Vous êtes sur la page 1sur 316

Structural Geology

Structural Features of Non-Tectonic Origin


General Introduction
At any point in its history, a sample of rock can be said to have "structure." Sedimentary rock,
whether highly contorted or flat-lying and undisturbed, is characterized by fabric, texture, and
tectonic setting.
In a very basic sense, the structural features of rocks can be divided into those that preceded the
onset of deformation-called sedimentary structures-and those that were the result of strain
imposed by external forces-structures of tectonic origin. Sedimentary structures are formed
during (1) deposition and (2) diagenesis that is uninfluenced by diastrophism. Tectonic structures,
in contrast, can originate both during and after diagenesis and are, of course, the major subject of
structural geology.
The relationship between tectonism and diagenesis, however, is often complex, which sometimes
makes the distinction between the two types of structures difficult or artificial. Deformational
forces often directly influence the lithification of sediments, particularly in areas where deposition
occurs in close proximity to growing mountain belts. Observed structures may exhibit an
unbroken progression from sedimentary to tectonic in origin.
Notice, however, that we use the term "external forces" to refer to the origin of tectonic features.
This, it turns out, is a highly useful simplification that justifies the distinction between the two
structural types. The idea that tectonic stress is applied from without onto relatively passive
sedimentary strata helps identify some sedimentary structures as both markers and tracers.
These can be used by geologists to determine the original relationships within a highly folded and
faulted lithologic sequence. They also frequently provide qualitative indications of deformational
intensity. Sedimentary structures are thus invaluable to geologists working in areas such as overthrust terrains, where great thicknesses of sedimentary strata are overturned or chaotically
disturbed. For example, a number of sedimentary features, particularly those generated by
current action, reveal which way is stratigraphically up in a section-the so-called younging
direction.
With respect to structural studies, detailed knowledge of sedimentary structures is most useful to
the field geologist or to the explorationist who may need to field-check interpretations. In normal
subsurface work, which is restricted to the "artificial exposures" provided by recovered cores, it is

sometimes difficult to distinguish sedimentary from tectonic structures. Generally, however, no


amount or orientation of deformation will produce the variety of depositional features commonly
found in sedimentary rocks. Therefore, only a brief review of the more common types is given
here.
A more complete overall discussion of depositional textures and features useful for the structural
geologist is offered by Hills (1971) and Reineck and Singh (1975).

Stratification
Layering is the most widespread structural feature of sedimentary rocks. It generally results from
some change or set of changes in the conditions of deposition-e.g., grain size, sorting, or
composition; current strength and direction; rate of sedimentation; water chemistry. It can also be
caused by postdepositional variations in cementation, compaction (thus packing), and fluid
composition; hence its origin is a complex, only partially understood phenomenon. For the
interested student, Allen (1970), Pettijohn, Potter, and Siever (1972), Blatt, Middleton, and Murray
(1972), and Reineck and Singh (1975) each offer discussions on the detailed theories of the
origin and significance of sedimentary layering.
While determination of true bedding is always critical, recognition of true bedding in highly
deformed areas is not always a simple matter, particularly in core samples ( Figure 1 , Illustration

of how original depositional layering can be obscured by secondary, tectonic layering ).

Figure 1

Diastrophism very often creates a secondary planar fabric. Penetrative fracturing in essentially
unmetamorphosed rocks is only one example of tectonic layering that can obscure bedding.
Thick, relatively homogeneous shale sequences or an originally massive sandstone may show no
primary layering for hundreds of vertical feet. In studying outcrops, logs, or cores in a structurally
complex area, the explorationist must make a constant effort to understand the origin and
significance of the layering that characterizes the strata under inspection.
In regions where the extent and influence of tectonism are unknown or questionable, we should
remember that original sedimentary dips rarely exceed 30 to 40, except in a few cases, such as
forereef talus, slumping, and local convoluted bedding. Dipmeter patterns are often very helpful in
determining true from apparent bedding, particularly in sandstone formations.
Overturned layering may go largely unrecognized without the use of additional sedimentary
structures that can distinguish the tops of beds and thus younging direction. Many of these,

shown in Figure 2 ,

Figure 2

Figure 3 ,

Figure 3

Figure 4 (Common sedimentary structures useful for determining the direction of younging ),

Figure 4

Figure 5 (General appearance of layering disturbed by a slumping, and b convolute bedding.

Figure 5

Upper truncation of slump interval can sometimes be used to distinguish younging ), Figure 6 ,
and Figure 7 (Appearance and proposed evolution of load structures.

Figure 6

Figure 6 is a small scale vertical section through load casts in the Langdale Slates, English Lake
District.

Figure 7

Figure 7 shows the development of "ball and pillow" casts proposed by Kuenen, 1965. Vibration
causes rupture and sinking of overlying sand into underlying, water-saturated muds) are currentgenerated features and are thus applicable to sandstones and clastic carbonates.

Cross-Bedding and Graded Bedding


For carbonates, cross-bedding is often the most useful for topping, where it can be found. Fossils
preserved in growth position are relatively rare. Shells, even when showing a predominant
orientation within a single layer, can never be entirely trusted. Though their hydrodynamic
preference is undeniably convex-downward (like an upside down bowl or plate), the high water
energy of shallow marine environments-frequently punctuated by storms-makes their final resting
orientation often uncertain.

In sandstones, crossbedding and graded bedding are the most reliable and widely used
indicators of bedding tops. These textural features are preserved and usually identifiable in
intensely deformed rocks and can often survive low- to medium-grade metamorphism. Graded
bedding is especially found in turbidite sandstones but can also occur in detrital carbonates,
volcanic ash, and varved clays. More generalized fining-upward gradations of clastic grain size
characterize channels in many environments (e.g., point bars, deep-sea fans).
Other surface features (e.g., desiccation or "mud" cracks, raindrop impressions, rootlet and
paleosoil horizons) may also make excellent younging horizons but are relatively rare in outcrop
and extremely so in cores.

Sole Markings
These are features that occur along the interface between a sandstone, siltstone, or clastic
carbonate and an underlying shale bed. They are preserved in negative relief on the under-sides
of the coarse-grained beds, so that what were originally grooves and depressions on the irregular
shale surface are now seen as ridges and protrusions. Consequently, sole marks make excellent
indicators of younging. The impressions can be caused by a variety of current movements
(erosion; objects being dragged, scraped, bounced, or rolled across the depositional surface), by
animal tracks and burrows, or by sediment loading.

Soft Sediment Deformational Features


Unconsolidated, water-saturated sediment can be disturbed into producing deformational features
of several types. Some of these are nearly contemporaneous with deposition; others occur as a
result of progressive compaction.
Slumping, for example, is fairly common in areas where rapid deposition takes place on an
inclined slope, such as in deltaic or submarine fan/turbidite environments. The actual slumping of
material can be caused by several factors, but the most important of these is the buildup of
abnormally high pore fluid pressures within impermeable depositional layers.
Slumping, particularly on a larger scale, seems more often to result from sudden detachment
within shale units and the consequent movement downslope of overlying layers. This detachment
can occur on many scales, but generally produces chaotic, incoherent bedding (part a of Figure 1
). The upper portion of a slump is commonly truncated by overlying sediments; this can
sometimes be used as an indicator of the younging direction.

Figure 1

Slumping can be triggered by a variety of phenomena that introduce sudden amounts of energy
into an environment, e.g., earthquakes or storms.
Convolute bedding (part b of Figure 1 ) is a common feature seen in the deposits of certain
facies, such as braided streams. This type of soft sediment deformation differs from slumping in
that it is more coherent (single laminae can sometimes be traced for significant distances) and
because disturbances die out upward within the same bed. Sometimes this also is a younging
indicator.
Such structures are not well understood by sedimentologists, who have proposed a variety of
explanations for their origin (Kuenen 1953). One of these emphasizes the existence of sediment
(usually clay-rich) that is thixotropic, i.e., that loses internal strength almost instantly under
disturbance (for example, during an earthquake) but regains it when the disturbance stops.
Convolute bedding and slumping are not always helpful for identifying bedding tops; they can,

however, be mistaken for tectonically induced deformation and therefore must be distinguished in
outcrop and core study.
Load casts ( Figure 2 ,

Figure 2

Small scale vertical section through load casts in the Langdale Slates, English Lake District, and
Figure 3 , Development of "ball and pillow" casts proposed by Kuenen, 1965.

Figure 3

Vibration causes rupture and sinking of overlying sand into underlying, water-saturated muds) are
bulbous protrusions that record the sinking of sand or silt down into hydroplastic shale. In contrast
to other sole marks, they are nondepositional and are caused exclusively by compaction and
readjustment of unconsolidated material after burial. They can be distinguished from flute marks
by their lack of linearity. Like these, however, they are excellent top/bottom indicators.

Supratenuous Folds
Sometimes called "compaction folds," these are structures generally thought to be the result of
the warping of sedimentary layers by differential compaction ( Figure 1 , Supratenuous fold

developed over a reef). This can occur around a single pebble, a boulder, or an exotic block
(fragments of rock up to tens of kilometers in length that have broken free and slumped into
unrelated sediments).

Figure 1

Such folding also frequently develops in layers deposited over an irregular erosional surface, into
a channel, and over the flanks of carbonate mounds or barrier bars. Because it is caused by
preexisting, generally local features, supratenuous folding dies out upward in a lithologic section.
Many geologists have also used the term supratenuous to refer specifically to folds that develop
as a result of differential sedimentation and subsequent compaction over erosional highs and
lows. Such folds are characterized by thinning at the crests, gradual thickening along the flanks,
and maximum thickening in the troughs. This type of folding is often accentuated by subsequent
tectonic deformation; such structures introduce local heterogeneity that can attract and focus later
deformation.

Depositional Tilting
The exploration geologist analyzing structure should also be aware of tilting in depositional units
that show thinning toward a buried "high," or basin margin and corresponding thickening into a
"low." Such tilting occurs on many scales: it can reflect original depositional slope over a broad
area or the presence of a local topographic inhomogeneity, such as a fault scarp or graben. It is
usually accentuated by subsequent burial and compaction. So-cal led compaction monoclines are
relatively common features in sedimentary units that have buried fault blocks. These structures
dip off the up-faulted block into an adjacent low.

Measurable dips for depositional tilting are usually low (less than 15), but can sometimes reach
30 or more locally. In all cases, these dips need to be identified as nontectonic. Where regional
in extent, they should be considered original structural parameters.
1. What geologic processes could trigger slumping?
2. In what type of geologic province would you expect it to occur continually, perhaps on a
relatively large scale?

Examples of Recent Structural Movements


Introduction
Many of the most informative pieces of evidence for the manner in which rocks deform come from
the small increments of movement to which geologists have been witness. Much of geology, in
fact, has in recent years been reviewed in light of a compromise between the previously
antithetical hypotheses of catastrophism and uniformitarianism.
For example, major components of total displacement along fault zones are now thought to be
the result of significant seismic events. Such events were previously thought to represent minor
catastrophes whose contribution to overall offset was secondary to the more continuous, slow,
and imperceptible failure and slippage of rocks called "creep." Strong and shallow seismic
disturbances, such as those that occurred near Shen Shu, China, in 1556, Lisbon in 1755, San
Francisco in 1906, Tokyo in 1923, Chile in 1960, and Sherman, Alaska, in 1964, have been able
to create large and essentially instantaneous changes in topography. In addition, Richter (1952)
offers a number of other well-known examples of historically documented movements associated
with major earthquakes.
Figure 1

Figure 1

and Figure 2 (Charts comparing the rates of common sedimentary and tectonic processes in

various parts of the world) display a comparative range of rates for present-day tectonic activity.

Figure 2

These are also measured relative to the slowest and fastest rates of sedimentation, and rapid
denudation. Obviously, in terms of geologic processes, tectonic deformation is often very fast, as
much as an order of magnitude beyond most depositional rates. It should be apparent, then, that
diastrophism deeply influences deposition. Often, the character and distribution of syntectonic
deposits, such as alluvial fans, are direct indicators of how and where mountain building took
place. Analyzing such deposits, therefore, is very often a central part of regional structural
analysis.
In its broadest sense, diastrophism can be divided into two main types on the basis of rate and
intensity. Orogenic movements are those normally associated with mountain building. They occur
most often in long, relatively narrow regions where high rates of intense deformation are
concentrated. Such regions have been called mobile belts because of the obvious horizontal and
vertical movement the rocks within them have undergone; this term, however, is no longer
extensively used.

Epeirogenic tectonism, in contrast, affects much larger areas with slower, more undulatory
movements. Figure 1 and Figure 2 can be used to compare the rates of epeirogeny with those of
several orogenic and depositional processes.
Epeirogeny is assumed to have multiple causes. The two most frequently cited origins for
epeirogenic uplift are isostatic compensation (due to the removal or addition of a crustal load) and
thermal expansion in the upper mantle (e.g., that related to crustal rifting). Only occasionally does
it appear to be directly associated with major faults in the crust. Long-standing debate has
surrounded the problem of distinguishing the effects of epeirogeny on the sedimentary record
from those due strictly to changes in sea level. The discussion is very pertinent to structural
geology: the origin of important regional unconformities, for example, is tied directly to broad
crustal movements. Epeirogenic movement is common in recent geologic history. Data from
leveling surveys conducted over the past century have revealed that almost every continent is
characterized by relatively gentle vertical movements. As shown in Figure 1 and Figure 2 , these
are generally measurable in mm/yr. The most conspicuous present-day form of epeirogeny,
however, is glacial isostatic rebound, which takes place at very high rates (cm/yr) and is very
short-lived (probably less than 2-3 my) (Flint 1971). During the last 8000 years, glacial rebound
has caused as much as 300 m of uplift (38 mm/yr) in the Hudson Bay area of Canada, and 520 m
(65 mm/yr) in the Baltic Peninsula with a calculated 210 m still to be achieved (Flint 1971) in
central Sweden and Norway.
Soviet and East European geologists have derived the term neotectonics to describe recent
epeirogenic-type movements of ancient shield areas. On the basis of precise leveling data,
Mescherikov (1968) reported continual uplifts and subsidences of between 10-100 m at rates of
about 10 mm/yr.
In North America, leveling lines indicate that portions of the Appalachian Mountain system are
rising at about 6 mm/yr. Such movement has been related to intraplate readjustment due to
drifting of North America from lower to higher latitudes (Sbar and Sykes 1973). Simultaneous
subsidence along the Atlantic coastal plain (see Figure 3 , Rates of present-day vertical crustal

movements in the United States) is, however, apparently due to crustal loading and consequent
bending as a result of the postglacial rise in sea level (Bloom 1978).

Figure 3

More recently, however, the term neotectonics has come to be used in reference to both very
recent and present-day deformations of any type, orogenic or epeirogenic. This is particularly true
in the literature dealing with the Mediterranean and western Pacific regions.

The San Andreas Fault System


As shown in Figure 1 ,

Figure 1

Figure 2 (Charts comparing the rates of common sedimentary and tectonic processes in various

parts of the world),

Figure 2

and Figure 3 (Calculated annual rates of displacement along several major faults in California.

Figure 3

Length of arrow corresponds to amount of movement see vector scale) movement along the
San Andreas fault represents one of the most rapid rates of tectonism anywhere in the world. As
a structure, the San Andreas is better described as a fault system which connects, through
continental crust, two oceanic spreading centers, one (the East Pacific Rise) in the Gulf of
California, the other (the Juan de Fuca Ridge) off the Pacific Northwest coast of Washington,
Oregon, and northern California.
The southern portion of the fault system apparently began less than 15 my ago, but has
undergone possibly 300 km of displacement (2 cm/yr) since then. Most of this displacement has
evidently occurred within only the last S my. Present movement along the fault averages between
2-5 cm/yr (Wallace 1970). Quaternary river terraces have been offset several kilometers, and
Pliocene strata indicate 23 km of displacement in the earthquake zone. Maximum creep along the
fault has been measured at 10 cm/yr (Dennis 1972).

Horizontal offsets measured after the 1872 and 1906 earthquakes in the Owens Valley and San
Francisco areas were as high as 7 m. Such "catastrophic" movements apparently occur at semi
regular intervals of between 100 and 150 years (Wallace 1970). This indicates that total offset
along the fault results about equally from creep and from sudden displacement accompanying
earthquakes.
In addition, subsidence in the highly productive Ventura basin, just east of Los Angeles, has been
directly linked to slip along the San Andreas and is estimated to have occurred at rates of about 4
mm/yr.

The San Andreas Fault System


As shown in Figure 1 ,

Figure 1

Figure 2 (Charts comparing the rates of common sedimentary and tectonic processes in various

parts of the world),

Figure 2

and Figure 3 (Calculated annual rates of displacement along several major faults in California.

Figure 3

Length of arrow corresponds to amount of movement see vector scale) movement along the
San Andreas fault represents one of the most rapid rates of tectonism anywhere in the world. As
a structure, the San Andreas is better described as a fault system which connects, through
continental crust, two oceanic spreading centers, one (the East Pacific Rise) in the Gulf of
California, the other (the Juan de Fuca Ridge) off the Pacific Northwest coast of Washington,
Oregon, and northern California.
The southern portion of the fault system apparently began less than 15 my ago, but has
undergone possibly 300 km of displacement (2 cm/yr) since then. Most of this displacement has
evidently occurred within only the last S my. Present movement along the fault averages between
2-5 cm/yr (Wallace 1970). Quaternary river terraces have been offset several kilometers, and
Pliocene strata indicate 23 km of displacement in the earthquake zone. Maximum creep along the
fault has been measured at 10 cm/yr (Dennis 1972).

Horizontal offsets measured after the 1872 and 1906 earthquakes in the Owens Valley and San
Francisco areas were as high as 7 m. Such "catastrophic" movements apparently occur at semi
regular intervals of between 100 and 150 years (Wallace 1970). This indicates that total offset
along the fault results about equally from creep and from sudden displacement accompanying
earthquakes.
In addition, subsidence in the highly productive Ventura basin, just east of Los Angeles, has been
directly linked to slip along the San Andreas and is estimated to have occurred at rates of about 4
mm/yr.

Alaska
The 1964 Sherman earthquake in southern Alaska provided geologists with an excellent
opportunity to study the sudden displacement-and consequent damage-of a large seismic event.
Mud slides released by the quake were reported in the Anchorage area, and large vertical scarps
cut jagged walls through several small towns. Figure 1 shows the location and range of
movement, which took only minutes to complete.

Figure 1

Maximum uplift was as much as 13 m, while subsidence did not exceed 2.5 m. In addition to the
effects mentioned above, tidal waves were generated, and large sections of ocean floor were
uplifted from the sea along steep faults (Plafker 1965). The quake was a shallow-focus event
related to subduction of the Pacific plate beneath Alaska (note location of the Aleutian trench axis
in Figure 1 ).
Figure 2 generalizes vertical movements across the United States, as determined from leveling
data.

Figure 2

The figure makes it apparent that continents can be spoken of as stable only in the most relative
terms. Most of the tectonism responsible for these movements can be classified as epeirogenic.
Causes are variously interpreted to be related to (1) crustal adjustments that reflect irregularities
in the upper asthenosphere over which the North American plate is moving, and (2) to stress
patterns initiated by the changing curvature of the planet's surface as North America has
migrated to higher latitudes (Turcotte and Schubert 1982).

Red Sea
The Red Sea rift is a highly active zone of ocean spreading that continues to separate the African
and Arabian plates and to drive a major collisional event occurring approximately 1500 km to the
northeast between Arabia and Eurasia. ( Figure 1 , Generalized map and cross section showing

continental breakup along the Red Sea rift and collision in the Zagros region of southeastern Iran.

Figure 1

Numbers indicate total estimated separation in km between Africa and Arabia.) Rates for
rifting in the Red Sea have been difficult to determine, due to the complexity of movement (Lowell
et al. 1975). In general, this region provides ample evidence for considering rifting a product of
both lateral, pull-apart motion and longitudinal (i.e., subparallel to the rift itself) slip motion. Early
determinations for the rate of continental separation in this area are on the order of 1-2 cm/yr
(Vine 1966).

Zagros Mountain Belt


The extensive and beautifully exposed Zagros Mountain system provides an excellent, if
somewhat inaccessible, geologic laboratory for the study of structures caused by continental
collision. One interesting artifact that reveals the local rate of uplift is an abandoned irrigation
canal - built approximately 1700 years ago over the Shaur anticline in what was then
Mesopotamia (presently southwestern Iran). Water running through the canal has cut downward

nearly 4 m into the original rock floor, which has risen a total of about 20 m (Dennis 1972). This
defines a vertical arching rate of over 1 cm/yr, among the highest documented for large-scale
folding anywhere. This should be considered to reflect the extreme intensity of deformation
resulting from collisional tectonics.

Japan
Figure 1 is based on detailed geodetic survey information and reveals the eastern side of Japan
as undergoing significant, continual movement.

Figure 1

Inoue (1960) has determined that anticlines in the Uetsu Mountains have been rising at roughly 12 mm/yr for the past half-century (Vertical displacements in southern Japan Honshu and

Kyushu islands between 1900 and 1960, as determined from levelling data). As in Alaska,
tectonism is associated with subduction of the Pacific plate.

Large seismic events, such as the 1923 Kanto earthquake that led to the destruction of Tokyo
and Yokohama, are also periodic and create frequent topographic change in Japan. Leveling data
compared from before and after the Kanto quake showed that onshore displacements averaging
1-5 m occurred throughout the region. ( Figure 2 , Direction and magnitude of subaerial horizontal

displacements resulting from the Great Kwanto earthquake of September 1, 1923.

Figure 2

Data derived from geodetic triangulation measurements. The city of Tokyo was destroyed as a
result of fires caused by the quake.) The quake epicenter was located in Sagami Bay, and an
extensive sounding program, employing four ships full time, was organized and carried out by the
Japanese navy (Richter 1952). The findings of this meticulous work (more than 83,000 total
soundings) were nothing less than incredible.

