit of ductile/weak rock between “basement” and layered


rocks of the “cover”

• This style is probably ubiquitous throughout the “stable” regions of continents


as intra-continental basins form and evolve

Examplesof Basement Fault-Block Patterns Basement / Cover (Sediments) Relationships


attenuation of strong layer
possible new layers
deposited during fault strong
movement

strong
basement
weak
Normal Fault

nearly continuous
strong layer strong

weak
basement
Vertical Fault
Figure 46
Structures typical of stretching and strong
faulting of strong
platform areas layer

weak
basement
Reverse Fault

Department of Petroleum Engineering, Heriot-Watt University 41


5. FURTHER LEARNING

Exercises

1. Collect together sketches and/or photographs of faults from: local outcrops, the
geological data you have around your office, or published examples in the literature.
In each case identify the hangingwall, footwall, sense of displacement, and throw. If
possible consider the 3-D aspects of the features. Try to classify the type of faults, and
explain their structural settings. You should have between 10 and 20 examples at a
range of scales.

2. Collect together sketches and/or photographs of fractures/CSBs from: local


outcrops, the geological data you have around your office, or published examples in
the literature. In each case identify the fracture sets, any fracture assemblages that
may be present, and seek to determine the directions of the lengthening and
shortening produced by the fracturing. You should have between 10 and 20 examples
at a range of scales.

3. Collect together sketches and/or photographs of folds/flexures from: local outcrops,


the geological data you have around your office, or published examples in the
literature. In each case identify the shape of the flexure, any associatd fractures or
faults, and seek to determine the tectonic setting. You should have between 10 and
20 examples at a range of scales. If the flexures are related to faults, think about the
sequence of formation.

4. Read several papers that describe a field. Identify the structural elements that are
important to create the trap, or that have a significant effect on production. Do all of
the features seem compatible within interpretated structural setting?

42

You might also like