The center of Sagami Bay is over 2000 m deep, and is partly the seaward extension of a faultbounded basin that continues onshore to the northwest as the Sakawa Valley. Vertical movement
along this fault system in the immediate vicinity of the shallow quake must have been truly
catastrophic. Soundings showed that a large area of the deeper bay subsided between 100 and
200 m; a maximum depth increase of 400 m was measured in one location. At the same time, in
the northern part of the bay, as much as 250 m of uplift was recorded.
Submarine slumping of sediment is presumed to have occurred. This, however, would only
account for decreases in the depth of the bay. One suggestion that plausibly accounts for the
great increases of depth is that resettling and compaction of water-saturated sediment may have
occurred as a result of the violent shaking. Certainly, this scale of movement, able to create
"instant topography" of dramatic relief, has never been documented on land. Some type of
catalyzed secondary effects were in all likelihood responsible for much of the depth change.

Stress & Strain


Basic Concepts
A rock is said to be in a state of stress when a force is applied to it. Earth stress is not simple, but
involves total force-per-unit-area for a particular point. What makes measurement complex is the
variety in the components of force (with their changing magnitudes and directions) that can occur
within a single volume of rock. Basically, however, we can think of stress as the threedimensional intensity of force acting at a specified point.
For any moment in its total history, a rock has structure. It is also true that for any such moment, it
is subject to stresses that tend to alter its properties. In reality, only in deep space would this
same rock be relatively free of stress. On earth, the forces that compose stress can be divided
into two main types: (1) body forces, which act at every point within the crust; and (2) surface
forces (also called applied forces), which act only at interfaces between objects and, therefore,
are defined only along surfaces.
We can better understand how these forces act by considering a single clastic grain buried within
the earth's crust. On this grain, two types of body forces act at all times: gravity and inertia. The
force due to gravity is called the lithostatic pressure, which results from the simple weight of
overburden transferred by grain-to-grain contact. If our particular grain becomes involved in
diastrophism, two types of surface forces will also exert their influence on it: pressure forces,
which act perpendicular to the surface of the clast, and viscous forces, which act parallel to these

same surfaces (Elliott 1976). (Note that viscous, as used here, does not imply anything about the
nature or behavior of the material composing the clastic particle.)
Figure 1 shows diagrammatically the breakdown of all these forces that compose the total stress
on this single grain.

Figure 1

Since gravity and inertia, by definition, act on every particle at every point in the crust, it is surface
forces that are primarily responsible for the creation of geologic structure and with which the
structural geologist is concerned.
A rock undergoes deformation when stress causes the displacement of particles within it. Such
stress can result from body or surface forces, or both. Deformation in nature is almost never
simple, since it results from a complex interaction between the chemical and physical properties

of a rock mass, its immediate environment, and the rate and intensity at which the displacing
forces are applied.
Geologists commonly make use of the concept of a stress field, which refers to the distribution of
stress acting within a defined body. Such a body can be a single folded layer or a sizeable portion
of a continent. A stress field is described as homogeneous if the stress at each point is equal.
The only situation that normally approximates this near the earth's surface is when stress is
almost totally due to lithostatic pressure. (We say "approximates," since there is nearly always
some component, however small, of lateral stress in the crust.)
The principal deformational effect of this body force is the compaction of sedimentary grains.
However, simple overburden can also create abnormally high pore fluid pressures that, in turn,
may lead to slumping and "growth" faulting.

The Three Principal Stresses


As a measure of force-per-area, stress is a vector quantity and thus may be expressed as the
sum of various components. ( Figure 1 , Resolution of unidirectional force (F) acting on a cube

face into the basic normal (N) and shear (S) components of stress.

Figure 1

Also shown is the idealized physical effect of each component.) If we trade our clastic grain for a
hypothetical cube of uniform composition and subject this cube to a progressive deformation, we
see that the total force (F) can be resolved into contributions that act perpendicular to the faces of
the cube, and along (parallel to) them. ( Figure 1 and Figure 2 , Resolution of normal and shear

stresses on a cube undergoing simple progressive deformation.

Figure 2

The direction and magnitude of force (F) remain constant. As deformation progresses, the relative
magnitude of the shear stress decreases, while that of the normal stress increases. )
On every face of our hypothetical cube, then, there will usually exist both normal stresses (the
pressure-force- per-unit area) and shear stresses (the viscous-force-per-unit area). Normal
stresses can be either compressional or tensional and tend to compact or separate particles
within a body. The effect of shear stresses, on the other hand, is to move particles past one
another ( Figure 1 ). During deformation, the relative magnitudes of these two stress types will
change, such that one type may decrease as the other increases ( Figures 2 ).
Materials science, through its detailed analysis of stress conditions, has derived an important
conclusion that has proven very useful in structural geology: for any point in a homogeneous
stress field, there exist three mutually orthogonal planes along which all shear stresses vanish
and thus only the components of normal stress exist. These three planes are known as principal

planes of stress, and the axes of their intersection are thus the principal axes of stress. These
axes, then, are used to describe what are referred to as the three principal stresses. This ideal
triaxial system makes everything simpler, since it allows us to speak in terms of only normal
stresses (see Figure 3 ,

Figure 3

Figure 4 and Figure 5 ), i.e., compression ("squeezing") or tension ("pulling-apart").

Figure 4

(Generalized orientation of the principal axes of stress for three basic geologic structures. In

Figure 3 and Figure 4, horizontal compression can be thought of as the actual applied force,
while Figure 5 shows the effect of horizontal extension.

Figure 5

)
In geology, principal stress is always spoken of in terms of compression, which is taken as
positive (for materials and engineering science, the opposite is true, i.e., tension is positive,
compression negative). The three principal axes, or stress directions, are correspondingly written
as 1 (maximum principal stress), 2 (intermediate principal stress), and 3 (least or minimum,
principal stress). For the purposes of structural geology, it is useful to understand four special
states of stress:

1. Uniaxial stress, where two principal stresses are zero and the other is nonzero.
2. Biaxial stress, where two principal stresses are nonzero and the other is zero.

3. Triaxial stress, where all three principal stresses are nonzero.


4. Pure shear stress, where 1 = 3 and is nonzero, while 2 is zero. This is a special case
of biaxial stress.
Within the earth's crust, the most common stress situation is triaxial, with 1 > 2> 3 >0.
In terms of actual physical effects, 1 and 3 can be thought of as representing the relative
compression and tension that act to deform a rock body by compacting particles into each other
and stretching them away from each other. We say "relative" because, very often, these stresses
are defined only in relation to each other. For example, a strong extensional event in the crust will
have a 1 associated with it that is secondary; i.e., the compression is passive, related mainly to
overburden pressure in this case.
The relationship between the three principal axes of stress can be pictured by considering the
triaxial examples shown in Figure 3 , Figure 4 and Figure 5 which place our hypothetical cube in
evolving fold and fault structures. Again, 1 represents the direction of maximum effective
compression; 3, extension. At this point, we should not be concerned with the specific angle that
the faults make with respect to these directions of stress. Note, however, that they, and 2, are
normal to the plane of the paper. In physical terms, 2 is parallel to the surface of the earth. It
basically represents the presence of essentially "infinite" neighboring material. Thus, with these
three axes, the conditions of stress at any point, or-as geologists often apply it-at a certain
location within a defined plate tectonic regime, can be described and related to actual geologic
structures.
Though there are no shear stresses acting on the surfaces of our cube, the fact that 1, 2 and 3
differ in magnitude and direction means that shear stress is generated within the cube. In fact, the
quantity 1- 3 (called the differential stress) is sometimes used as a general indicator of shearing
stress. If we were to measure the shearing stresses generated within our cube, we would find that
they reach a maximum along planes inclined at 45 to 1 and 3 that include 2. ( Figure 6 ,

Diagram showing planes of maximum shear stress in an idealized triaxial stress system that is
applied to a homogeneous block of material.

Figure 6

Note that these planes are inclined at 45 between any two principal stress axes.) This represents
the ideal case, and, as we will see later on, it helps explain why rocks do not fracture randomly
when stressed to the point of rupture.
Despite the extensive and detailed literature available on stress analysis in various rock types,
little is known about stress fields that govern deformation in progress. Though techniques exist to
determine in situ stresses, these only give us information about the final structure and the
stresses either stored within it or associated with its specific setting within a larger tectonic
environment. Such techniques do not directly tell us about the evolution of a particular structure.
Figure 7

Figure 7

and Figure 8 show two computer-generated interpretations of normal and shear stresses in a
folding layer that has already suffered approximately 630/c total shortening.

Figure 8

The short lines in Figure 7 are drawn perpendicular to 1 and therefore closely approximate 3.
We might just quickly note three important aspects to Figures 7 and 8: (1) how 3 changes
direction over different parts of the folding layer; (2) how this compares to the changing directions
of shearing stress shown in Figure 8 ; and (3) how the sense of these shearing stresses reverses
across the folding layer. All of these reveal significant elements in the folding process.

Types of Strain
Much of the terminology derived for understanding stress has also been applied to strain. Our
hypothetical cube is said to have suffered homogeneous strain (also called uniform strain) when
the strain is the same at all points within it. This means that originally straight lines remain straight
after deformation (part a of Figure 1 ).

Figure 1

Thus, for example, a cube becomes a rhombohedron, while a sphere inscribed within it becomes
an ellipsoid.
This should also help make clear the basic concept of inhomogeneous strain (part b of Figure 1 ),
by far the most common in naturally deformed rock. This type of strain involves some amount of
rotation in the position of particles, which means that originally straight lines become warped and
detailed analysis becomes impracticable. It is, therefore, almost always useful to find some way in
which natural deformation can be approximated as homogeneous. The most common approach
is to consider geologic structures as the summation of many localized homogeneous strain fields.
This method has proved especially helpful in the explanation of secondary rock fabrics, such as
mineral alignment and fracturing.
Such fabrics often provide invaluable clues to both small- and large-scale structural patterns.
Natural fracturing, of course, is of particular importance to petroleum geology, and knowledge of
the stress-strain relationship associated with it can be very useful. As we will see, fracture
patterns very often have a direct causal relation to major structures, such as faults and folds. The
summation-of-local-strains method, therefore, will usually reveal this and permit the geologist to
predict patterns in adjacent, undrilled areas. Several examples of this are given later on, when we
look at fracturing in detail.
There are a number of basic ways in which deformation by homogeneous strain at constant
volume occurs. Those that involve simple flattening and stretching are shown in Figure 2 .

Figure 2

(Several basic types of homogeneous strain imposed on a cube of ideally uniform composition. In

each case, the inscribed circle and ellipse represent cross sections through the strain ellipsoid
before and after deformation. The types of strain shown are: a. uniform extension; b. uniform
flattening; and c. plane strain. Note that no strain occurs in the intermediate direction. )
To understand how geologists treat natural deformation, however, it is also necessary that we
look at the two basic types of shear strain. ( Figure 3 , Hypothetical cross section and diagram to

illustrate domains of pure and simple shear in a series of folds that show progressively greater
total strain.

Figure 3

The shape of the strain ellipse can be the same for either type of shear and cannot be used to
derive detailed strain history.) Both types help us explain a great many large- and small-scale
features seen in rocks.
Pure shear is a form of strain in which no rotation of the strain axes takes place. It is often
referred to as an irrotational deformation." It results from uniform extension in one direction and
contraction perpendicular to it ( Figure 4 , Particle paths in simple and pure shear. Note rotation

involved in pure shear). Strain that approximates pure shear is seen in many folds.

Figure 4

In simple shear, all particles within a body are displaced in one direction. This is our cube pressed
into a rhombohedron again; this time, however, we need to take note of the rotation in the strain
ellipsoid. Simple shear can be visualized by imagining the result of placing our cube (with its
inscribed sphere) between the two surfaces of an active fault. The shearing motion created by
these two surfaces stretches and flattens the sphere into a strain ellipsoid whose long axis is
progressively rotated until it is nearly parallel to the fault plane itself. Displacement within such a
body takes place by slippage along closely spaced planes ( Figure 4 ).
In actual materials, this can be accomplished in a number of ways, for example, by slippage
between grains or crystals, or by actual flow at elevated temperatures and pressures. As we shall
see, this style of deformation has widespread application to geologic structure.

Strain Ellipsoid

Let us refer again to the sphere inscribed within our cube that undergoes homogeneous strain.
Giving this sphere axes 1, 2 and 3, we find that these remain orthogonal through the
homogeneous deformation and can be used to describe the long, medium, and short dimensions
of the resulting ellipsoid. ( Figure 1 , Several basic types of homogeneous strain imposed on a

cube of ideally uniform composition.

Figure 1

In each case, the inscribed circle and ellipse represent cross sections through the strain ellipsoid
before and after deformation. The types of strain shown are: (a) uniform extension; (b) uniform
flattening; and (c) plane strain. Note that no strain occurs in the intermediate direction.)
Together, these are called the principal strain axes of the strain ellipsoid. Just as stress can be
said to exist at every point within a body, so is there a corresponding strain ellipsoid for these
points, once deformation has taken place. Thus, comparison between the "before" and "after"

shape and axes of the sphere inscribed within our cube provides us with a measure of the
amount and type of strain.
It is standard practice for geologists to derive the principal axes of stress by superimposing them
on the strain ellipsoid. ( Figure 2 , Note that this superposition assumes homogeneous strain, i.e.,

maximum shortening occurs in the direction of maximum principal stress, and maximum
extension in the direction of minimum principal stress ).

Figure 2

This is an optimistic simplification (as we have seen, the principal axes of stress and strain
coincide only under conditions of homogeneity), but can be very useful. It has, for example,
offered considerable insight into basic mechanisms and patterns of deformation-particularly

faulting and fracturing-on many scales. This will become increasingly clear in the following two
sections on folding and faulting.
Because of their frequent use of structural cross sections, geologists have also found it
advantageous to make use of a strain ellipse-essentially a cross section through the strain
ellipsoid along the 1 3 plane (i.e., the one that involves 1 and 3). The justification for this is,
again, dependent on the assumption of homogeneous strain. Because of the regional nature of
most tectonism, and the layered nature of lithologic sequences, many geologic examples of strain
can be considered to approximate plane strain (see part c of Figure 1 ). In this type of
deformation, the intermediate axis remains the diameter of the "original" sphere (the 2 axis-parallel to 2-- in part c of Figure 1 ), while shortening and stretching occur along the other two
axes.
Thus, two dimensions are sufficient to describe the strain at a particular point. If we are ready to
accept the assumption of homogeneous strain, the strain ellipse becomes one of our principal
indicators for the summation- of-local-strains method. ( Figure 3 Hypothetical cross section and

diagram to illustrate domains of pure and simple shear in a series of folds that show progressively
greater total strain.

Figure 3

The shape of the strain ellipse can be the same for either type of shear and cannot be used to
derive detailed strain history.)
As shown by Figure 4 , (Example of how the strain ellipse here constructed from deformed

oolites can be used as a descriptive guide to deformation intensity and orientation.

Figure 4

Shown is a cross section through the south Mountain fold of western Maryland, U.S ) the strain
ellipse can be an important guide to the general degree and style of deformation.
Some natural materials, such as ooids, spherulites, pebbles, certain fossils, and reduction spots
in shales, can be used as qualitative ellipses or, in some cases, ellipsoids. However, because
volume changes frequently occur during deformation (especially in carbonates) any quantitative
determinations of strain based on such materials must be used with caution.
In principle, any object whose initial shape is known can act as a strain indicator. Such an
indicator can be important to the subsurface explorationist, since it may be the only direct
evidence available for how much strain has affected the fabric-and thus porosity and
permeability-of a lithologic section. In most cases, the degree of tectonic influence on grain
texture is fairly apparent from petrographic study. Strain indicators are primarily useful where this
may not be clear and where special circumstances warrant mathematical determinations of strain.
Specific techniques for measuring finite strain from oolites and spherulites are given by Ramsay
(1967) and Ramsay and Huber (1985).
Nearly all deformation in nature is in-homogeneous. Not only do originally planar surfaces
become complexly curved, but volume changes that involve both loss and addition of material
frequently take place. Because of their pronounced heterogeneity in composition, thickness, and
thus strength, rocks do not behave passively during deformation, but adjust in complex ways.
Some units become strain-hardened and are able to withstand and transfer greater and greater

amounts of stress as deformation progresses; other lithologies, in contrast, are fated to absorb
stress by flowage, recrystallization, and the development of secondary fabrics such as cleavage.
Again, despite the dominance of inhomogeneity in nature, both local and regional deformational
history can be reconstructed by assuming near-homogeneous strain domains. On a large scale,
this often establishes the regional nature of stress and strain. Geologists often estimate a regional
strain ellipse based on the orientation of major structural trends. This is called the mean strain
ellipse and is often useful in explaining such trends in terms of plate interactions.

Rock Strength
A substantial literature exists with regard to experimental "rock crushing" in the laboratory, the
intent of which is to simulate and analyze the process and effects of deformation. Studies are
usually performed on cylindrical, corelike samples, which are subjected to compressive or tensile
stresses in a chamber whose pressure and temperature are regulated. From these studies,
scientists have determined that deformation generally progresses through three main stages
(Nadai 1950). These are defined by the behavior of the deformed material, and are most clearly
and simply shown by the use of stress-strain diagrams ( Figure 1 ).

Figure 1

In order, the three stages are as follows:

ElasticDuring this initial stage, the stress-strain relationship is linear. If stress is


removed, the body reverts to its original dimensions. No deformation (permanent strain)
results; strain is said to be completely reversible ( Figure 1 , segment A).

Plastic As stress continues to increase, it will eventually reach some limit beyond
which the body suffers permanent strain. This is termed the elastic limit (or yield stress),
and beyond it material is said to behave plastically; any increase in stress brings a
corresponding increase in deformation ( Figure 1 , segment B).

Rupture With continued increase of stress, the body will eventually fracture (rupture) (
Figure 1 , point C).

These three stages through which deformation normally progresses are idealized as the behavior
of three hypothetical "bodies" that are subjected to stress, as shown in Figure 2 (Stress-strain

relationships for several ideal materials: (a) Hookean (elastic) body; (b) St.

Figure 2

Venant (plastic) body; (c) Newtonian (viscous) body ). Each style of behavior is denoted by the
name of a well-known mathematician. Each should be apparent from the stress-strain graphs
shown.
A Hookean body knows only elastic strain before rupture, and is approximated by a simple elastic
spring attached to a fixed body. A St. Venant body, in contrast, shows elastic strain up to a yield
stress and then deforms indefinitely by shear strain at that same stress. This type of behavior is
approximated by a weight that is pulled across some surface by an attached spring: the spring
stretches elastically up to the point where friction on the table top is overcome and the weight
begins to slide. The last example, called a Newtonian body (or, sometimes, Newtonian liquid),
has no shear strength at all and therefore exhibits no elastic strain. It deforms by what is called

"viscous strain." In part c of Figure 2 , this type of body is represented as a porous piston pulled
through a fluid-filled cylinder. A Newtonian body, then, will deform indefinitely in response to any
shear stress, with the total strain being directly proportional to the amount of elapsed time.
These three types of ideal bodies, then, help describe the components in progressive deformation
when it is caused by increasing amounts of stress. We might compare them to the graphs shown
in Figure 3 , which describe two generalized stress-strain relationships for known materials.

Figure 3

The principal difference between ideality ( Figure 2 ) and reality ( Figure 3 ) lies in the behavior
known as strain hardening.
Such hardening is often the result of complex readjustment-even recrystallization-of the material
suffering strain. We can envision it on one level as being due to the compaction and realignment

of particles during progressive deformation. In sandstones, for example, stress concentrations


along grain boundaries result in extensive grain fracturing and dissolution. This causes the
progressive filling of pore space with grain fragments and recrystallized quartz. All other
conditions (e.g., temperature, pressure) being equal, this will increase rock strength; that is,
greater and greater amounts of stress become necessary to impose the same increment of
deformation ( Figure 3 , upper curve). We expect certain clastic lithologies to respond in this
manner.
On the other hand, strain hardening can occur up to some ultimate strength, after which the
stress necessary to cause a given strain decreases continually ( Figure 3 , lower curve).
Lithologies such as salt and gypsum-anhydrite may behave this way under certain conditions.

Brittleness and Ductility


On the basis of the differences between elastic and plastic behavior, we are able to characterize
the general response of materials as either brittle or ductile. Brittle material ruptures before any
significant plastic deformation occurs. Such behavior in rocks is marked by the development of
breakage discontinuities along the planes that represent maximum shear strain. These are not
necessarily faults or fractures visible to the unaided eye, but may take place between individual
grains or within crystal lattices. In contrast, material is described as ductile if it is able to undergo
a large amount of plastic deformation before failing.
In every situation, whether a rock responds in a brittle or ductile manner depends on several
major controls: composition, effective confining pressure, temperature, and strain rate. It is, in
fact, relatively meaningless to speak of the true "strength" of a rock without reference to these
parameters. Near the surface, at low temperatures and pressures, rock will tend to act in a more
brittle manner. With growing overburden (which increases both pressure and temperature),
ductility generally increases. At the same time, however, the element of time is crucial; if small
amounts of stress are applied over sufficiently long periods of time, almost any rock will deform
plastically (only those lithologies with very low resistance to shear, such as basalt, may not). But
given normal rates of tectonic deformation ( Figure 1 and Figure 2

Figure 2

, Charts comparing the rates of common sedimentary and tectonic processes in various parts of

the world), there is some transitional depth range over which the response of a particular rock
type will grade from dominantly brittle to ductile.

Figure 1

Figure 3 is a schematic representation of the spectrum from brittle to ductile behavior in


limestone, as determined by laboratory tests.

Figure 3

Such testing applies a uniaxial stress (either compression or tension) to a cylindrical sample that
is sealed within a chamber whose temperature and confining pressure can be regulated.
Limestone and marble have been favored as samples, since these lithologies are more isotropic
than clastic rock types.
It should not be assumed that brittleness guarantees faulting, or that ductility inevitably leads to
folding. Most lithologic sequences are highly anisotropic, and, as a result, respond to tectonic
stress in a manner that combines both styles of behavior. For example, the very shallow Shaur
anticline amply demonstrates relatively brittle bending, which involves rigid body rotation and
translation at shallow levels in response to high strain rates. At the same time, the extensive
fracture system that characterizes this and many of the Zagros folds-and, in fact, is responsible
for the very high rates of production (up to 80,000 bbl per day per well) in the Iranian Asmari oil
fields-indicates simultaneous brittle rupture.

Figure 4 ,

Figure 4

Figure 5 , and Figure 6 indicate how laboratory analysis has shown deformation to vary as a
function of pressure,

Figure 5

temperature, and strain rate.

Figure 6

The other factor that determines a rock's response to stress is, in one sense, the most obvious-its composition. Figure 7 is a diagram derived by Handin et al.

Figure 7

(1963) from extensive lab experiments on natural rock. Shown are measured ductilities for four
common sedimentary lithologies in a water-saturated condition.
One interesting point made by this diagram is that, with only one kilometer of burial, marked
differences in ductility already exist between limestone and other lithologies. This is partly the
result of pore pressure effects that strongly reduce ultimate strength. It is also, however, directly
related to the mineral structure of calcite, which allows for significant intracrystal line gliding, even
at moderate pressures and temperatures.
The same trend of ductility increase is true for sandstones, and is often aided by pore pressure
effects. We might also note that dolomite-while slightly more ductile than quartzite at depths
below two kilometers-shows the tightest range of permanent strain before rupture. Due to its
mineral structure, dolomite does not deform readily by intracrystalline gliding.

What is normally referred to as "rock strength," then, is not a fixed property, but a relative
response, determined by a specific set of immediate environmental conditions. Since both
engineers and geologists make use of the terms compressive strength, tensile strength, and
shear strength, three major points should be made in relation to these. First, a material that
behaves in a brittle manner can fail only under tension or shear. Second, almost all materials are
weaker under tension than compression ( Table 1 , below). And third, it is very often the shearing
strength of materials that determines when and how they will fail.

Average crushing strength Tensile strength Shearing strength


Sandstone

740

10-30

100-200

Limestone

960

30-60

150-250

Quartzite

2020

30-90

100-300

Granite

1480

30-50

150-300

Basalt

2500

-------

50-150

Table 1 . Measured strength of various rock types at standard temperature and pressure. Note

the relative similarity in tensile and shear strength among materials with widely differing crushing
strength. Note also the data for basalt. (Courtesy AGI)

The Mechanical Properties Log as an Estimate of Formation Strength


The mechanical strength of a particular formation can be estimated from a mechanical properties
log, such as that in Figure 1 .

Figure 1

It can be very important to determine the strength in reservoir sandstones in order to prevent
sand production problems. Both experimental and field test data indicate that a good correlation
exists between rock strength, rock density, and certain dynamic elastic constants derived from
compressional and shear velocities as recorded on sonic logs. Intrinsic strength is dependent on
several major factors, including grain texture, cement, and the effective stress, which is defined
as the overburden pressure minus the pore pressure.
In the example shown, G represents the shear modules, defined as the applied stress divided by
the resulting shear strain; Cb is the bulk compressibility, defined as dilatation divided by the
hydrostatic pressure; andz is the total intergranular space, derived from the interpretation of
sonic logs. As explained by Tixier, Loveless, and Anderson (1973), the quantity G/C b is the most
sensitive indicator of formation strength. Note the corresponding changes in all log curves from
the weak sand to the relatively tight lime unit below.

The Concept of Rock Competence

In addition to the concepts of brittleness and ductility, structural geologists frequently describe the
relative strength of rock material in terms of competence.
This term is used to indicate contrasting behavior, or "ductility contrast," within a layered
sequence of different rock units. Competent units are those that deform in a more brittle manner
relative to incompetent units, which show more ductile behavior. A simple example of ductility
contrast would be an alternation of thick sandstone and shale layers. During compression, folding
of the competent sand units would control the deformation in the weaker incompetent shales,
which tend to flow and accommodate the evolving fold geometry.
"Competent" and "incompetent," therefore, are strictly qualitative and, more importantly, relative
terms. Looking at Figure 1 we can see that while limestone and dolomite share similar
competencies in relation to quartzite at shallow levels, they differ dramatically at depth.

Figure 1

Below 4 km, limestone becomes highly incompetent relative to both quartzite and dolomite.

Conjugate Fracturing and Shear Couples


During experimental rupture tests (part (a) of Figure 1 ), rock often fails in the form of conjugate
fractures.

Figure 1

(Part (a): Shear fractures typical of failure in dry limestone and marble: 1Marble, brittle failure at

25 C, 35 bars, and 1% strain; IIMarble, "transitional" failure at 25 C, 280 bars, 20% strain; III
Solenhofen limestone, near-ductile failure at 25 C, 1000 bars, 11.2% strain; IVSolenhofen
limestone, ductile behavior followed by rupture at 150 C, 6500 bars, 9.1% strain. Part (b):Diagram
illustrating angles of rupture in relation to maximum and minimum axes of stress. ).
These represent full development of the discontinuities depicted in Figure 2 .

Figure 2

(Schematic diagram showing the spectrum from brittle fracture to ductile flow during typical

compression and extension experiments on rock material. Typical strains before rupture as given,
as are stress strain curves for uniaxial compression and extension. Ruled portions on the latter
indicate the variation common for each case.).
The angle fractures make with respect to 1

and 2 is almost never 45 (see part b of Figure 1

and Table 1, below). Rock is generally a heterogeneous material and a great variety of internal
physical properties determine the actual orientation of fractures. Generally, however, fractures
occur at low angles (roughly 15-30 degrees) toand complementary high angles (60-75
degrees) to 3. Part a of Figure 1 portrays several examples of faulting created during testing.
Table 1, below, gives the predicted angles of fracturing for common sedimentary and igneous
lithologies. Comparing these data with those presented in Table 2, below, it would seem that a
general correspondence exists between fracture orientation and shearing strength. This, of
course, is not surprising.

Movement along conjugate fractures takes place by shearing, and the opposing motion across
them is said to represent a shear couple. It is important to remember the basic orientation of 1,
2 and 3 in relation to shearing, since shear fracture is one of the most common types of strain
observed in rocks. The effects of shearing are well documented on all scales, from displacement
within calcite crystal lattices to the motion between lithospheric plates. Figure 3 and Figure 4 (San

Andreas fault system,

Figure 3

California) offer two examples of regions where the resolution of principal stress directions is fairly
straightforward.

Figure 4

Geologists have found it useful to delimit shear couples in a wide variety of tectonic
environments. In particularly complex regions, recognition of direction and attitude of the principal
stress axes has proved its worth as both an explanatory and predictive tool.
Table 1. Predicted angles of rupture in relation to principal axes of compression and tension.

Mean shear strength values are from Table 2. (Courtesy AGI)

Rock Type

Shale (calcareous)

13 77

Siltstone

20 70

Sandstone

21 69

Limestone

Main shear strength kg/cm2

150
200

fine-grained

16 74

oolitic

23 67

Granite

17 73

275

Basalt

21 69

100

Table 2. Measured strength of various rock types at standard temperature and pressure. Note the

relative similarity in tensile and shear strength among materials with widely differing crushing
strength. Note also the data for basalt. (Courtesy AGI)

Average crushing

Tensile

Shearing

strength kg/cm2

strength

strength

Sandstone

740

10-30

100-200

Limestone

960

30-60

150-250

Quartzite

2020

30-90

100-300

Granite

1480

30-50

150-300

Basalt

2500

-------

50-150

Figure 1

In Figure 1 (Generalized stress-strain curves for geologic materials), what portion of the graph
indicates strain hardening?

3-Folds
General Introduction
Certainly, folds are visually intriguing geologic structures. Geologists have long been drawn to the
drama and spectacle of folded rocks. There is, as Hans Cloos (1939) observed, a special
satisfaction and awe to be found in the evidence of earth stress having deformed originally planar
and rigid rock into coherent arches many hundreds of feet high ( Figure 1 , Folded strata).

Figure 1

Even for the most seasoned structural geologist, the curving lines of folded strata never lose their
particular aesthetic appeal. From a descriptive-analytic point of view, this degree of attention is
justified by a pronounced variation in specific fold style, geometry, and generating mechanism, by
attendant suites of complex but informative minor structures, and by the overall importance that
fold structures have for petroleum exploration.
The specific attraction for geologists in folding lies in its direct implications concerning the origin,
rate, evolution, and diversity of crustal movement. Several lines of evidence are considered in
unraveling tectonic mechanisms:
the geometry of folds.
the experimental simulation of folding.
the specific relationship of folding to faulting.

the relationship of fold belts to other tectonic provinces (analysis of structural style).
the small-scale structures associated with folding.
To some degree, modern petroleum exploration was born with the anticlinal theory. I.C. White's
formulation of this hypothesis in 1885 represents the first conceptual approach to systematic and
successful prediction of oil and gas occurrence. The elegance of this theory lies in its simplicity
and in its enduring practical application, which, with the aid of modification and expansion,
continues today. The anticlinal theory, however, is not the only direct benefit that exploration and
production geologists have derived from a knowledge of folding. A few others include:
the prediction of fracture trends.
the prediction of fracture distribution and intensity.
the explanation and prediction of tectonically induced variations in reservoir thickness,
porosity, pressure, and general fluid behavior.
the prediction of oilfield shape and extent.
the delineation of constraints on contour and isopach work.
the prediction of deeper accumulations beneath known fields.
the understanding and prediction of significant faulting and its influence.
the understanding and prediction of the general orientation and style of neighboring
structures, and, therefore, the location of potential traps in other parts of a specific region.

Description of Folds
A fold is produced when initially planar layers become bent, i.e., nonplanar. To the naked eye,
folds in rock appear to be continuous; layering is preserved, not truncated. For many years,
geologists took the coherent curvilinear geometry of folds in sedimentary strata to mean that
strain had accumulated gradually and continuously. Within the past seventy-five years, however,
it has become evident that a great deal of fold deformation is accomplished by discontinuous,
incremental gliding along bedding planes, within layers, and between and inside of individual
grains and crystals. Thus, the strain involved in folding imposes change in both the geometry and
internal physical properties of the layers, which normally remain relatively coherent.

Folding is seldom the result of a single, discrete deformational episode.


Diastrophism nearly always occurs in what can be described as pulses: during the evolution of an
orogen, stages of intense tectonic activity wax, wane, and frequently overlap at irregular or semiregular periods that can affect different parts of a specific region at different times. The reasons
for this have not yet become clear, but they are assumed to be more closely tied to the nature of
stress generation within the earth's crust, rather than to the mechanics of strain response.
The most basic types of folds are the anticline, in which right-side-up bedding dips away from the
fold crest, and the syncline, in which bedding dips inward. Regional arching with associated
folding produces an anticlinorium. Similarly, large-scale downwarping of a fold belt creates a
synclinorium. ( Figure 1 , Example of anticlinorium and synclinorium from the Rhineland region of

Germany.

Figure 1

Note the large scale of structures. Dashed lines indicate cleavage. ) Generally, the orientation of
the anticlinorium or synclinorium roughly parallels that of the folds within it.
The terms anticline and syncline can only be used once the younging direction is known. In
strongly folded regions, beds may be completely overturned. Thus, during the early stages of
analysis, before stratigraphic relationships have been established, it is common practice among

structural geologists to make use of the terms anti form and synform, neither of which implies the
direction of younging.
To describe the great variety in fold morphology, a correspondingly diverse nomenclature exists.
This discussion uses only more fundamental and commonly employed terms. The reader who is
interested in more detailed study is referred to the extensive treatments given by Whitten (1966),
Ramsay (1967), and Davis (1984).

Basic Fold Geometry and Orientation


Figure 1 displays the principal descriptive features of a simple fold.

Figure 1

Several of these represent imaginary lines and surfaces that help us analyze a fold as a
geometric phenomenon. Except for those that require the third dimension for their depiction,

these features are most often shown on cross sections drawn perpendicular to the long direction
of the fold. This type of section is called a fold profile. In Figure 1 , the definition of each term
should be fairly obvious from the geometry shown.
The hinge of a fold is its point of maximum curvature at one particular location (cross section)
along its length. The fold axis, or hinge line, contains all hinge points in a folded layer. The axial
surface, then, connects all these axes within a single fold composed of multiple layers. In some
cases, the axial surface is planar and can be called the axial plane. The plunge of a fold is the
angle between its axis and a horizontal plane. ( Figure 2 , Illustration of fold plunge and its

change along strike. Such culminations and depressions characterize folds of many scales in
mountain belts.)

Figure 2

For the purposes of practical simplification, geologists often treat folds as if they were cylindrical,
i.e., generated by a line moved parallel to itself in space ( Figure 3 , Perfect cylindrical fold

system), or conical, i.e., generated by the same line with one end fixed.

Figure 3

Noncylindrical and nonconical folds are common in nature, but can often be considered the
summation of cylindrical or conical parts ( Figure 4 , Complex folding in the Vanige region of

southwestern France).

Figure 4

As a result, for extrapolation to greater or shallower depths, geologists use the interlimb angle to
project fold geometry ( Figure 5 , Classification of folds on basis of interlimb angle. Lower figure

indicates how this angle is determined for curvilinear folds).

Figure 5

In petroleum geology, folds are most often drawn in vertical cross section (which is different from
a fold profile if the structure is plunging) and are shown as cylindrical, even to depth. This must be
recognized as an approximation, especially in areas where thick shale or evaporite intervals exist.
In profile, folds are commonly described as symmetrical or asymmetrical on the basis of whether
their axial surface is a plane of symmetry or not. More specifically, as shown in Figure 6 , folds
can be classified as upright, overturned, or recumbent on the basis of the inclination in their axial
planes.

Figure 6

Large-scale recumbent fold structures, generally accompanied by thrust faulting, are often called
nappes (German "decke"; English "sheet" or "cover"), a term originated by Alpine geologists to
describe large, thin, flat-lying structures that have been transported like rumpled sheets for
considerable distances ( Figure 6 and Figure 7 ).

Figure 7

This style of deformation in unmetamorphosed rocks is generally unusual except in the European
Alps.
Geometrically, one of the simplest forms of folding is the monocline.
( Figure 8 , Simple block diagram of monocline.

Figure 8

Note how upper layers are passively draped over faulted basement.) This is frequently a steplike
drape in sedimentary cover over faulting in crystalline basement rocks. It basically represents,
therefore, a form of passive bending, in contrast to the active buckling that produces most
anticlines and synclines in orogenic zones.

Fold Closure
The principal importance of folds in petroleum geology is very often related to the concept of
structural closure. As shown in Figure 9 , (Definition of structural closure.

Figure 9

Part a shows that it is the crest of the fold that determines closure (not the hinge); b illustrates
how the geometry of closure varies with that of the fold.) closure is the pocket formed between
the crest of the fold and the lowest closed contour. It is measured in terms of the vertical height of
this pocket and is an overall indicator of a fold's total capacity to hold oil and gas, given good
reservoir properties. Structural closure should be contrasted with structural relief, which is the
relative height to which the fold rises above the regional slope.

Basic Fold Classification Parallel and Similar Geometry


In addition to describing folds according to their geometry and orientation, geologists have found
it useful to classify them on the basis of morphological criteria that reflect mechanisms of
deformation. This classification scheme examines folds in profile. Its principal attempt is to
distinguish structures that have formed by flexure and layer-parallel shearing from those that

have resulted from flow. Figure 10 shows two idealized profiles of these types.

Figure 10

Parallel (also called "concentric") folds are so named because the upper and lower surfaces of
individual layers remain parallel during folding. Note in the figure that this does not mean all
layers necessarily remain parallel to each other. It does mean, however, that they maintain a
constant thickness throughout the fold (part a of Figure 10 ). This makes for crowding of hinge
areas, which either causes the fold to die out rapidly with depth or to compensate the increasing
lack of space by faulting ( Figure 11 , Hypothetical cross section to illustrate how parallel folding

creates overcrowding and faulting in hinge areas.

Figure 11

Note also the extensional fracturing along the crest of the fold) and basal detachment (also called
"decollement") ( Figure 12 , More detailed prediction model based on observed structures in

productive areas of the Rocky Mountain region of Colorado and Wyoming).

Figure 12

In Figure 10 , this crowding is shown as an increased crumpling in the thinner layers. This is one
form of what is termed disharmonic folding, so called because of the difference in wavelength.
Another way to view this hinge crowding is to note that with depth in a parallel fold, anticlines
decrease in size, while synclines increase. Generally speaking, parallel folding is more common
when deformation is of moderate or low intensity and involves thick, competent units.
In similar folding, by contrast, substantial flowage of material is assumed to take place. This can
occur in two basic ways: (1) by material transfer out of the fold limbs and into hinge areas, and (2)
by movement along planes of cleavage that develop parallel to the axial plane of a fold.
Geometrically, the shape and size of a similar fold is retained at depth (part b of Figure 10 )
without faulting. The flowage required for similar folding is most common in incompetent

lithologies (e.g., shales and evaporites) and at elevated temperatures and pressures. Slippage
along axial plane cleavage, as we will see, often leads to similar folds in shales.
Though both parallel and similar end-member types are seen in the field (similar folds being
particularly characteristic of metamorphic rocks), folds in sedimentary rocks more often show a
compromise between them. For example, folds that have concentric limbs, often show thickening
at their hinges. Particularly in stratigraphic sequences with high ductility contrast for example,
competent sandstone or carbonate units separated by thick shale or salt intervals-substantial flow
from sheared limbs into hinge areas should be expected.
In such cases, competent layers show more or less parallel folding. This fact is often important to
keep in mind when structures are projected to depth in cross sections. Even a moderate degree
of similar folding will prove depth predictions based on parallel geometry to be in substantial error.
In general, folds formed under low confining pressures-such as most of those that trap
hydrocarbons-are mainly the result of rigid body rotation and layer-parallel shearing. Such a
mechanism produces parallel folds in competent layers, but may be accompanied by flow in
incompetent lithologies. Figure 13 and Figure 14 give two examples of such folds.

Figure 13

(Figure 13 is a cross section through Spring Creek field in northwest Wyoming (western margin of

the Big Horn basin), showing many of the structural features predicted in Figure 12.

Figure 14

Note how closure changes with depth due to faulting. Figure 14 is a classic section through the
Jura Mountains of eastern France showing sharp, boxlike folds developed above a shallow
decollement plane. Note how these folds show a combination of parallel and similar geometry, as
well as pronounced flowage of Triassic gypsiferous marls into hinge areas. )

Fold Mechanisms and the Distribution of Strain


The distinction between parallel and similar folding is very useful as a general guide to fold
geometry in many petroleum provinces with relatively simple structural traps. However, in more
complex regions, where rocks have developed secondary tectonic features (e.g., cleavage or
fractures) that can influence entrapment, it is sometimes necessary to determine mechanisms of
folding in order to predict the potential existence and geometry of traps. Fracturing is often

genetically related to folding; therefore an understanding of mechanisms will to some degree also
explain the existence, orientation, and trends of fracturing.
Geologists working mostly in regions of sedimentary rock emphasize the bending or buckling of
layers, while those more familiar with metamorphic terrains have stressed the ability of rocks to
flow. The great diversity of fold structures in the earth has resulted in the development of
competing terms and concepts. Here, we attempt to simplify these and make them easier to
understand by looking first at mechanisms that describe how individual layers become folded,
and then at mechanisms that relate to the folding of multiple layers.
To explain what is seen in nature, we can identify three main processes that seem to explain the
folds we might observe in single layers of any lithology. The principal importance of these
processes for our understanding of folding is that they describe what happens within a particular
layer.
1. Pure bending (part b of Figure 1 ) occurs by the simultaneous stretching of material
along the outer arc of the folded layer and the compression of material along the inner
arc.

Figure 1

Between these zones of relative tension and compression is a "neutral surface" where no
strain occurs. The compensation of stretching by shortening means that the resulting fold
will be parallel in its geometry.
2. Flexural shear (part c of Figure 1 ) deforms a layer by internal shear along planes that
are parallel to the folding layer. As a whole, the layer can be thought of as undergoing a
shear couple, with the relative sense shown in part a of Figure 1 .
Note that with this mechanism, no deformation (such as stretching) occurs along the plane of the
fold layer. Slip takes place along discrete surfaces within the layer.
A special type of flexural shear folding is termed flexural flow. This is said to occur when
deformation takes place on the scale of individual grains, and there are no discrete surfaces of
displacement. A certain amount of flow into hinge areas often occurs during this type of folding.
This is generally small and depends on the lithology involved.

Flexural shear produces folds with mostly parallel geometry.


3. Shear folding (part d of Figure 1 ) occurs when shearing takes place within a layer
along closely spaced planes oriented obliquely to the layer itself. These planes are often
roughly perpendicular to bedding in hinge areas and oblique to it along fold limbs. This
mechanism involves simple shear and produces similar folds. It is much more common in
ductile lithologies and in medium- or high-grade metamorphic rocks. In thick shale
intervals, shear folding is apparently a late-stage mechanism that begins after the
development of penetrative axial plane cleavage during flexural shear.
Each of these mechanisms produces plane strain in single layers. Each, admittedly, represents
an ideal condition and is only occasionally seen in nature in its "pure" state. This is because the
process of folding is strongly determined by the character of the entire stratigraphic sequence
involved. Through the study of laboratory models, geologists have confirmed their intuition that
folding is often the result of progressive mechanical adjustment both within and between
individual layers.
In many cases, it appears that single layers can shorten (through compaction, among other
processes) by as much as 20% before folding actually begins. This means that a certain amount
of strain hardening may occur very early. At the same time, most stratigraphic sequences are
characterized by high ductility contrast, with sandstone and carbonate units frequently being the
more competent, and shale or salt the less competent, intervals. Generally, it is the former that
fold first and control the subsequent deformation to a large extent. Response of less competent
layers, therefore, is relatively passive.
With regard to the folding of multilayered sequences, it is important to be familiar with three basic
mechanisms. These are the following:
1. Flexural slip refers to the shear-or slip-that occurs between layers as a fold develops.
As implied by part a of Figure 2 , this slip compensates the bending and rotation of
comparatively rigid layers that do not undergo large amounts of internal plastic
deformation.

Figure 2

It therefore occurs principally while a fold is still open. Thinner, more ductile layers
between competent layers are subjected to relatively intense and penetrative shearing.
Notice that this style of folding requires that tectonic compression occur parallel or
subparallel to bedding. Looking at the mean strain ellipses in part a of Figure 2 , we might
note how similar are their orientations to ellipses in the limbs of a fold developed by
flexural shear (part c of Figure 1 ). Flexural slip folds are the most common folds in
nonmetamorphosed sedimentary rocks.
2. Shear folding is a topic we discussed with respect to the internal deformation of single
layers. In multilayered sequences, shear planes (such as cleavage) can cut across
competent layers and thus allow for a certain (usually minor) amount of slip (see part b of
Figure 2 ). Such folding, obviously, would succeed earlier deformation that created the
shear planes. In general, shear folding becomes significant at higher grades of
metamorphism.

3. Kink folding (part c of Figure 2 ), also called chevron folding, is caused by a


combination of flexural slip between layers and sharp, localized yielding (with some
ductile flow) in hinge areas. Kinking is most often seen on a mesoscopic (hand-sample)
scale in the field and commonly occurs as a conjugate set of single kink bands. (The term
"chevron fold" is most often reserved for larger structures, with amplitudes measurable in
meters rather than centimeters.) The details of kink band formation are somehow linked
to the anisotropy present in thin-bedded sequences of alternating competent and
incompetent layers. As a general rule, kinking is of relatively minor importance to
petroleum exploration, though a few important exceptions exist, such as the Ventura
anticline in southern California, with its reserves of nearly a billion barrels.
One final component of the folding process should be mentioned. In addition to the various
mechanisms we have discussed, many folds also undergo a component of flattening. This
additional shortening, shown in Figure 3 , takes place approximately normal to 1 with
simultaneous extension in the 3 plane.

Figure 3

As the figure shows, sufficient flattening may "tighten" a fold by pure shear strain bringing all
strain ellipses into relative parallelism and originally flexural slip folds into more or less similar fold
geometry.
Some geologists believe that almost all folding involves this type of additional strain (see Ramsay
1967). Relevant experimental and theoretical work has, thus far, been limited, but has apparently
confirmed the assumption that flattening is more important in sequences of low ductility contrast.
Given the an isotropy and consequent ductility variation in most lithologic sections, the folding
process takes place by sometimes complex combinations of these end-member mechanisms.
This we should expect. A relatively competent siltstone layer, for example, may deform internally
by a degree of flexural shear and "externally" by flexural slip. A very general rule of thumb
predicts that with increasing depth, and thus ductility, a given layer will show transition from
parallel to similar folding.
For the petroleum explorationist, knowledge of what might be called the structural character of the
stratigraphy in a specific area is essential. Such knowledge will act as a basis for developing an
intuitive sense of what types and geometries of structure may be expected.
Moreover, different fold mechanisms have different effects on the petrophysical character of rock.
The greater amount of internal shear a sandstone unit has suffered, for example, the more its
porosity-and possibly permeability-will be directly related to position on a fold.
More generally, tectonic compaction can substantially reduce porosity along fold limbs and in
hinge areas; pressure solution in carbonates can cause considerable permeability loss. At the
same time, the flattening or stretching of clastic grains can create directional permeabilities.
Pressure gradients produced during folding can also cause fluid migration, with simultaneous
dissolution of material and porosity enhancement. Fracturing can occur at many stages of the
folding process. If interconnected, fractures can result in tremendous amplification of reservoir
quality. These are a few of the main effects that folding can have on prospective reservoir
formations.
As we have mentioned, most exploration takes place at relatively shallow structural levels in the
earth. Thus, competent units (sandstones, thicker carbonates) are usually assumed to deform
into parallel folds (see Figure 5 and Figure 6 ).

Figure 5

Particularly when such units are relatively thick, the extra space created between them in fold
hinge areas is assumed to be filled by intervening ductile material, especially shale, marl, and,
where available, salt or gypsum.

Figure 6

The strong component of flexural slip between competent units is thought to be the major
influence forcing weaker material out of fold limbs and into hinge areas. The actual amount of
material displacement is rarely known; resulting cross sections, therefore, very often portray folds
as parallel, with near-ideal geometry. This is a practical simplification. However, if the
stratigraphic section is well known, folds can be drawn to reflect the flowage of material that we
would expect to occur in the most ductile intervals. This is often necessary in order to accurately
project fold geometry-and thus closure-to depth.
Pronounced ductility contrasts between relatively thin, brittle layers and thicker, incompetent units
often result in disharmonic folding (see part a of Figure 4 ). Here, the wavelength of folds varies
considerably between layers. As mentioned above, disharmonic folding is a frequent and
predictable occurrence in the hinge areas of concentric folds, due to the increased crowding that
occurs with depth.

Figure 4

In certain orogenic provinces (especially fold-thrust belts), this type of deformation is very
common on both a mesoscopic and macroscopic scale. Comparatively thin sandstone or
carbonate layers (1-2 m) within thick shales (50-100 m) are often intensely crumpled within large,
open folds that have wavelengths of a kilometer or more. Such layers thus indicate directly the
amount of flow that has occurred. They are, however, local features, and should not be mistaken
as signs of the general intensity or style of deformation.
In general, overcrowding in the hinge areas of flexural folds can greatly disrupt layering by
disharmonic folding, fracturing, and minor faulting. This will be more true for folds with interlimb
angles less than about 40-50. Within such a highly deformed hinge zone, dipmeter data may
show a relatively sudden change from a continuous pattern to a random, "bag o' nails" motif.
With respect to potential reservoir formations, the concentration of hinge strain can have several
effects. It can, for example, enhance porosity through fracturing, or can destroy it by tectonically
compacting the stronger, more resistant grains. Such strain may also cause the release and
mobilization of solutions that will dissolve soluble material out of the rock matrix and recrystallize

it in fractures, thus sealing them completely. Carbonate material, in a calcareous sandstone, for
example, is particularly susceptible to this type of local mobilization.

Minor Structures Associated with Folding


A great variety of minor, secondary structures generated during folding have been identified by
geologists working in the field and subsurface. These features can be extremely useful for the
unraveling of complex fold geometries. Wilson (1982) offers a comprehensive listing of them.
Only a few of the most common are relevant to subsurface exploration and thus are included in
the discussion below.

Kinematic Axes and the Description of Minor Structures


Figure 1 indicates what structural geologists call the kinematic ("motion-describing") axes of a
flexural slip fold.

Figure 1

These axes are generally used to indicate what a particular minor structure reveals about the
geometry of the host fold. As shown, the orientation of the three axes changes across the fold.
Note that the a-axis represents the direction of slip between layers (called "S surfaces"); the baxis is parallel to the axis of the host fold; and the c-axis is the direction perpendicular to (across)
layering.
The a-axis is sometimes referred to as the direction of tectonic transport, with the a-c plane
describing the plane in which most deformation (folding, flattening, stretching) takes place.
Interlayer slip occurs along the a-b plane.
The general term "S surface" is used instead of "bedding" for two reasons: (1) slip sometimes
occurs within single layers, thus paralleling but not necessarily coinciding with bedding, and (2) in
highly deformed rocks, bedding is not always obvious; secondary planar fabrics (S 1, S2, and so
forth) may exist and need to be distinguished from true primary layering (S0)
Where cleavage is well developed, the kinematic axes are oriented as shown in Figure 2 .

Figure 2

To some degree, this is based on the assumption-usually justified-that a significant portion of later
deformation in such folds occurs by shear along cleavage planes.
Figure 3 shows a variety of common minor structures, including joints, minor folds, cleavage, and
lineations, that can be used to derive the geometry of the host fold.

Figure 3

Among the most important of these to identify and measure are b-axis lineations, i.e., those that
parallel the main fold axis. For example, in structurally complex areas, the intersection of bedding
and cleavage can sometimes be seen and measured in oriented core samples. Elongated
pebbles, flattened ooids, and deformed calcite and quartz grains can all be interpreted in terms of
kinematic axes.

Minor Folds
Figure 4 shows the assumed origin of what are known as parasitic folds.

Figure 4

This is also sometimes called "drag folding," but the more common use of the term drag folding in
the petroleum industry is in reference to folds presumably caused by faulting. Parasitic folds
frequently characterize disharmonic layers within flexural slip folds. Shearing between thicker,
competent intervals is taken up by the development of these structures. Their sense of shear
indicates which limb of the host fold they occupy, and their axes generally approximate that of the
host fold.

Cleavage
This is a complex phenomenon of deformed sedimentary rocks that can be basically described as
a tectonically induced, planar fabric that imparts a mechanical anisotropy to a rock. A
comprehensive review of the mechanisms proposed by various researchers to explain the origin
of cleavage is given by Wood (1974). For our purposes, two basic types deserve mention.
Slaty cleavage, also called "continuous" or "flow" cleavage, results from the parallel alignment of
all platy (clay and mica) minerals perpendicular to the direction of maximum shortening. It

apparently occurs as the combined result of de-watering, grain rotation under stress, and a
certain amount of recrystallization in response to flattening.
Slaty cleavage exists throughout a specified material and completely dominates its mechanical
properties: it is therefore called "penetrative." Where especially well-developed, it accompanies
the low-grade metamorphism of shale into slate. A good slate, such as that from the quarries of
North Wales, could be split with the ideal tools into sheets as thin as-or even thinner than-a single
sheet of paper. In folds, slaty cleavage develops parallel or subparallel to the axial plane, and is
therefore often referred to as an "axial plane foliation" (see Figure 2 and Figure 5 , Slaty cleavage

in a small, overturned fold, Canadian Rockies).

Figure 5

It can also, however, develop as a result of intense shearing between thick, competent units
deformed by flexural slip, or in association with faults. The precise stage of folding at which slaty
cleavage is formed appears to be variable; however, it does seem to be a relatively early

development. Many structures show a fanning in the orientation of cleavage planes that seems to
be caused by continued folding after cleavage formation ( Figure 6 , Progressive development of

slaty cleavage, showing late-stage minor shearing along cleavage planes. Note the thickening of
individual layers as a result of shortening).

Figure 6

Fracture cleavage, also called spaced cleavage, is commonly a form of closely spaced jointing,
and results from the development of discrete planes of mechanical discontinuity within a relatively
competent bed ( Figure 7 , Diagram showing the proposed origin of fracture cleavage due to

intralayer stress generated by flexural shear folding.

Figure 7

Normally S1 fracture plane develops; S2 does not. Continued shearing will rotate fractures
hingeward.). Generally, no grain reorientation is involved. Actual parting may be incipient or fully
developed, and movement along the planes of fracture (which transforms such cleavages into
"microfaults") may also have occurred ( Figure 8 , Sketch of observed fractures in sandstone

layer enclosed by cleaved shale, Delaware Watergap, eastern Pennsylvania).

Figure 8

Fracture cleavage is thought to be the result of shearing within competent units deformed by
flexural slip (Billings 1972; Dennis 1972; Hobbs, Means, and Williams 1976; Park 1983). Its
spacing is related to how brittlely a layer behaved during deformation: massive, competent units
will show wide spacing; weaker, thinner layers may show as much as 150 fracture planes within a
single centimeter (Wilson 1982). Figure 9 shows how the attitude in cleavage planes can vary
with lithology.

Figure 9

Within a single fold, this change across layers is usually termed refraction of cleavage.
As a form of fracturing in competent units, this type of cleavage can have obvious significance for
petroleum accumulation and migration. Once its geometric relationship to bedding is accurately
determined, its orientation can be predicted on the basis of position in a fold. Inversely, it can
itself be used to indicate what portion of a fold is being examined.

Fractures
In addition to cleavage, discrete sets of fractures often characterize folded competent layers.
Stearns (1967,1977) has identified five major fracture patterns useful for a detailed understanding
of brittle behavior in fold deformation. The common occurrence and interconnected nature shown
by two of these patterns makes them important to understand. Their consistent relationship to
bedding attitude indicates that they are, in fact, the result of folding.

Pressure Solution and Stylolitization


In addition to cleavage, certain changes in rock fabric occur as a result of tectonic stress. Two of
these, pressure solution and stylolitization, can have especially large effects on rock porosity and
permeability. Though both are found in association with faults as well as folds, they appear to be
more common in highly folded rocks that have transmitted and absorbed stress internally.

Pressure Solution
This is a phenomenon that involves the dissolution of material at points where there is a
concentration of stress, and the reprecipitation of that material in places of lower stress
concentration. The process of pressure solution is usually explained in terms of Riecke's
principle, which states that (1) material will dissolve under pressure on the sides of objects (single
grains, crystals, pebbles) that face the principal compressive stress, and (2) this dissolved
material will reprecipitate on those sides facing the least principal stress.
Basically, solution occurs because the tectonically induced pressure exceeds the hydraulic
pressure of the interstitial fluid. It is most common along grain, crystal, or pebble contacts, and its
main physical effect is to reduce pore space and more tightly weld a rock together. If extensive,
pressure solution may cause consistent elongation of individual grains or cobbles. Where the
original shape of these is known or can be determined, the principal axes of stress can be
derived.
Pressure solution is a very common feature of folded sandstones and carbonate rocks, especially
limestones. Due to the relatively high water solubility of calcite, material goes into solution quite
readily during intense folding. Quartz grains and pebbles also frequently show evidence of
pressure solution, even in rocks that have been only gently folded. Surrounding conditions, such
as pressure, temperature, and the composition of pore fluids, naturally have a strong influence on
the degree of pressure solution that results.

Stylolitization
This is a special case of pressure solution that occurs predominantly in carbonate rocks to
produce thin, sawtoothlike seams or contact surfaces. In cross section, these seams resemble
the irregular tracing of a stylus. It is said to mark a zone of solution resulting from differential
vertical movement that may or may not be related to the folding process. Some stylolites, for
example, are parallel to bedding and have apparently originated during diagenesis.

Stylolites are especially common in relatively homogeneous carbonate lithologies. Dissolution


involves the removal of the most soluble (carbonate) material from the seam, which is most often
filled by residual insolubles such as clay, sand, iron oxides, and organic material. This sometimes
allows their presence to be detected on natural gamma ray logs as peaks within otherwise
relatively unradioactive carbonate units.
Stylolites are known to be permeability barriers. Where common, laterally extensive, and parallel
to bedding, they can significantly influence the patterns of fluid flow within a prospective reservoir.
The extent of individual stylolites is generally unpredictable, and can range up to hundreds of
meters. More often, however, they are relatively local. They frequently mark boundaries between
similar lithologies of different texture, but can terminate, begin again, overlap, or bifurcate
unpredictably.
Field geologists mapping in areas of limited exposure often make use of a "bedding-cleavage"
rule that utilizes the geometric relationship between bedding and cleavage to help determine
whether a section is right side up or not. The main assumption under-lying this rule is that the
observed slaty cleavage remains roughly parallel to the axial plane of folds.
Briefly state this rule and sketch an overturned fold to illustrate.
On what portion of a flexural slip fold might we most expect tectonically induced porosity loss?

Faults
General Introduction
Much of the basic nomenclature relating to faults was derived from coal mining in the British
Midlands during the late eighteenth and early nineteenth centuries. In fact, the word "fault" itself
was originally used by miners to describe the sudden, unexpected, and trouble-causing
termination of a coal seam. Thus, "fault" then carried much of its vernacular sense; basically,
some sort of a mistake had been made.
Geologists like Murchison and Lyell, however, were quick to realize that a fault was, in reality, a
fracture along which displacement had occurred, and that simple geometric methods could be
used to predict where a particular seam might be found again, either above or below. The
bounding surface of a fault presented a "wall" to the disgruntled miner, who was normally forced
to continue his heading a short way into solid rock and then sink a new shaft in order to relocate
the seam. The wall of the fault plane was almost always inclined, which meant that the miner

could hang his lamp from the rocks above the fault and rest his foot on those below it ( Figure 1 ).

Figure 1

Thus, the terms hangingwall and footwall then as now simply label the two sides of the fault and
imply nothing about displacement. A fault was "normal" to the miner's experience when it was
inclined toward the hanging-wall; a coal seam that ended against such a fault could be found
again if the miner continued his tunnel a short distance in the same direction and then sunk a
shaft downward ( Figure 1 ). (In this area of Britain, such faults are far more common.) At times,
the reverse was true, and thus the corresponding fault was designated as such. These terms
remain in use.
Geologists have long known that faults can extend laterally for much greater distances than folds,
and that they sometimes represent structures of far greater significance. This has been confirmed
by the theory of plate tectonics, which has not only "discovered" and explained a new class of

faults (transform faults), but has emphasized the importance of faulting over folding in general,
especially with respect to deciphering the origin of tectonism.
Structural style often provides a broad context for understanding the patterns of faulting that may
be expected in a region. At the same time, however, the explorationist must generate prospects,
and this sometimes requires a good deal of knowledge about detailed lithologic and structural
constraints.

Description and Basic Terminology


By modern definition, rocks are said to be faulted when they have suffered observable
displacement along a plane or interval of rupture. Such rupture occurs mainly by shear. The fault
plane can be relatively simple (part a of Figure 1 )

Figure 1

or it may consist of a large number of individual offset surfaces and thus be more accurately
described as a fault zone (part b of Figure 1 ). Less frequently, rocks may become displaced by a
form of shearing that causes loss of internal cohesion but not actual rupture of lithological layers
(part c of Figure 1 ). Geologists also use the term shear zone to label fault zones in which the
individual planes of displacement are extremely closely spaced. The great majority of faults,
however, more closely approximate planes of slip along which shearing has taken place as a
result of movement.
The location, style, and geometry of faulting are due to the interaction between the applied
stresses and the initial properties of the affected rocks. Inhomogeneities resulting from
sedimentation or diagenesis, such as the abrupt facies changes at a reef margin or the draping of
units over a barrier bar, can localize faulting. Faults are often genetically related to folding. A
great number of possibilities exist, and it is often helpful to the explorationist to delimit those that
appear relevant to a particular area, since this knowledge can be used to predict or explain the
occurrence of faulting in other parts of an area.
Faults are most generally classified on the basis of their relative sense of displacement. Yet, as
for most of geology, certain settings have encouraged the development and use of more specific
terminology, often related to proposed mechanisms. Each major type of fault discussed
immediately below is given a more specific nomenclature when discussed in relation to a
particular structural style.
Net slip along a fault is measured by a vector that traces the displacement between originally
adjacent points ( Figure 2 , Basic terminology for fault offset, showing strike slip (ss) and dip slip

(ds) components of net slip (ns).

Figure 2

The fault itself is oblique slip; points P and P' were coincident before faulting.). It is most often
resolved into dip slip and strike slip components. Faults in which one or the other of these
components is dominant are correspondingly named.
For nonvertical dip slip faults, geologists find it useful to again divide the displacement into
components, this time into vertical and horizontal contributions ( Figure 3 ).

Figure 3

From coal mining terminology, the former is sometimes called the "throw," the latter, the "heave,"
and the angle between the fault plane and the vertical, the "hade." These terms are somewhat
archaic, and remain more frequently used in mining than in petroleum geology. Petroleum
geologists, and structural geologists in general, more often refer to the vertical and horizontal
separation across a dip slip fault, while the inclination of the fault plane is simply said to be either
high angle or low angle with reference to a horizontal datum. In addition, the normal stratigraphic
separation, shown in Figure 3 , is often important to determine, particularly with reference to log
data.
Part a of Figure 4 shows the four types of dip slip faults (Major types of dip slip faults.

Figure 4

H and F refer to hangingwall and footwall ). In most cases, it is assumed that simple shear has
acted as the principal strain within the fault plane (part b of Figure 4 , Diagram illustrating the

progressive simple shear within an ideal fault plane. Most faults have complex histories of offset
involving intermittent, incremental, and pulselike periods of movement.).
It should be mentioned that fault type may not always be obvious. Where separation cannot be
determined or appears to change along strike, determining the nature of the fault can be difficult.
In some regions, a diversity of interrelated faulting styles exists in close juxtaposition;
differentiating specific structures can prove very difficult.
In addition, many intracontinental basins are characterized by near-vertical, high-angle faults of
small displacement that cannot be easily identified as normal or reverse. Furthermore, as
mentioned, multiple episodes of tectonism often affect a single region. Such episodes may
involve contrasting stress regimes, and displacement along early faults can be reversed. One of

the more important occurrences of this phenomenon involves normal-fault-dominant provinces


being subjected to compressive stresses as a result of changing plate boundary interactions. This
creates "inverted" structures, which we shall look at.
Normal faults can either be planar or listric (concave upward). On seismic sections, planar faults
often appear curved, due to the effect of increasing velocity with depth. In general, however,
normal faults are more easily identified on seismic sections than other types of faults. This is
because of their frequent occurrence in deep, dominantly marine basins characterized by
otherwise relatively undeformed sediments.
Most rupture along normal faults is by shear. Failure due to tension fracturing does occur, but
only at shallow levels where confining pressures are very low. Tension fractures tend to be nearly
vertical and are commonly removed by erosion. Only in areas where faulting has begun very
recently (e.g., certain portions of Iceland) will such fractures be visible.
The term thrust fault deserves special discussion. A thrust has often been defined as a reverse
fault dipping less than 45 However, as used by most geologists today, "thrust" is a genetic term,
said to imply near-horizontal, tangential compression and a zone of movement dominated by
simple shear strain (part b of Figure 4 ). Shortening is involved, and this can be as great as tens
of kilometers or more for single faults.
Thrust faults can (and do) dip at any angle, and individual thrust planes often show complex,
curving geometries. They may be essentially horizontal (along bedding planes) for kilometers, but
curl up at their termini to become nearly vertical or even overturned. In the latter case, thrusts
become apparent "normal" faults, geometrically speaking. Section 6.5 discusses these faults in
detail.
In relation to both normal and thrust faults, we should become familiar with several other
commonly used terms. These were originally derived by German geologists to describe the
structural features that characterize the great valley of the Rhine, as well as those of the Bavarian
and Austrian Alps. Normal faulting on a regional scale is most often referred to as block faulting,
since it results in the creation of horst ("high") and graben ("trench") structural topography (part a
of Figure 5 ).

Figure 5

On the other hand, as shown in part b of Figure 5 , a cover of thrusted rocks can become locally
eroded to the extent that a fenster or window to the underlying footwall lithologies is created, or
an isolated remnant called a klippe is left. Either of these can range from tens of meters to
hundreds of square kilometers in size.
Strike slip faults can be simply classified as left-lateral or right-lateral
( Figure 6 ) on the basis of the shear sense.

Figure 6

When facing the fault, the direction in which the opposite side appears to have moved gives this
sense. The San Andreas ( Figure 7 ) is one of the best-known examples of an active right-lateral
strike slip fault.

Figure 7

Others include the Alpine fault of New Zealand and the Atacama system of may be essentially
horizontal (along bedding planes) for kilometers, but curl up at their termini to become nearly
vertical or even overturned. In the latter case, thrusts become apparent "normal" faults,
geometrically speaking. Section 6.5 discusses these faults in detail.
In relation to both normal and thrust faults, we should become familiar with several other
commonly used terms. These were originally derived by German geologists to describe the
structural features that characterize the great valley of the Rhine, as well as those of the Bavarian
and Austrian Alps. Normal faulting on a regional scale is most often referred to as block faulting,
since it results in the creation of horst ("high") and graben ("trench") structural topography (part a
of Figure 5 ). On the other hand, as shown in part b of Figure 5 , a cover of thrusted rocks can
become locally eroded to the extent that a fenster or window to the underlying footwall lithologies
is created, or an isolated remnant called a klippe is left. Either of these can range from tens of
meters to hundreds of square kilometers in size.

Strike slip faults can be simply classified as left-lateral or right-lateral


( Figure 6 ) on the basis of the shear sense. When facing the fault, the direction in which the
opposite side appears to have moved gives this sense. The San Andreas ( Figure 7 ) is one of
the best-known examples of an active right-lateral strike slip fault. Others include the Alpine fault
of New Zealand and the Atacama system of Chile ( Figure 8 and Figure 9 ,

Figure 8

Examples of major strike slip vaults in various parts of the world.

Figure 9

Dots mark the site of active volcanoes.). Left-lateral faults of such scale are well represented by
the Great Glen fault of Scotland and the Philippine fault. Such large strike slip structures are often
called wrench faults, in general, and transcurrent faults more specifically if they cut across
regional structural trends. In cross section, the symbols "A" and "T" are used to mark which side
of the fault has been displaced away from and which side toward the observer.

Minor Structures Associated with Faults


Like folds, faults are frequently large and complex enough in their deformation to generate a
characteristic set of legible minor structures. These can often be of considerable help in the
analysis of both individual and regional structures.

Slickensides

Perhaps the most common features associated with faulting are the polished and striated
surfaces called slickensides that directly record the sliding movement of rock surfaces past each
other. They commonly occur as parallel grooves, striations, stretched and flattened mineral grains
(especially quartz and calcite), or as combinations of these that parallel the direction of relative
motion when they formed. They have been observed along fault zones involving almost every
type of rock.
Slickensides show a puzzling diversity in their specific characteristics that cannot be simply
attributed to differences in rock type and fault movement. They often show characteristic steps (
Figure 1 ), and are sometimes penetrative over a narrow interval (Drawing of a slickenside

sample in shale from southern New York Note the characteristic steps and ribbing of mineral
grains.).

Figure 1

Within this zone, however, they can indicate more than one direction of movement. Furthermore,
if the direction of fault movement changes, they will become overprinted.

As a phenomenon related to shearing, slickensides are not confined to fault zones, but are also
associated with flexural slip folding, being most well-developed along contacts between thick,
competent units. They can also occur along fractures called shear joints that have suffered smallscale offset.

Drag Folding
A second common minor structure associated with faults is known as drag folding. This is used to
describe the warping of layers in the immediate vicinity of a fault zone. This term, despite its
ubiquity, is often misleading, since it implies that faulting occurs first and causes folding by
friction. In reality, as pointed out by Hobbs, Means, and Williams (1976) and Park (1983), ductile
warping following rupture is far less likely than the reverse. ( Figure 2 , Schematic illustration of

the "preferred" origin of fault-generated drag folding. Prefaulting strain results in offset layers
dipping up or down into the fault plane toward each other.)

Figure 2

Drag is described as "normal," then, if it accords with what would be expected if folding occurred
first; "reverse drag," is in the opposite direction. Here, obviously, is another case where
terminology can easily be confused: "normal" and "reverse" when used to describe drag folding
have no connection to the sense of displacement along the relevant fault. In Figure 3 examples of
normal drag for both normal and reverse faults are given.

Figure 3

Reverse drag is more commonly seen in association with listric normal faults; displacement tends
to roll beds in the hanging wall over to form a gentle anticline. Such "rollover' " has proved
extremely important to petroleum accumulation.
Transverse folds, sometimes of considerable size, as well as both normal and reverse faults are
known to be associated with large strike slip faults such as the San Andreas. Figure 4

Figure 4

and Figure 5 relates the range of these parasitic structures to the mean strain ellipse of a major
wrench fault.

Figure 5

Certain other secondary structures, also generated by folding, are found in association with faults,
especially reverse faults. Shear and extension joints, for example, are frequently related to
faulting. In addition, thrusting is capable of inducing cleavage in incompetent layers, both above
and below the plane of movement (Wood 1974).

Basic Interpretation of Fault-Associated Structures


The specific suite of secondary structures associated with a particular fault is the result of a
complex interaction among the following parameters: (1) the internal properties of the rocks
involved; (2) the initial orientation of these rocks relative to the fault plane; (3) the temperature
and pressure conditions of the surrounding subsurface environment; (4) the rate, magnitude,
direction, and duration of the relevant stresses. Careful study of minor structures will, therefore,
provide us with at least some amount of information about these factors. Obviously, such

information can be useful to an understanding of how faulting may have affected a potential
reservoir interval. Moreover, as in the case of folding, geometric relationships may also become
clear from statistical analysis of secondary features. Fractures, micro-faults, minor folds,
slickensides, fault cleavage, and gouge or breccia are readily discernible in cores, if these can be
recovered. Fault gouge or breccia is often apparent during mud logging.
In the case of fractures and minor faults (fractures along which movement has taken place), their
orientation with regard to the main displacement may or may not be simple and straightforward.
Minor folds are most often caused by fault shear, and thus their axial planes should roughly
parallel the fault plane itself. Cleavage that is generated by faulting often trends slightly oblique to
the fault plane, however. If it cuts through several layers, the line that marks the intersection
between cleavage and bedding planes should roughly parallel the line marking the intersection of
the fault with bedding.

Principal Stress Directions in the Development of Faulting and Fracturing


Figure 1 displays the idealized orientation of the principal stress axes during reverse, normal, and
strike slip faulting.

Figure 1

Because faulting and fracturing both represent the brittle rupture of rocks, they are often
discussed in relation to similar stress systems.
Figure 2 shows the generalized orientation of actual fractures formed in an experimental triaxial
test, in which a block of Solenhofen limestone was shortened by 1 % at room temperature.

Figure 2

The two categories of fractures that geologists usually speak of-shear fractures and extension
fractures-were generated by this test and are indicated. This is how rock generally ruptures in the
laboratory. Note that a broad correspondence exists between ideal faulting and shear fracturing
with relation to stress orientation. As we have said, rocks are weakest in shear.
Two mutually orthogonal types of extensional ruptures are shown in Figure 2 These have been
related to the stresses generated during loading and,

1 and 3 unloading, and, in each case,

they develop normal to 3. Notice that during unloading exchange the orientations they had
during loading. What this tells us is that faults can conceivably develop at very low angles to

1 ,

and that we should expect the relaxation of tectonic stress to generate late-stage features,
especially fractures, in a rock body.
More generally, it can be understood from Figure 1 that a region undergoing a high degree of
faulting is characterized by numerous local stress fields that change in both orientation and

magnitude as diastrophism progresses. The geometry and attitude of particular fault planes,
therefore, may not always appear directly related to the stress axes that apply to large-scale
structural trends. The stress orientations shown in these figures do not, by themselves, always
explain the specific angles at which faults develop or the shape of fault planes. For example,
many fault planes are curved and cannot be explained simply in terms of stress axes alone.
At smaller and smaller levels of scale, many faults reveal increasingly complex components of
displacement, usually involving shear couples. Figure 3 shows the detailed resolution of shear
fractures in an evolving monocline. The fractures shown are those which immediately precede
propagation of a high-angle reverse fault from basement into the overlying strata.

Figure 3

While in their basic geometry monoclines represent one of the simpler geologic structures, the
details of the deformation associated with them are not simple at all.

We should expect, then, that rocks in the vicinity of a fault or fault zone will be complexly sheared
for some distance on either side of the fault plane itself. This distance may be measurable in
centimeters to kilometers, depending on the lithologies, the type of fault, and the amount and
environment of displacement. As with folds, the summation-of-local-strains method is often the
most accurate approach to analyzing complicated fault systems.
In Figure 3 , three sets of fractures are evident: two seem to be conjugate shear fractures; the
third, a bedding plane extensional fracture. If our scale were on the order of miles instead of
meters, there would most likely be smaller, third-order structures associated with the shear
fractures. Thus, we would build our portrait of the overall structure in steps, beginning with what
we could observe and measure most easily.
Figure 4

Figure 4

and Figure 5

Figure 5

give examples of regional strike slip and normal faulting ( Figure 4 shows the resolution of
maximum and minimum principal stresses for the San Andreas fault system, California. Figure 5

is a generalized cross section through a region of extensional (rift) deformation showing the
resolution of maximum and minimum principal stresses.) . Reverse faulting on a comparable scale
occurs most often in long and narrow thrust fault provinces called fore/and thrust and fold belts. (
Figure 6 , Generalized thrust best structure showing variety of faults and folds and the main types

of associated petroleum traps.

Figure 6

Cross section covers horizontal distance of about 100-150 km and is based on detailed studies of
the Canadian Rocky Monitions.) Significant hydrocarbon entrapment occurs in such provinces
throughout the world. Given this and the highly complex nature of the structures involved, we
should look in more detail at the nature of thrusting.

Thrust Faults
For over a century, the existence of low-angle fault planes, along which rocks have sometimes
been transported for up to 100 km, has baffled and intrigued structural geologists. Mechanisms
proposed for such movement have included rigid push, large-scale gravity sliding, and, more
recently, convergence between plates (at subduction zones) and within plates, as well as collision
between continental portions of plates (see Figure 1 , Generalized map and cross section

showing continental breakup along the Red Sea rift and collision in the Zagros region of
southeastern Iran.

Figure 1

Numbers indicate total estimated separation in km between Africa and Arabia). The debate
continues for particular provinces, although the last three mechanisms are now generally
regarded as most important.

The Concept of Relative Shortening


Given the structures of foreland thrust and fold belts above a regional plane of detachment (
Figure 2 , Generalized thrust belt structure showing variety of faults and folds and the main types

of associated petroleum traps.

Figure 2

Cross section covers horizontal distance of about 100-150 km and is based on detailed studies of
the Canadian Rocky Mountains.), it does not seem possible to distinguish between
underthrusting- in which the lower, relatively undeformed "block" is active-and overthrusting-in
which it is the upper block, in which deformation is concentrated, that actually moves. Relative
shortening is what geologists measure in thrust terrains.
In foreland belts, this has been determined to range up to 150 km (Price and Mount-joy 1970),
and is probably greater in certain mountain systems, such as the Himalayas. To some degree,
these provinces pose the culminative challenge to structural geologic analysis, involving, as they
do, the entire spectrum of major and minor structures seen in nonmetamorphosed sedimentary
rocks. Nearly all styles of folding and faulting are present and intimately related. This has both
created structural traps for giant petroleum accumulations as well as destroyed significant
hydrocarbon potential by breaking up and exposing source and reservoir rocks. To understand
how the basic features of thrusting are described and analyzed by geologists, we should become
familiar with a few more commonly used terms.

Thrust Fault Geometry and the Influence of Lithology


Rocks that have been transported from their original location ("root zone") are said to be
allochthonous ("other earth"), while those that remain in place are called autochthonous ("same
earth"). The allochthon is also variously known as a "thrust sheet" or "plate" (not lithospheric), or,

in certain circumstances, a "nappe." Most thrusts define a listric plane that flattens with depth.
Several or more subsidiary thrusts commonly occur within a single allochthon, either as splays off
the "sole" fault or as local ruptures in the cores of anticlines. In general, the thrusts in foreland
belts are very rarely simple listric planes, but are themselves folded and, at times, truncated by
younger thrusts ( Figure 3 , Simplified map and cross section from same general areas as Figure

2, showing complex thrust-fold deformation in the southern Canadian Rockies.

Figure 3

Note the apparent reversals in the direction of thrusting due to refraction of fault planes through
lithologic sequence of variable ductility. Other features on the map include a Klippe (K) and an
anticline that becomes a thrust along strike (A)).
The actual zones of movement may be mylonitic at depth, or, if the thrust passes through a
ductile unit such as a thick shale, very narrow, unbrecciated, and essentially parallel to bedding

for distances that can range up to tens of kilometers. As discussed by Dahlstrom (1970) and Elliot
(1976), the geometry and location of thrust faults in thick sedimentary sequences is largely
determined by the distribution of competent and incompetent layers. Four simple rules and Figure
4,

Figure 4

Figure 5

Figure 5

and Figure 6 summarize this influence (Figure 4: Diagrams illustrating present-day geometry

along a major thrust fault in the Canadian foreland and a detailed reconstruction of its upward
path through various stratigraphic units.

Figure 6

Figure 5: Diagram using the concept of fault preference to show how thrust development is often
directly a function of stratigraphy. Lithologies shown are those of Figure 4. Fault preference is
defined as the length of a thrust within a unit divided by the thickness of that unit, multiplied by an
arbitrary constant. Note here the strong preference at greater depth for major thrusts to exist in
the Kootenay and Fernie shales. Figure 6: Ramp model for thrust development. Note that it is the
lower plate that is shown to move):

thrusting cuts up-section in the direction of displacement (this is often called the direction of
"tectonic transport," or, in older nomenclature, the "facing" direction)

thrusting tends to parallel bedding in incompetent layers, occurring near contacts with
competent units, and to cut obliquely up-section in thicker, more brittle units

the age of major faults is younger in the direction of thrusting


major thrust faults do not overlap appreciably

The Evolution of Thrusting: Ramping, Decollement, and Imbrication


In most fore-land belts, therefore, the overall evolution is for thrusting to begin at deeper levels
and to progress upward and outward (i.e., away from the root zone) from a major sedimentary
basin. Ductility contrast in the stratigraphic section encourages a stair-step evolution of thrusts,
often called ramping ( Figure 6 ) A relatively flat portion of a thrust plane, particularly where it
remains parallel to bedding within a single lithology, is referred to as a decollement, a term we
have used previously to describe the basal detachment that sometimes occurs in parallel folds.
The progressive stacking of thrust faults may develop in "piggyback" fashion-where younger
thrusts form in the foot-wall-or alternatively in "overstep" fashion, where thrusts become younger
toward the root zone. As mentioned, piggybacking appears to prevail on a regional scale and is
by far the more significant progression. At the same time, both piggyback and overstep thrusting
occur on a more local level, often as a result of imbrication.
Imbricates occur most often in two structural positions of high stress concentration: (1) near the
toe of a major thrust, and (2) above ramps in a thrust plane. They dip steeply as they approach
the surface, and stack slice after slice of the same stratigraphic section along listric faults, which
sole out into a major thrust plane. Continued movement along this plane after imbricates have
formed will rotate them so that they can become vertical and overturned.
As discussed by Dahlstrom (1970), imbrication actually offers a basic model for foreland
thrusting: as the scale of a cross section is increased to become more regional, major thrusts
themselves become imbricates of the largest faults (i.e., those with the greatest displacement).
These, in turn, can be thought of as subsidiary faults to a basal detachment or decollement plane
that marks the structural boundary between basement (usually crystalline metamorphic or
plutonic rocks) and sedimentary cover (see Figure 2 ).

Thin-versus Thick-Skinned Tectonics


The concept of basal decollement is the fundamental structural principle in the hypothesis known
as thin-skinned tectonics; this is often contrasted with the thick-skinned hypothesis, which
postulates no sole fault and, therefore, direct involvement of basement in each major thrust (
Figure 7 , lower section, Cross sections illustrating the thin-skinned (upper) and thick-skinned

(lower) hypotheses for regional thrust faulting in the Appalachians ).

Figure 7

The debate between these two schools of thought is a historical one that continues today. On the
basis of drilling and seismic data-both of which have proved the flattening of thrusts at depth in
foreland areas-most geologists now favor the thin-skinned hypothesis for at least the more medial
and distal portions of thrust belts. However, it is well known that toward the metamorphic core of
many such belts, basement rocks are heavily involved in thrusting. Such involvement, as in the
case of the Himalayas, can be very extensive and may, in fact, control the overall style and
evolution of thrusting.
Yet some recent studies based on deep-reflection seismic profiles (Cook 1982; Cook et al. 1979)
have strongly favored the thin-skinned hypothesis for basement thrusting as well. Thus, the
debate has expanded to focus on two major questions: (1) whether thick-skinned faulting occurs
at all in foreland belts; and (2) if it does, what is the nature of the transition between it and the

decollement tectonics that characterize the sedimentary cover. The controversy represents one of
the major areas of research in contemporary structural geology.
Finally, as we have mentioned, the development of foreland thrust and fold belts seems to occur
in pulses. Each episode intensifies the deformation in existing structures and also generates new
features that modify these structures. This superposition of structures can create a number of
seemingly anomalous age and thickness relationships. Figure 8 displays several of these.

Figure 8

Tear Faults

During its movement, a large thrust sheet or reverse-faulted block may tear along near-vertical
planes oriented transverse to the principal tectonic transport direction ( Figure 9 ,

Figure 9

Orientation of major tear faults (numbered dashed lines) in the Jura mountains, French-Swiss
border and Figure 10 , Tear faults along the northeast corner of the Beartooth Mountains,
Montana). Normally, these either form in this direction or at angles to it that invite explanation of
them as conjugate shear faults.
Various genetic schemes have been proposed for tear faults. They have been mapped as striking
at both low and high angles to regional , and thus appear to be generated by compression or
extension within the allochthon. These stresses can be explained as the result of differential
movement. Due to a variety of factors, such as stratigraphic changes along strike, large thrust
sheets apparently advance at different rates over portions of their length (Elliot 1976). In addition,

the direction of transport can also vary along strike (see Figure 10 ); thrust terminations often
show bow-shaped patterns that indicate a type of radial displacement (Dahlstrom 1970; Davis
1984; Lowell 1985).

Figure 10

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

The currently accepted interpretation of transform faults was first put forth by Wilson (1965), who
sought to reconcile the offset of ocean ridge segments and magnetic anomalies with earthquake
first-motion data. Such data indicated that actual movement along the faults was opposite to that
indicated by the apparent physical displacement of ridge segments. Wilson's analysis showed
that this displacement represents an original "frozen" geometry of continental separation. Actual
fault motion due to sea-floor spreading, however, has the opposite sense ( Figure 1 ,

Figure 1

Schematic illustration of transform fault development and Figure 2 , Block diagram showing
relationship between rifting and transform faulting.

Figure 2

Note that offset of rift segments is opposite to the direction of actual displacement across the
transform fault (shown by arrows). Crustal attenuation is also illustrated. ). Continued spreading,
then, will not affect the offset of ridge segments.
Such transform faults have come to be known as ridge-ridge transforms, and represent one of
several possible types of plate-boundary connections. Figure 3 shows two other types of
transforms and the resulting displacement across them.

Figure 3

In Figure 4 , we see examples of a ridge-ridge and ridge-trench transform (San Andreas and Fairweather fault systems, respectively).

Figure 4

Note that the northern corner of the Cocos plate must also be marked by a ridge-trench
transform.
It is speculated that ridge-ridge transforms act to decrease the dynamic resistance to spreading
(Bally and Oldow 1983). This resistance is at least partially amplified by differential rates of
spreading along a ridge axis. The faults are said to provide avenues of low shear resistance for
separation, and to segment the diverging plate margins into orthogonal steps that reduce the
effective surface area. They trend at right angles to ridge segments and therefore trace the actual
direction of plate motion. With respect to subduction and collision boundaries, transforms appear
to develop in response to the differential velocities and directions of converging plates. As such,
they frequently represent a type of metastable condition that will continue until major plate
reorganization (Bally and Oldow 1983).
Where transform faults cut continental crust, the term wrench fault is used. The San Andreas fault
is a well-known example that connects the Juan de Fuca Ridge off the U.S. Pacific Northwest
with the Baja spreading ridge in the Gulf of California. This type of transform is of some
importance to petroleum exploration. Associated folding and normal faulting have created
structural traps for large hydrocarbon accumulations, such as those found in the Bakersfield area
of southern California (Harding 1976).

Basic Influence of Faulting on Logs

Figure 1 ,

Figure 1

Figure 2 ,

Figure 2

and Figure 3 show the predicted effect of normal and reverse faulting on log curves, along with
two field examples.

Figure 3

Figure 3 shows a cutout of section from a well in Alaska's giant Prudhoe Bay field. Figure 2 is
from a well in western Iran and shows, on the other hand, the repetition of section in gamma ray,
resistivity, and dipmeter logs caused by relatively low-angle thrusting. Note how the dip log
responds to drag immediately below the fault. An anticline in the upper limb is clearly indicated by
the upward-decreasing dip pattern, while the fault plane itself is marked by a zone of chaotic dips.
1. A very simple and extremely useful technique in structural analysis is known as "down-plunge
viewing." This recognizes that if complex structures on a geologic map (outcrop or subcrop) are
Viewed in the direction and at the angle of their plunge, a reasonable approximation of a cross
section can be seen.
For example, the map in Figure 1 has been drawn assuming vertical faults and axial planes and a
plunge of 30.

Figure 1

Viewing from the right, with the plane of the paper at about 300 to the line of sight, we see the
units in cross section as they would appear projected on a plane perpendicular to the fold axes.
Down-structure viewing is also very useful for discerning relative displacement along faults. In
Figure 1 this is revealed by a View along the bedding planes in the northern half.
2. All faults in the figure are vertical, as stated. Which are normal and which are reverse?
3. What type of offset does the down-bedding view of faults accurately show? What, then, is the
major limitation with this method?
4. In general, is the down-structure viewing technique better suited to regions of high or low
relief?

6, Joints & Fractures

General Introduction
Like faults, joints result from a brittle response in rocks to stress. They are the most conspicuous
and omnipresent secondary structure of rocks exposed at the earth's surface. In subsurface work,
petroleum geologists use the term "fractures" to refer to local ruptures of almost any kind that do
not show enough offset to be called faults. Geologists in the field distinguish joints from faults in a
similar fashion, i.e., on the sole criterion that no visible displacement has occurred along the
planes of parting. Ironically, the term "joint" was coined over two hundred years ago by workers in
British stone quarries. These men believed that the pleasingly regular, mutually perpendicular
planes were those across which individual blocks of rock were joined.
Most often, joints occur in sets of semi-regular spacing. They may be fully developed and
conspicuous or incipient ( Figure 1 , Illustration of joint face structures and nomenclature).

Figure 1

Frequently related to regional deformation patterns, joints are undoubtedly the source of reservoir

fracturing in many areas. They may also result from non-orogenic stresses initiated by draping
and differential compaction.
Joints are generally classified by nomenclature that reflects particular mechanisms of formation.
Extension joints (a subset of extension fractures) develop perpendicular to 3. Sheeting or
exfoliation joints are basically a type of extension jointing that develops parallel to the surface of
the earth. Such fractures are presumably caused, at least partially, by the release of confining
pressure in a fairly isotropic rock body (e.g., an igneous intrusion) as its overburden is decreased
through erosion. Shear joints, as a form of shear fractures, develop at acute angles to 1.
Within a particular set, joints need not be parallel. Typically, jointing orientation is related to
position on a particular fold. Moreover, two or more sets frequently occur together, comprising a
joint system that essentially splits a specific rock body into an assembled mosaic of blocks.
Microfractures, which require microscopic examination for their identification, are more common
in some competent lithologies. On the other hand, visible joint sets in sandstones or dolomites
can show wide spacing (tens of meters) and can exert a dominant control on surface topography.
Fracturing is a relatively shallow structural phenomenon in the earth's crust and is highly
dependent on lithology. It remains unclear at what depths joints can be generated. On the basis
of mathematical considerations, tension fracturing is predicted to a maximum of about 3 km
(Hobbs, Means, and Williams 1976). However, given a sufficiently low geothermal gradient, high
pore pressures, and a large differential stress (1 -3), both extension and shear fractures can be
produced at considerably greater depths.

Types of Fractures and the Influence of Lithology


As a general rule, factors that tend to increase rock ductility decrease the overall contribution of
rupture to deformation. In a stratigraphic sequence of high ductility contrast, its spacing and
orientation can vary substantially between major rock types. To help quantify and better analyze
this variation, Stearns (1967) has derived the concept of "fracture number." This he defines
simply as the average number of parallel fractures per 100 ft (33 m) normal to the fracture plane.
For this type of measure, any linear distance could, of course, be used. This, then, makes
fracture number a potentially useful quantity for core and microscopic analysis. Average fracture
numbers for several competent lithologies are shown in Figure 1 .

Figure 1

According to Stearns and Friedman (1972), fracture systems that show consistent orientations
and that pervade a large volume of rock can be divided into two broad types: (1) those related to
specific structures, and (2) regional orthogonal fractures. The former type can most often be
explained in terms of the stress system that created the host structure. The latter type, however,
is not well understood. Some geologists have related it to epeirogenic movements, most notably
plateau uplift, but this does not appear to explain most cases of regional fracturing. Both types of
fractures, however, can enhance the reservoir quality of prospective formations.
Fracturing can increase the effective permeability of rock by as much as several orders of
magnitude. It is responsible for the creation of reservoirs in rocks that normally lack sufficient
porosity and permeability to hold hydrocarbons, e.g., shales, quartzites, and even igneous and
metamorphic lithologies. Generally, this permeability increase is greater along extension
fractures, since these often undergo a small amount of separation.

By contrast, there is a component of compression along shear fracture planes. This means that
certain preliminary assumptions about directional permeability can be made in cases where
fractures can be tied to a specific stress system. Usually, this involves an analysis of the
relationship between fracturing and known structures, such as faults and folds.

Basic Techniques for Determining Fracture Orientation


Any determination of fracture orientation should be based on statistical analysis. Several
techniques exist and examples of each for a single joint system are shown in Figure 1 (Part a is a

rose diagram, part b is a strike histogram and part c is a stereogram, showing poles to fracture
planes.).

Figure 1

In cases where fractures are mostly vertical or near-vertical (which is usually the case), rose
diagrams and histograms showing strike frequency are used.
Rose diagrams are often preferred, since they can be plotted directly on maps to show the actual
orientations and relative dominance of different fracture trends ( Figure 2 , Major fracture patterns

in uplift areas of the Colorado Plateau.

Figure 2

Note the general NW-SE and NE-SW trends. These have been interpreted by some researchers
as indicating conjugate failure on a regional scale.). However, use of stereo-grams is essential in
cases where fractures dip at substantially less than 90. Stereograms also allow fracture and fault
or fold data to be plotted and compared on the same figure. This is extremely useful for
establishing structural relationships in fracture analysis. Notice that in part c of Figure 1 , fracture
set Ill closely parallels the host fold axis (point B). As a preliminary assumption, we might

consider this set to represent extensional fractures; sets I and II, therefore, most likely indicate
conjugate shear fractures.

Basic Techniques for Determining Fracture Orientation


Any determination of fracture orientation should be based on statistical analysis. Several
techniques exist and examples of each for a single joint system are shown in Figure 1 (Part a is a

rose diagram, part b is a strike histogram and part c is a stereogram, showing poles to fracture
planes.).

Figure 1

In cases where fractures are mostly vertical or near-vertical (which is usually the case), rose
diagrams and histograms showing strike frequency are used.

Rose diagrams are often preferred, since they can be plotted directly on maps to show the actual
orientations and relative dominance of different fracture trends ( Figure 2 , Major fracture patterns

in uplift areas of the Colorado Plateau.

Figure 2

Note the general NW-SE and NE-SW trends. These have been interpreted by some researchers
as indicating conjugate failure on a regional scale.). However, use of stereo-grams is essential in
cases where fractures dip at substantially less than 90. Stereograms also allow fracture and fault
or fold data to be plotted and compared on the same figure. This is extremely useful for
establishing structural relationships in fracture analysis. Notice that in part c of Figure 1 , fracture
set Ill closely parallels the host fold axis (point B). As a preliminary assumption, we might
consider this set to represent extensional fractures; sets I and II, therefore, most likely indicate
conjugate shear fractures.

Fault-Associated Fractures
Faults and fractures both represent stress-induced rupture of rock, and that when they occur in
association they can generally be related to the same stress field.
Fault movement itself generates shearing stresses that can induce fracturing. In many instances,
therefore, shear fractures can be considered miniature versions of a particular fault. Thus,
knowing the orientation of a fault means that one can sometimes predict associated fracture
trends. This also means that fracture orientation will change with fault attitude.
As a general rule of thumb, faults that develop at shallow levels in especially competent
lithologies (e.g., dolomite) are more likely to have fractures associated with them. However, it
should be emphasized that no necessary relationship exists between the displacement along a
fault and the amount or intensity of fracturing.
Shear fractures are more likely to undergo displacement when associated with faulting than with
folding. Such movement can either increase permeability by creating a poor fit between the two
sides of the fracture, or decrease it by sealing the fractures with gouge or even mylonite. As a
rule of thumb, the intensity of fracturing can be expected to be relatively equal in both upthrown
and downthrown blocks of a normal fault, but somewhat higher in the upthrown block of a reverse
or thrust fault.
According to Stearns and Friedman (1972), several basic rules can be applied to drilling a well
such that the greatest number of natural fractures are encountered. As shown in Figure 1
(Principal fracture patterns and their respective strain ellipses associated with normal and reverse

faults.

Figure 1

In each case, there are three possible fracture sets: two conjugate shear fractures and one
extensional fracture. One conjugate parallels the fault; the other is antithetic to it. Note that the
strike of fractures theoretically parallels that of the host faults or faults.), the strikes of all three
potential fractures will generally parallel that of the host fault. For low-dipping faults, no deflection
of the borehole is needed to intercept the greatest number of fractures. As the fault attitude
steepens, deflection of a well toward the fault plane becomes more necessary ( Figure 2 ,

Diagram showing the dependence of fracture orientation on fault attitude. N refers to normal fault,
R to reverse fault. Angles between shear conjugates are idealized. ).

Figure 2

Fractures can develop during the early stages of deformation and may thus become rotated.
Later folding or faulting of a fractured section may obscure original structural relationships. Most
orogenic regions experience multiple episodes of deformation; thus, later trends often overprint
earlier structures. At times, the true connection between fracturing and faulting will become clear
only after these later deformational effects have been "removed."

Fold-Associated Fractures
Other sets of fractures besides cleavage often characterize folded competent layers. As an
example, Figure 1 shows the variety of fractures seen in a small anticline in southern Germany.

Figure 1

The fold shows an early stage of cleavage formation in finer-grained lithologies, and various
shear and extension fractures in the thicker, more competent sandstone layers.
In all, five fold-related fracture patterns have been identified and analyzed (see Stearns 1967).
Figure 2 shows the derived axes of greatest and least principal stresses for the two most
common patterns (Patterns associated with folding (mostly parallel); note that both patterns show

a consistent geometric relation to bedding.).

Figure 2

As with faulting, these patterns consist of conjugate shear fractures, plus an extension fracture.
Pattern A indicates stretching along strike, parallel or subparallel to the host fold axis. This has
been documented in folds of many styles and scales and is frequently indicated by b-axis
lineations, such as grain or cobble elongation (see Figure 3 , Various common minor structures

useful for analysis of fold geometry.

Figure 3

Shown are b-axis lineations and elongated cobbles; a-axis striations due to interlayer slip; minor
folds with axes parallel to that of the host fold (B); a-c joints; and cleavage planes.). Pattern B is
essentially the same set of fractures rotated 90, and indicates that stretching takes place in the
plane of dip. This pattern should recall the distribution of strain shown in part b of Figure 4 (pure
buckling).

Figure 4

Fracture cleavage, can be thought of as a shear fracture pattern.


According to Stearns and Friedman (1972), both patterns A and B can characterize a single bed.
Individual shear fractures of pattern A often occur as relatively isolated features. They can cross
an entire fold and extend hundreds of feet vertically, but are also seen on many scales, even that
of single grains. They show exceptional consistency in their orientation on all scales, however,
which means that statistical plots show nearly identical patterns, whether data are taken from
aerial photographs or thin sections. Pattern B fractures are smaller (up to several meters long),
but all three fracture sets usually occur together.
With relation to permeability enhancement, the size and isolation of pattern A fractures mean that
any of the three fracture sets might predominate in a well. This lowers the predictability of the
resulting directional permeability. For pattern B, however, as the figure shows, avenues for fluid
communication will tend to parallel the trend of the host structure.

Relationship between Fracture Porosity and Permeability to Structure Curvature


In cases where pattern B is dominant, a set of simple expressions has been derived to relate
fracture porosity and permeability to bedding thickness and structural curvature (Murray 1968).
These are given in Figure 5 (Part a reveals basic equations relating structural curvature and

petrophysical character.

Figure 5

Part b is a graph showing influence of bed thickness on fracture permeability. Note that doubling
this thickness increases permeability by a factor of ten.). The specific structure for which they
were derived was an asymmetric anticline in the Williston Basin of North Dakota. A structural
contour map on top of the productive Mississippian Bakken formation is shown in Figure 6 , and
makes obvious the correlation between structural curvature and well productivity.

Figure 6

What various fracture patterns tell us in a more general sense is that the same body of rock is
often subjected to several different states of stress during the folding process. In every case,
therefore, the specific relationship between a fracture pattern and fold geometry must be carefully
established, usually by statistical techniques.

Wellbore Breakouts and In Situ Stress


Within the past twenty years, regionally consistent patterns of borehole elongation have been
observed in different provinces. More recently, the four-arm dipmeter caliper has been able to
detect the long axis of such elongation. Images processed from ultrasonic borehole televiewer
surveys have, meanwhile, revealed the actual shape and dimensions of the borehole wall. The
televiewer is a tool that emits ultrasonic pulses from a rotating piezoelectric transducer at a rate of

600 per revolution. It produces an "unwrapped" image of the wellbore surface that has proved
useful for identifying and studying natural fractures.
Wellbore breakout is the term used to describe the spalling of rock that appears to create
elongation. To date, the data indicate that breakouts are relatively broad, flat curvilinear surfaces
that enlarge the wellbore on opposite sides to produce a final elliptical shape ( Figure 1 ,

Proposed cause of wellbore breakout).

Figure 1

The two most current and accepted interpretations of this phenomenon attribute breakout to (1)
the intersection of the wellbore with natural fractures, and (2) compressive shear failure due to
stress relief, such that the direction of elongation is parallel to the in situ minimum horizontal
compressive stress ( Figure 1 ) (Zoback et al. 1985). A growing consensus based on recent
analyses strongly favors the second interpretation.

The stress relief hypothesis is especially attractive, since it allows for relatively straightforward
derivation of the basic in situ stress field. This can have obvious importance for explaining and
predicting the orientation of hydraulically induced fractures in low permeability reservoirs.

Unconformities
General Introduction
Unconformities are primary structures whose identification and tracing are as important to
structural geology as to stratigraphy. There are several reasons for this.
First, the period of erosion or nondeposition indicated by an unconformity marks a fundamental
change in environment, one that is very often due to tectonic influence (this, however, must
always be established).
Second, unconformities are invaluable markers for the deciphering of orogenic or epeirogenic
events. Within a thick stratigraphic section, they divide the overall geologic history into a series of
individual subhistories that can, to some degree, be analyzed separately. Mapping the distribution
and characteristics of regional unconformities, in particular, is absolutely necessary to the
understanding of geologic structure. For example, as shown in Figure 1 , reconstruction of
paleoevents by the stepwise removal of tectonic disturbance (sometimes called "pal inspastic
reconstruction") is best done incrementally, by using each unconformity as a paleosurface.

Figure 1

Third, unconformities can be confused with certain types of faults, especially low-angle
decollement thrusts through lithologies that have been subsequently intensely folded.
The principal types of unconformities are all seen on the cross section shown in Figure 2 (Cross

section through Mills Ranch gas field, southern Anadarko basin, showing examples of an angular
unconformity (A) several disconformities (D), and a nonconformity (N)) .

Figure 2

The term disconformity is used where formations are parallel across a surface of nondeposition;
angular unconformity is used when such formations display angular discordance; nonconformity
is reserved for unconformities developed on crystalline igneous or metamorphic rocks.
The term hiatus refers to the total interval of geologic time that is unrepresented at a specific
location along an unconformity. This almost always varies. Along regional unconformities, the
differences can be as much as tens of millions of years. Hiatuses are most commonly measured
by one or more of the following: qualitative geological time units (i.e., stages, epochs);
biostratigraphy; paleomagnetic reversal correlations; and, where data for these are unavailable,
lithostratigraphic correlations (e.g., individual members, formations, or groups). Examples of
change in hiatus and how this is represented on stratigraphic cross sections are shown in Figure
3.

Figure 3

This figure also shows how major unconformities divide a lithologic section into relatively
separable sequences.

Recognizing Unconformities
Unconformities are often identifiable from surface, well log, core, and seismic information.
Erosional truncation that involves discordance between units is very often nicely shown by
seismic profiles. Most exploration, in fact, has come to rely heavily on seismic mapping of
unconformities before drilling begins. At the same time, seismic data can only give us very
general information about the thickness and specific character of an unconformity. Geologic
description, therefore, is most often necessary.

Surface Recognition

Observation of outcrops remains the best means for analyzing the precise character of
unconformities, once they are identified. Recognition, however, is not always straightforward, and
the geologist must often be alert to clues in the form of the detailed effects of erosion. Angular
truncation is one of the most obvious of criteria, and can be evident both in outcrop and in aerial
photos. Discordance, however, is not by itself definitive evidence of an unconformity, as we shall
discuss in a moment.
Marked changes in facies type or in the degree of induration across an identified contact are often
good indicators of hiatus. Telltale signs of erosion, such as a basal conglomerate (containing
pebbles of the underlying formation), chemical alteration (e.g., paleosoil profiles), discoloration,
and irregular relief in the top of a formation, may also exist. Certainly, abrupt truncation of
structures by relatively flat-lying units is diagnostic, as are significant differences in the intensity of
deformation on either side of a contact. At the same time, however, some degree of structural
variation may also result from normal ductility changes within a lithologic sequence, e.g., between
a succession of dolomitized carbonates and a thick shale interval.
Paleontologic evidence can act as a deciding factor. Fossils sometimes indicate gaps of millions
of years within lithologies showing no other obvious changes. In such cases, it must be shown
that faunal successions are, in fact, interrupted, in order to guard against the influence of sample
bias.
In certain areas, as mentioned, unconformities can be difficult to distinguish from faults. Such a
case is shown in Figure 1 (Unconformity or fault? Continuity of a younger unit (Y) above an older

unit (O) supports the first interpretation. This, however, is not conclusive.

Figure 1

). Where beds are tilted, perhaps even overturned, the true nature of the contact is impossible to
discern by outcrop observation alone.
It is often helpful to remember, however, that with unconformities, younger beds will run parallel
to the contact; thus the stratigraphic units above them should not be truncated. Nonetheless,
decollement zones within a single, ductile lithology will not disturb original concordance for as
much as tens of kilometers. Evidence for actual shearing due to movement may exist only within
the immediate vicinity of the fault plane and in the sheared microscopic fabric of the rock. In all
situations of doubt, evidence from as many sources as practicable should be considered before
final determinations are made.

Subsurface Recognition
All criteria discussed regarding evidence for erosion also apply to examination of cores and
cuttings. Narrow intervals of conglomerate do not always indicate paleoerosion zones; however,

where such intervals separate well-indurated units of significantly different lithology, there is a
good chance that an unconformity exists. Again, paleontologic work is often able to determine
gaps in the rock record.
With sufficient supporting information, well log character can be used to trace the stratigraphic
position of unconformities between wells. The detailed character of the unconformity and,
especially, the lithologies on either side determine which specific logs will reveal the most
identifiable response. As mentioned, sharp changes in rock type occur across many regional
paleoerosion surfaces, and this will often influence an abrupt change in log character. Figure 1 ,

Figure 1

Figure 2 ,

Figure 2

and Figure 3 show examples of how gamma ray, sonic, SP, resistivity, and dipmeter logs respond
to unconformities in several large productive fields.

Figure 3

In most cases where actual paleoerosion surfaces are present and identified in a section, the
combination of one resistivity and one porosity log is sufficient to help locate and trace an
unconformity. However, disconformities, especially those within single lithologies-most often
produce no traceable change in log character.
In Figure 1 , note that the greatest change in log character takes place across the regional
unconformity at the base of the Cretaceous (Long curves in the Ninian field, North Sea, showing

multiple unconformities. Lithology is indicated by central column.). Here, lithology changes from
dark, organic-rich shale to limestone. The difference is most conspicuous on the sonic log, which
shows a sharp decrease in interval transit time from the shale into the carbonate, as we would
expect. Likewise, as shown by the change in log response at the top of the Callovian, local
unconformities can just as easily separate unrelated lithologies with very different petrophysical
properties. In the case of angular unconformities, juxtaposed units continually change along the

paleoerosion surface, and thus so does log character, as shown in Figure 2 (SP and resistivity

logs through the Delhi field, northeastern Louisiana, showing multiple unconformities ).
Figure 3 gives an example of how dipmeter logs can help identify unconformities ( Resistivity and

dip logs for well in northern Algeria, showing truncation of azimuth trends at unconformities).
Abrupt discontinuity between dip azimuth trends is characteristic of angular unconformities. If
trends are similar on both sides of the unconformity, however, the change in dip amount,
especially if slight, may appear to represent a depositional structure. At very low angles of
discordance or across disconformities, no significant change may be visible. A weathered zone
associated with a surface of paleoerosion may, however, be recorded as a zone of incoherent
dips if a small correlation interval is used.
As mentioned, unconformities are very often identifiable on seismic profiles. This means that they
can be traced, and their significance to some extent described, early in the exploration process.
Discordance between lithologic sequences can be obvious, as we see in parts a and b of Figure 4
,

Figure 4

or more subtle, as part c of Figure 4 shows (Seismic profiles showing examples of a

nonconformity, b angular unconformity and c relatively subtle angular discordance decreasing to


conformity, right to left ). In most cases, systematic termination of underlying reflections is strong
evidence for either an unconformity or faulting. Figure 5 gives an example where terminations
due to faulting and unconformities can be compared (Uninterpreted and interpreted seismic

section from offshore western Africa, showing a complex diversity of both erosional and structural
terminations in reflection patterns.

Figure 5

In bottom figure, the letters indicate ages of units: TPL=Tertiary, Pliocene; TM=Miocene;
TP&E=Paleocene and Eocene; K=Cretaceous; J=Jurassic; TR=Triassic. Note especially the
angular discordance between unit K1.1 (a deep-water clastic unit and underlying Jurassic strata;

between units TP&E and K1.4, K2; and within the Tertiary section. In addition, an irregular
unconformity separates the basal Jurassic from Triassic beds.).
Where the reflections that mark zones of discordance are continuous but show highly variable
amplitudes, they are more likely to represent erosional unconformities. In many cases,
surrounding structural and depositional patterns will help dictate whether faulting or paleoerosion
has been responsible for the angular relationships observed.
Where lithologic change across a paleoerosion surface is substantial, and the associated
velocity-density contrast is large, the unconformity may itself generate a reflection. This is true for
both angular unconformities and disconformities. Often, the latter are identified by tracing out the
former into deeper basin areas, as shown by part c of Figure 4 .

Structural Styles
General Introduction
The area of geologic study known as tectonics, or sometimes geotectonics, is best thought of as
structural geology in its broadest sense. It examines regional structural features-their
interrelationships, evolution, and effects on sedimentation.
The concept of "structural style" is based on comparative tectonics. Its greatest utility lies in
identifying certain basic patterns of deformation that are repeated in geologic provinces
throughout the world. Studies of structural style have the considerable advantage of relating
these patterns to present-day plate tectonic habitats and thereby to predictive models of origin
and evolution. This means that, prior to exploration, the types of source and reservoir rock,
migration paths, and hydrocarbon traps can be predicted, or at least anticipated.
In classical structural geology, surface mapping, well data, and laboratory experiments were the
main sources of information for understanding tectonic features and their genesis. This has
changed dramatically within the past several decades. Seismic reflection profiling has provided
geologists with a tool that can reveal continuous structural data to considerable depths. It has
opened up both submarine provinces and the deep continental subsurface to eager scrutiny.

Classification of Styles
As developed by Harding and Lowell (1979), Bally et al. (1983), Bally and Oldow (1983), and
Lowell (1985), structural styles are first differentiated on the basis of whether basement is

involved or uninvolved directly in deformation. In the case of noninvolvement, structures primarily


develop within a ""detached" sedimentary cover.
These two basic criteria are, in turn, applied to the four major types of tectonic provinces: (1)
compressional; (2) extensional; (3) strike slip (wrench); and (4) intracratonal (vertical). For the
most part, these last two types of province are thought to involve basement in nearly every case.
Table 1 (below) lists the various structural styles currently used in exploration work.

Structural Style

Dominant

Transportation Mode

Deformational Force
Extensional
Extensional fault
blocks

Extension

Extension

faults" and others)


Salt structures

dip slip of blocks and slab


Subhorizonal to high-angle

Detached normal fault


assemblages ("growth

High to low-angle divergent

divergent dip slip of


sedimentary cover in
sheets, wedges, and lobes

Density contrast

Vertical and horizontal flow

Differential loading

of mobile evaporites with


arching and /or piercement
of sedimentary cover

Shale structures

Density contrast

Dominantly vertical flow of

Differential loading

mobile shales with arching


and/or piercement of
sedimentay cover

Gravity structures

Slope instability

Differential loading
Downslope gliding on
decollement

Compressional

Compressive fault

Compression

High to low-angle

blocks and basement

convergent dip slip of

thrusts

blocks, slabs, and sheets

Decollement thrust-

Compression

fold assemblages

Subhorizontal to high-angle
convergent dip slip of
sedimentary cover in
sheets and slabs

Decollement thrust-

Compresssion

fold assemblages

Subhorizontal to high-angle
convergent dip slip of
sedimentary cover in
sheets and slabs

Strike Slip
Wrench fault

Shear couple

Strike slip of subregional to


regional plates

Vertical
Basement warps:

Multiple deep-seated

Subvertical uplift and

arches, domes, sags

processes(thermal

subsidence of solitary

events, flowage,

undulations

isostasy, etc)
Structural Styles

Primary

Secondary

Extensional fault

Divergent

Convergent boundaries

blocks

boundaries:

Extensional

1. Trench outer slope


1. Completed
rifts

2. Aborted rifts :
aulacogens
Intraplate rifts

2. Arc massif
3. Stable flank of
foreland and forearc basins

4. Back-arc marginal
seas (with

spreading)
Transform boundaries:

1. With component of
divergence

2. Stable flank of
wrench basins

Detached normal fault

Passive

assemblages ("growth

boundaries(deltas)

faults" and others)


Salt structures

Divergent boundaries:

Regions of intense
deformation containing

1. Completed

mobile evaporite sequence

rifts and their


passive
margin sags

2. Aborted rifts;
aulacogens

Shale structures

Passive

Regions of intense

boundaries(deltas)

deformation containing
mobile shale sequence

Gravity structures

Passive boundaries

Convergent boundaries:

1. Trench outer slope


2. Fore-arc basins
3. Back-arc basins
Compressional
Compressive fault
blocks and thrusts

Convergent

Transform boundaries (with


component of convergence)

basement

boundaries:
1. Foreland basins
2. Orogenic belt cores
3. Trench inner slopes

Decollement thrust-

Convergent

Transform boundaries (with

fold assemblages

boundaries:

component of convergence)

1. Mobile flank
(orogenic belt)
of forelands

2. Trench inner
slopes and
outer highs

Strike Slip
Wrench fault

Transform boundaries

Convergent boundaries:
Foreland basins
2. Orogenic belts

Arc massif
Divergent boundaries:
1. Offset spreading centers
Vertical
Basement warps
arches, domes, sags

Plate interiors

Divergent, convergent, and


transform boundaries
Passive boundaries

Table 1: Structural styles and their various plate-tectonic habitats

In our discussion, we will follow the basic organization in volumes II and III of A.W. Bally's
excellent and extensive Seismic Expression of Structural Styles: A Picture and Work Atlas (1983).
This series is itself a major contribution to contemporary comparative tectonics and should be
consulted for more detailed and varied examples of each style.
First of all, we need to define what is meant by "'basement." For petroleum geology, this is usually
taken to be structural basement, which means rigid crystalline igneous or metamorphic rock. The
degree and manner of its involvement in the creation of structures are important to understand,
since these determine the overall context for structural entrapment in overlying sediments. In fact,
in many hydrocarbon provinces, it is the basement structure that is the key to both overlying
deformational and depositional patterns.
These provinces and the structural styles within them are related genetically to one or more of the
following plate boundary classifications:
1. divergent, which includes intracontinental rifts, protooceanic (new ocean) basins, and
oceanic basins;
2. convergent, which includes intra-oceanic arc (oceanic-oceanic) systems, continental
margin arc (oceanic- continent) systems, closing ocean (continent-continent) basins, and
both arc-continent and continent-continent collision zones;
3. transform, which includes ridge-ridge transform faults and wrench faults.
Many of the new ideas about the relationship between orogenesis and plate interactions focus on
the events that occur along convergent boundaries. These boundaries have been divided into two
main types ( Figure 1 , A- and B-type subduction. Two types of A- subduction are shown.).

Figure 1

"B," or Benioff-type, subduction is said to characterize regions where oceanic lithosphere dips
under an island arc or continental margin. These boundaries are Cenozoic in age and have
Benioff zones of shallow-intermediate-deep focus earthquakes. They most often "face" outward,
i.e., oceanward, from the arc or continental margin beneath which lithosphere is subducted.
"A," or Alpine-type, subduction has been invoked to explain many of the world's foreland fold and
thrust belts (e.g., the Canadian Rocky Mountains and Western Overthrust Belt, the Himalayas,
the Zagros Orogenic Belt, the Alps, the Appalachians), which are characterized by continental
crust dipping beneath a regional decollement fold-thrust belt, usually toward a high-grade
metamorphic core. A-type subduction zones generally face continentward; i.e., the direction of
thrusting and overturning of folds within them is mainly toward the continental interior. They are
primarily of Mid Paleozoic to Early Cenozoic age. Most examples have been explained as the
result of continent-continent or continent- island arc collision, and thus as successors to B-type

subduction. Several major provinces of this type, however-most notably the overthrust belts of
western North America-have not been explained, and remain highly enigmatic in terms of their
origin.
Because there are many more examples of B-type subduction than A-type, and since B-type is
"in progress" everywhere, a considerable amount of specific nomenclature has evolved from
recent analyses of its features and processes. The more important terms are shown in Figure 2
and Figure 3 ,

Figure 2

which portray the details of convergent margins (Figure 2: Regional setting of an active B-

subduction zone, including a backarc (rifted) basin.

Figure 3

Figure 3: Enlargement of forearc region in Figure 2, showing common basin and trench
nomenclature.). In general, because it appears that many of the world's major mountain systems
have resulted from collisional events-i.e., they have evolved from B-type to A-type subductionmost researchers feel that an understanding of B subduction will help lead to more accurate and
useful explanations of A subduction.
In our introduction to structural styles, we are by no means limited to discussing regions of
mountain building. As we will see, five of the nine structural styles categorized thus far occur
where diastrophic influence is comparatively slight, even absent. For most styles, we will examine
both "evolving," and older, "completed" tectonic provinces. In addition, field examples will be
given. In each case, it should be kept in mind that structure alone does not define hydrocarbon
potential, nor does the existence of good reservoir rock. Certain regions, such as the Makran fold
belt that we will discuss presently, have all the apparent prerequisites for accumulation, but may

be characterized by low heat flow. Thus, a particular structural style must often be viewed in a
larger tectonic framework in order to best evaluate its hydrocarbon potential.

Extension Styles Basement Involved


Regional block faulting is perhaps the most widespread structural style in the earth's crust,
characterizing a majority of the world's passive continental margins as well as numerous, linear
intracratonal grabens that occur on nearly every continent.
For these cases, block faulting is directly related to the process of rifting (i.e., continental
breakup). In the few possible exceptions (e.g., the basin and range province of western North
America), crustal extension still demonstrates the principal features of rift tectonics and thus may,
in fact, represent unique or anomalous circumstances of this same basic process. It should be
noted, however, that block faulting also occurs along some convergent plate margins and in
backarc settings.
Extensional provinces that involve basement in their deformation can be divided into two basic
types:
Actual rift grabens and upwarps (whether active or failed).
Passive continental margins, which represent one side of a successful rift that has been
buried by contemporaneous and subsequent nonmarine and marine sedimentation
Tectonism in these provinces may be thought of as dominantly tensional in nature. At the same
time, rifting is a complex process; rarely, in fact, during its initial stages does separation occur
exclusively at right angles to the central rift axis. Components of wrenching and compression are
common; the former may be related to transform offset of ridge segments. Individual faults, or
fault assemblages, may therefore show considerable strike slip displacement: local folding and
reverse faulting may exist.
Perhaps the most important characteristic that must be kept in mind with regard to normal faulting
is its deceptive simplicity. Though normal faults most often appear relatively simple in cross
section, they are most often extremely complex in map view. Individual faults can be straight,
cuspate, or can alternate between these. Individual blocks bounded by faults can vary
substantially in size and can be tilted or rotated in different directions to different angles within a
few square miles. In places, faults may appear to die out in ways that seem contrary to the
predictions of geometry or rock mechanics. In a broad sense, much of this can be thought of as

resulting from the intricate readjustment that occurs due to the creation of "extra space" by
extension. In many cases, the precise pattern of structure related to a particular prospect or play
may not become clear until years of detailed work have been done.

Block Faulting in Rifts


Figure 1 shows three basic models that have been proposed to explain the geometry of riftrelated faulting.

Figure 1

Much controversy continues to surround this subject. The degree to which basement-involved
faults show listric geometry appears to vary between regions. Such variation is believed to result
from differences in how rifting evolved and in the type and thickness of crust involved.

Due to the excellent exposure and relative youth of its features, the Red Sea graben is often used
to point out the principal aspects of this structural style. According to Lowell et al. (1975), the
larger, regional faults vary in strike and frequently intersect, but demonstrate a dominant
orientation that parallels the central graben. Two important secondary fault trends appear-to form
a conjugate set that is bisected by the primary trend. Displacement along them, however, is
almost entirely dip slip. Both trends appear to have developed simultaneously during various
stages of rifting. Together, they represent the principal style of rift-related faulting.
Active extension, and thus faulting, appears to have been episodic. As a result, considerable
overprinting of older structures by newer ones has occurred. The cumulative effect is great
complexity on a local scale, despite the overall consistency of major structural trends.
Most faults in this region are interpreted to show listric geometry at depth in cross section ( Figure
2 , Evolution of Red Sea graben.

Figure 2

Steepness of faulting is interpreted to decrease and listric geometry to predominate at depth) and
a curving trace in plan view ( Figure 3 , Gulf of Suez graben: Part a is a block diagram of east

central portion of the graben showing normal fault patterns.

Figure 3

Part b is a rose diagram fault plot for the same region. ). Seismic profiles, however, have been
interpreted to show more or less straight, relatively high-angle normal faults ( Figure 4 ,

Uninterpreted and interpreted seismic profile in western portion of the southern Red Sea.

Figure 4

Note tilting of lower sediments in each graben.), at least in several areas. Since reflection profiles
generally exaggerated the curved geometry of faults at depth (due to velocity increase), the
linearity of the fault planes in Figure 4 appears all the more distinct. The question of detailed fault
geometry, therefore, remains partly unanswered in this region.
Fault-bounded blocks are often rotated in stair-step fashion toward the central graben axis.
Seldom, however, is the structure this simple. Antithetic faults-those that dip opposite to the major
faults (i.e., away from the rift axis)-are also common. These, too, can be thought of as
conjugates. Fault wedges and basement blocks can be rotated in any number of directions, and
both the geometry and magnitude of displacement often vary considerably, even between
adjacent blocks. Sufficient rotation can cause reverse drag folding in hangingwall sediments,
such that beds "roll over" and thus imitate the structure of many growth faults.

In addition, a phenomenon known as "footwall uplift," where the footwall is physically displaced
upwards as a result of isostatic and elastic rebound, has recently been identified as accounting
for about 10% of the total offset along basement block faults (Jackson and McKenzie 1983). Such
uplift can lead to local thickening of sediments over the hangingwall block, thinning over the
footwall block, and to local bathymetric highs that can localize reef growth or other shallow marine
sedimentation.
Figure 2 shows the inferred structural evolution of the southern Red Sea rift as interpreted by
Lowell et al. (1975). We might note that all stages-but particularly those that mark the beginning
of tectonism-involve significant vertical uplift. The broad arching and thinning of continental
lithosphere is accomplished by extensional adjustment in its upper, brittle portion (i.e., the crust)
and is accompanied by high thermal gradients. This results in a central graben with flanking
plateaus that remain elevated above surrounding terrane. These uplifts, therefore, deflect major
drainage away from the embryonic ocean basin but contribute their own coarse detritus in the
form of alluvial fans.
The general lack of clastic influx encourages the accumulation of a thick evaporite sequence.
This can have two consequences of major importance to hydrocarbon entrapment: such
evaporites can provide an excellent seal for underlying porous intervals and can, after
subsequent burial, be mobilized into salt structures that create a host of trapping possibilities.
Figure 5

Figure 5

and Figure 6 show the distribution and nature of the major hydrocarbon accumulations in the Gulf
of Suez portion of the Red Sea.

Figure 6

Note the number of stratigraphic traps, including those associated with the unconformity
separating the Cretaceous and Tertiary sections. This type of major unconformity between
pregraben (generally Pre-Miocene) and graben fill deposits is typical of rift provinces and has
considerable importance to exploration. Its overall significance is best discussed in terms of
passive continental margins.
A look at the Piper oil field in the central North Sea graben ( Figure 7 ) reveals several other
major features consistently seen in this setting.

Figure 7

Figure 8 (Productive fault blocks ),

Figure 8

Figure 9 (Structure contour map ),

Figure 9

Figure 10 (Cross section through Piper field),

Figure 10

Figure 11 (Interpreted seismic profile through Piper field),

Figure 11

and Figure 12 (Sedimentation model )are taken from Maher (1980) and show the relationship of
production to block fault structure in the Piper field.

Figure 12

Note location of the oil-water contact. Given that migration and entrapment most likely occurred
after faulting and that erosion tilted and sealed the Piper formation, the faults within the sand
body are probably nonsealing.
Rifting in this region began during the Early Mesozoic, as part of the crustal thinning and
magmatic intrusion that led to the opening of the North Atlantic. By the end of Cretaceous time,
however, the North Sea arm of the mid-Atlantic rift system had largely failed. The graben,
therefore, has remained an intracratonic basin that has been below sea level since MidCretaceous time.
Cross section and seismic data ( Figure 11 ) reveal how strata are draped over basement blocks.
As shown in Figure 12 , these blocks exercised a dominant control over depositional
environments. In particular, sedimentation of the high-energy, marginal-marine Piper sand was
largely determined by the location of structurally positive horsts. Moreover, later movement along
faults created migration paths into the sand body. Thus, rifting was instrumental in creating
conditions conducive to both the deposition and the favorable structural geometry of source and
reservoir rocks.

These bear detailed comparison to the structural features and patterns of this discussion. At the
same time, their specific tectonic and depositional histories are quite different. For example,
where the East African rift system has had sporadic activity over much of the Tertiary period, the
basin and range has undergone almost 250 km of continuous extension since the Late MioceneEarly Pliocene Age.
In general, rifts have excellent hydrocarbon potential in comparison to other structural styles.
Their evolution generates an abundance of source, migration, and trapping possibilities. These
last relate not only to faults and the drape closures over them, but to salt sealing. Equally
important is the existence of high heat flow, which is regional and continuous for millions of years.
Fault intersections can result in trap-door blocks with overlying closure-an ideal structural trap.
Drag folding along faults may create others. Unconformities created during the starved basin
stages can also act as seals. The highest portions of rotated fault blocks can localize the growth
of reefs or the deposition of shallow-water and porous sands, as at Piper field. Sediments
deposited during these stages often include organic-rich lacustrine and swamp material. These
form excellent source rocks.
Basement block faulting also occurs in settings of plate convergence and in association with
transform faults. This structural style characterizes backarc basins and sometimes the inner
margin of forearc basins. To date, the former have been found to be more petroleum-productive.
With regard to transforms that cut through continental crust, such as the San Andreas, strike slip
motion sometimes generates extension and normal faulting on the cratonal side of related basins.
In the southeastern Great Valley of California, such faults were generated during the episode of
most rapid movement along the San Andreas, and have acted to trap significant amounts of
petroleum.

Extensional Structural Styles in Passive Margins


Passive continental margins represent one side of a successful rift separation. The term passive
refers to the general assumption that most orogenic tectonism has ceased along such a margin.
Passive margins are post-Triassic in age, having developed as a result of the breakup of the
supercontinent Pangaea. They are characterized by seaward-prograding wedges of mostly
shallow marine sediments that can reach 14 km in thickness (Bally and Oldow 1983). These
deposits reflect in their overall character the basic stages of rifting and continental separation.
The ideal stratigraphic section (see Figure 13 , Idealized evolution of Atlantic continental margin

from Triassic to the present, showing progressive burial of rifted basement structures.

Figure 13

The features shown are considered typical of most passive margins , Figure 14 ,

Figure 14

and Figure 15 Block diagrams illustrating the general succession of depositional environments

that develop along passive continental margins characterized by mainly clastic deposition.

Figure 15

) begins with coarse, syntectonic, nonmarine clastics (alluvial fan, fan delta, braided stream),
possibly interbedded with fine-grained, organic lacustrine shales. These sediments are commonly
intruded by, and interbedded with, volcanic material, mostly of mafic composition. This sequence
accumulates in half-grabens between major fault blocks, as well as within the central rift. It thus
becomes broken up, tilted, and rotated as faulting continues.
Evaporites mark the earliest stages of subsequent marine incursion, as the rift expands to
intersect an ocean basin. These may also occur at the base of the rifting section, if marine
incursion occurs at the initiation of breakup. Thicknesses of salt may therefore be confined to
halfgrabens, or may occur as more extensive sheets overlapping these local basins. As plate
separation proceeds, and a new ocean is created, continental margin sedimentation begins and
marine displacement results. A thick wedge of clastic-carbonate deposits then begins to prograde
over the older, largely nonmarine sediments. In most cases, a regional unconformity separates
these two sequences.

Basically, then, the evolution of passive margins can be divided into an early, rift-related phase
and a later, drift-related phase. Earlier structures are predominantly associated with basement
block faulting; later phase deformation is concentrated in the overlying sedimentary wedge and is
controlled mostly by gravity (e.g., growth faulting, salt and shale structures).
The early phase lithotectonic sequence is confined to the Triassic-Lower Jurassic units, while the
bulk of the later, continental margin phase is represented by Cretaceous and Tertiary sediments
and the structures developed within them. The Cretaceous-Tertiary boundary (between units
TP&E and K1 .4, K2) is marked by a sloping unconformity. This is interpreted to have been the
result of submarine erosion, possibly related to changes in spreading rates and the relative height
of sea level (see Vail et al. 1977).
With regard to structure, where the Triassic-Jurassic sequence is characterized by basement
faulting, the Cretaceous sequence shows large-scale detachment faulting, and the Tertiary
sequence is, so to speak, faultless. We might notice how the detachment faulting appears to be
broadly related to deeper basement block faults. These, however, have not been reactivated such
that they cut up into the Cretaceous sequence.
Figure 16

Figure 16

and Figure 17 show a map view and cross section (based on seismic profiles and data from
several test wells) of the Baltimore Canyon area on the Atlantic margin of North America.

Figure 17

Here, we see on the interpreted cross section how sediment loading has presumably depressed
the transitional (i.e., between continental and oceanic) crust and accentuated the continental shelf
edge into a hinge zone. This is typical for most passive margins. In addition, we might note the
complex pattern of relatively long, narrow basins that exists both onshore and offshore.
Figure 18 shows a seismic section through this same general area of the Baltimore Canyon
trough.

Figure 18

Here, note how deep the normal (growth?) faults extend, and also how the reefoid mass has
been tilted strongly landward. This structural framework has been interpreted to be the result of
plastic flow in a limy shale interval beneath the thick carbonate section. These shales were
presumably rendered highly ductile by increasing overburden. With time, they were forced to flow
basinward, thus removing material from behind the reefoid mass. This withdrawal of material
rotated the overlying strata down, while simultaneously pushing the reefoid mass up, creating the
large anticlinal structure and faulting seen. The resulting structural setting appears to offer several
excellent potential structural traps.
Seismic reflection data also appear to indicate that many of the local basins shown in Figure 16
are actually half-grabens, separated by rotated fault blocks and joined transversely by transform
faults (Bally and Oldow 1983, Harding 1984). Figure 19

Figure 19

and Figure 20 offer a close-up view of one such shallow basement half-graben filled by
nonmarine deposits and covered by shelf sedimentation.

Figure 20

Such basins are presumably common features on nearly all passive margins. ( Figure 19 ,

Example of multichannel seismic profiles in the Baltimore Canyon area showing local halfgrabens. Note tilting and deformation of synrift sediments within each half-graben and their
general noncontinuous (nonmarine?) character. Figure 20 : Interpretation of depositional fill in
Newark-type rift basins. Lacustrine (possibly swamp-derived in part) shales are highly organic.
The petrophysical character of alluvial fan and fluvial-deltaic deposits is highly variable. This
same type of sedimentary sequence is presumed to fill basins such as those shown in Figure 19
.)
As indicated by structures that have been studied in the Newark rift system of the eastern United
States, the detailed history of displacement along the faults bounding these half-grabens can be
quite complex, involving substantial amounts of strike slip motion. Figure 3 supports the notion
that, due to the local variability of fault trends, these basins are also likely to be localized and,
therefore, difficult to map along strike.

The nonmarine basin fill of the half-grabens of the Newark system is interpreted as shown in
Figure 20 . The vertical and lateral juxtaposition of organic-rich lacustrine shales with alluvial fan
and fluvial plain deposits may hold significant hydrocarbon potential.
Possible hydrocarbon traps in basement-involved fault blocks associated with passive margins
include those mentioned for rift provinces in general. Deep burial may, however, modify some of
these, as will salt mobilization. The more important role of this tectonic setting, vis-a-vis petroleum
potential, has been to act as the site for the deposition of great clastic wedges (especially deltaic)
that are characterized by relatively shallow detachment faulting, as well as salt, shale, and gravity
sliding structures that have trapped enormous amounts of hydrocarbons.

Detached Normal Faulting


Normal faulting that does not involve basement occurs both as a regional structural style in its
own right, and as secondary deformation to other major styles. Cases of the latter include areas
of extension on the crests of large folds, on the overhanging leading edges of major thrust sheets,
or above diapirs involving salt, shale, or igneous intrusions.
By far the greatest number of hydrocarbon traps of this type have been found along passive
margins in association with growth faulting, so named because of its syndepositional nature.
Other common names for growth faults include "down-to-basin,"' 'contemporaneous," and, less
frequently, "regional" and "down-to-the-gulf" faults. They are best developed in regions of deltaic
or continuous clastic shelf deposition, where sediment accumulates rapidly and remains
semiconsolidated to depths of several kilometers ( Figure 1 , Diagrammatic cross section through

northern portion of the Gulf of Mexico, showing basic condition of metastability (due to density
inversion) that leads to the consistent development of detachment structures ).

Figure 1

High pore pressures in shales (due to retarded dewatering) is a frequently cited influence on the
generation of growth fault structures. Regional crustal subsidence due to sediment loading
provides at least a broad context for extension. However, a great many growth faults are also
directly associated with salt diapirs, which essentially act as large, locally upthrown blocks.
Historically, the Gulf of Mexico has served as the type locality for this structural style.
Figure 2 shows an offshore growth fault, as interpreted on a seismic profile ( Excellent example of

a typical growth fault structure in the Gulf of Mexico).

Figure 2

Some of its more important characteristic features are:

the major fault plane is listric in shape, dips into the basin, and becomes parallel to
bedding with depth;

displacement along the plane increases with depth;


the thickness of individual units increases abruptly in the hangingwall, due to
syndepositional offset;

both secondary synthetic and antithetic faults complicate the basic structure. Many of
the former sole out along the main detachment plane;

back rotation of the downthrown block creates rollover that increases, then decreases
and shifts basinward with depth. (Some of the secondary faulting appears to be related to
extension in the crest of this structure);

in map view, the fault plane is cuspate, and is usually concave toward the basin center;
antithetic faults, however, often show the reverse geometry.
Figure 3 shows the most common hydrocarbon traps that result from growth faulting.

Figure 3

As with rift-related structures, cross sections appear relatively simple, but in map view, faults
intersect, die out, splay, and generally vary along strike to create complex patterns ( Figure 4 ).

Figure 4

In Figure 5 , a dip log from an offshore Louisiana oil and gas field shows the distinct patterns of
growth faulting at two levels. Note the upward-decreasing dip motifs of the two rollover zones.

Figure 5

Also apparent is a highly consistent azimuth trend to the northeast, except in the uppermost
shales. This is most likely due to tilting associated with the faults.
The flattening of faults with depth has been variously interpreted. Geologists have traditionally
pointed to the existence of overpressured, ductile shale intervals into which the faults often seem
to die out. Roux (1977), however, has attributed it to increasing compaction: faults originally
develop at steep angles that are flattened with burial as bed thickness decreases. Of these two
interpretations, the former continues to be preferred.
Shale units in which the effective in situ stress is especially low can be assumed to be preferential
zones for faulting. Pore pressure effects ensure very low shear resistance. This preference also
appears to influence fault geometry. As shown in Figure 6 and Figure 7

Figure 7

, faults flatten out into shale-dominant prodelta slope facies (Figure 6: Generalized line drawing

based on seismic profiles in the Gulf of Mexico showing relationship between depositional facies
and growth fault development.

Figure 6

Figure 7: Diagrams illustrating the proposed development of growth faulting as a result of local
"sinking" of sand-rich facies into underling clay-rich sediments.).
The pattern of such faulting can be directly or indirectly related to preexisting basement
structures, as we have mentioned. The rejuvenation of deep-seated faults by sediment loading
can directly lead to structural adjustment in the overlying cover. Preexisting basement structure
can more indirectly establish the context for growth faulting by influencing the development of
depositional hinge areas. This, in turn, can directly affect the location and trend of detachment
structures.
An example of indirect basement influence is given in Figure 8 , which shows a complex example
of large-scale, growth-type faults developing above the deepest portion of a rifted margin
(Generalized cross section though Baltimore Canyon trough, showing detached faulting above

presumed flowage of limey shales, A and B).

Figure 8

In addition, deep-seated faults can be rejuvenated by sediment loading and this may cause
structural adjustment in the overlying cover. Here, older, fault-generated basement topography
has determined thickness and facies patterns of later sedimentation and, thus, the location of
highly ductile limy shale intervals. Flowage of this lithology under the influence of gravity-induced
stress has played a significant part in the development of growth faulting (see also Figure 9 ,

Interpreted seismic profile through the Baltimore Canyon trough, showing many features
associated with passive continental margins. Note lump of material that has moved downslope
under the influence of gravity.).

Figure 9

Some researchers consider gravity to be the principal cause of growth faulting in areas such as
the Gulf Coast or Niger deltas. With respect to gravity, the maximum principal stress is vertical
and it is the shear component of this stress, as resolved along the regional depositional slope,
that leads to instability. In general, body forces within a volume of sediments deposited over a
sloping surface must be considered significant. Where this volume is very large, as in the case of
passive margin clastic wedges, these forces are very likely to be fully capable of generating largescale fault systems.
At the same time, a variety of other tectonic influences related to salt and shale diapirism,
possible basement fault rejuvenation, and so forth, exist in most passive margin settings. Since
growth faulting seems to be a regional response to overburden and slope-related instability (or
metastability), it may result from a number of specific causes.
Figure 10 and Figure 11 , meanwhile, show the fallacy of trying to strictly segregate structural
styles in every situation.

Figure 10

(Figure 10: Migrated time seismic section showing structure of the Hibernia discovery area,
approximately 250 miles due east of Newfoundland.

Figure 11

Note scale of growth-type faulting and how it cuts deeply into basement. Figure 11: Time-seismic
structure map on lower Cretaceous marker horizon, showing basic structure of the Hibernia
anticline.) The Hibernia discovery was made approximately 250 miles due east of Newfound-land
in a narrow, Late Mesozoic graben (Jeanne d'Arc basin) that cuts into the northern portion of the
Grand Banks platform. The graben developed as part of the active rifting phase that separated
Africa and North America between the Late Triassic and Cretaceous ages. The structural style, at
least in Lower Cretaceous beds ( Figure 11 ), is growth faulting, with large-scale rollover
(compare Figure 2 and Figure 10 ). The faults, in this case, developed as part of the rifting
episode and though syndepositional to some extent, are not all detachment structures. The more
major faults extend deep into the Paleozoic section and most likely continue into basement
(Arthur et al. 1983). Hibernia, therefore, presents a case in which the geometry of features
assigned to a non-basement-involved structural style (growth faulting) also appears in the
features of a basement-involved style (rift fault block).

Salt Structures
The upward movement of salt in the form of diapirs within sedimentary sequences has long been
assumed to be caused by the lower density of salt relative to the surrounding and overlying rock
(Nettleton 1934). This density difference renders deeply buried salt buoyant and mobile. Due to
this general metastable condition, its diapiric rise-often called halokinesis ("salt movement")-may
or may not be initiated by tectonic influence. On the other hand, diapirism is, itself, a
deformational force, piercing, drag-folding, and faulting the sedimentary layers through which it
penetrates.
1. In general, the original deposition of salt is very often associated with rifting, which, in
its initial stages, creates relatively narrow grabens with restricted circulation. Thus, salt
structures are common along passive margins and in rift zones. Most often, the thickness
of deposited salt increases into the host basin or graben depocenter. There is a
progressive change in both the style of salt mobilization and its effect on the overlying
sedimentary cover, which has been described as follows (refer to Figure 1 , Principal

types of salt structures; contours in arbitrary units)

Figure 1

2. A withdrawal of salt from the original depositional edge, with a consequent


downbending of overlying sedimentary layers
3. A contemporaneous pillowing of salt that may result in isolated swells or low,
continuous ridges
4. The formation of low domes and anticlines that can have sufficient structural relief to
pierce overlying sedimentary layers and form local salt stocks
5. The formation of piercement salt walls, diapirs, and detached features
Two well-studied successions of actual salt structures, showing the variation from basin margin to
center, are shown in Figure 2 (Salt structures of northwest Germany.

Figure 2

Salt is from the Permian Zechstein group, a synrift depositional sequence) and Figure 3 (General
inferred relationship between evolution of a basin and that of salt structures for a prograding
continental margin).

Figure 3

Note the basin-ward increase in the size of individual structures. Figure 3 relates the style of
diapirism and resulting salt structures to the rate of sedimentation and, therefore, basin evolution.
As shown, continued regression along a continental margin would presumably cause an evolution
of diapirism from swell structures to spines.
The overall orientation of salt diapirs, however, does not always reveal a consistent relationship
to basin margins (see Figure 4 , Distribution of salt structures in the northwestern Gulf of Mexico).

Figure 4

Where uninfluenced by tectonism, diapirs tend to cluster somewhat randomly, or in subparallel


ridges whose alignment may be determined by overburden. In cases where mobilization results
from tectonic influence (e.g., rejuvenated basement faulting, or thrusting or folding in collision
zones), patterns should appear that roughly indicate deformational trends (Trusheim 1960). Due
to its extremely low ductility, salt (and evaporites in general) can serve as a plane of decollement.
In mobile belts, it has often served to localize deformation, forming the cores of large, flexural-slip
anticlines whose limbs become at least partially detached along contacts with the salt intervals.
A hypothesis that refines the traditional model of diapirism has been proposed by Bishop (1978).
In this view, salt is mobilized as a series of waves flowing both basinward and vertically upward
before a front of prograding overburden. Such a scenario can be inferred from Figure 5 , which
apparently shows the main thickness of salt pushed up in front of the great weight of Late
Mesozoic and Tertiary deposits (Regional NW-SE cross section through Gulf of Mexico).

Figure 5

The traditional concept of salt diapirism has assumed the upward movement of large masses of
salt through a thick overburden. Despite its broad acceptance, this concept suffers from what
might be termed a space problem. Basically, it does not adequately predict what happens to the
column of material displaced by the salt. The observed folding and fault displacement along the
margins of a diapir do not appear to account for the total space occupied by many piercement
diapirs.
In a discussion of this problem, Woodbury, Murray, and Osborne (1980) propose that salt diapirs
do not "jam" their way upward like volcanic plugs, but rather remain at about the same vertical
position relative to the sedimentary cover. As this cover thickens by normal deposition, and the
relevant basin subsides, a diapir presumably builds downward by drawing salt from the mother
bed. The primary motion, therefore, is of sediments sinking past the salt body. Differential
subsidence is assumed to create tension and normal faulting above the diapir. Figure 6 shows a

comparison between the concepts of upward-building and downward growth of diapirs (Sequence

a: proposed for structures in northern Germany, shows active intrusion of salt through overlying
sedimentary cover.

Figure 6

Sequence b: proposed for the Gulf Coast, the "diapir" remains at the same height relative to the
earth's surface and grows downward during concurrent sediment loading basin subsidence.).
Overall, individual halokinetic structures are extremely variable. Their development, location, and
relationship to structures in the basement, as well as the character of the overlying sedimentary
cover, are determined by the specific architecture and history of the basin in which they were
deposited.

Extension above salt structures causes anticlinal warping and normal faulting. The latter is
sometimes termed keystone, for reasons that should be apparent in Figure 7 (Seismic section of

salt diapir showing keystone normal faulting in overlying units ).

Figure 7

In map view, the faults are most often radial in their pattern ( Figure 8 , Structure map showing

complex radial fault system associated with a productive salt dome, offshore Louisiana.

Figure 8

Note greater well density in area of particularly complex faulting, where numerous separate
blocks exist.). However, as with normal faults in other structural styles, local complexity can be
great. Antithetical faulting, for example, is also common.
Figure 9 shows a series of salt structures developed in the Aquitaine basin of southwestern
France, on the eastern passive margin of the Atlantic.

Figure 9

Sediment thickening and thinning due to diapirism, as well as anticlinal warping and the
development of a major unconformity, are all shown as the consequences of salt mobilization. In
addition, the loci of mobilization appear to be over earlier, rift-related basement faults.
The basic scheme of hydrocarbon traps associated with salt structures is shown in Figure 10
(1=anticline, 2=graben caused by extension, 3=porous cap rock, 4=flank sand pinchout,

5=overhang, 6=nonoverhanging wall of structure, 7=angular unconformity, 8 and 9=normal


faulting along flank of structure).

Figure 10

In many cases, the largest traps are in reservoirs not pierced by the salt; for example, numbers 1,
2, 8, and 9 in the figure. This, however, is not a hard and fast rule.

Shale Structures
The conditions that encourage the flowage of shale within a thick wedge of clastic deposits are
more specific than are those for salt. They are primarily related to retarded dewatering of shale
and to the overall sequence of sediments in a regressive (progradational) shelf.
During rapid burial, sands dewater and compact faster than do shales. Impermeability acts to
inhibit water loss. An exception to this occurs at the contact between a shale interval and an
overlying porous sand, but here the expulsion of pore water into the sand creates an even more
impermeable upper seal to continued water loss from the underlying shale. Trapped, excess pore
water keeps the clay mineral grains apart (thus countering lithostatic pressure) and keeps the

rock in a weak, ductile state, susceptible to flow. This is the classic explanation for
"overpressured" shales so commonly and, often, unfortunately encountered along passive
margins buried by thick, regressive clastic sediments.
Along with Jackson and Galloway (1984), we might think of the entire depositional body as a
"spoon" (see Figure 1 , Concept of depositional "spoon," composed of compactible, sand-rich

half-spoon overlying noncompacible shale-rich sequence), composed of a proximal, sanddominated "megafacies" wedge that overlies a more distal, shale-dominated one.

Figure 1

This setting leads to what is known as differential compaction. As the sands compact and
become more dense, they subside in-to the less compacted shales.

Experimental and theoretical work have both shown that the stress resulting from differential
compaction and loading is greatest at the front of the load and that the more mobile,
overpressured material beneath will flow laterally to allow continued load subsidence. In geologic
terms, this means that, due to the depositional slope, shales will be forced basin-ward. This can
lead to growth faulting on a local scale and the creation of shale mounds. Such mounds can
subsequently act to localize sand deposition, as pro-gradation continues, and, thus, the
generation of other growth faults ( Figure 2 and Figure 3 , Evolution of shale mound through

growth faulting).

Figure 2

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

Figure 3

Differential compaction apparently leads to the transfer of great masses of shale, sometimes
ahead of a series of growth faults. Bruce (1973) has described shale ridges in south Texas with
dimensions of over 20 km in length, 40 km in width, and 3000 m in height, with a trend that is
roughly parallel to regional strike ( Figure 4 , part (a) Seismic expression of shale structures.Shale

mounds in northern Gulf of Mexico.

Figure 4

Shale diapirs also seem to be related to gravity sliding along master growth faults, which become
planes of detachment at depth ( Figure 4 , part (b) Relationship between growth faulting and

shale flowage).
Shale structures are not usually of great significance to petroleum exploration (the Beaufort Sea
represents an important exception). However, they are intimately associated with growth faulting,
salt diapirism, and the overall structural evolution of thick clastic wedges along passive
continental margins. This makes knowledge of them necessary to both regional and local
exploration in such settings.

Gravity Sliding
Gravity-induced downslope movement has been invoked by structural geologists to explain an
almost dizzying variety of phenomena on almost all geologic scales. It has been used, for

example, as an interpretation of the rumpled-cover nappe structures of the Alps, the decollement
thrusting of the Canadian cordillera, the chaotic sedimentary sequences known as "melange" (Fr.
"mixture") or "olistostromes," the disturbed layering of single depositional layers, and the
generation of growth faulting.
Gravity structures can very broadly be divided into those that are presumably related directly to
orogenesis and those that are generated as a result of normal deposition. De Sitter (1964) and
De Jong and Scholten (1973) offer comprehensive summaries of the role that gravity has been
thought to play in the tectonics of foreland belts. Much of this, however, has become obsolete
because of recent advances in plate tectonic analysis and theory.
Figure 1 (Gravity tectonics as envisioned for the creation of structures observed in the Apennines

of central Italy.

Figure 1

This generalized evolution shows the basic features previously thought to cause foreland fold and

thrust belt structures.) is a simplified diagram that illustrates the basic assumptions of gravity
tectonics as a mechanism for thrust generation. The principal reasons such a mechanism fails to
explain how laterally extensive sheets of sedimentary cover are transported toward a cratonal
interior are as follows:
1. Basement slopes are usually in the wrong direction.
2. Regions of "tectonic denudation" are not generally seen.
3. Predicted thrust development is chronologically reversed from what is actually
observed in foreland belts.
4. Sliding off uplifted regions would inevitably create relatively local tectonic aprons
incompatible with the vast, laterally continuous thrust belts that parallel plate boundaries
for thousands of kilometers.
5. No such mechanism is observed in any of the world's presently active foreland belts.
As a result, gravity is now invoked to explain specific structures of lesser scale. In many
instances, pore pressure effects are thought to play a primary role in initiating actual detachment.
Tilting due to tectonic uplift will create slopes along which the force of gravity will be resolved as
shear stress. This has been applied to specific decollement structures (see the classic
explanation of the Heart Mountain fault by Pierce 1957), and also to both local and regional
detachment features along Gulf Coast-type passive margins characterized by high rates of
sedimentation. Among the latter are rotational slides observed high on upper delta front slopes
and larger-scale, slower-moving rotational blocks that show growth fault-type structures at their
head. These we will review in a moment. A third common setting for apparent gravity-related
displacement is in thrust belts, where topography can become oversteepened due to the
continued advance of thrust sheets. Here, local normal faults will relieve the resulting stresses,
which are almost entirely due to the simple weight of the material involved (i.e., the maximum
principal stress is gravity and, therefore, vertical).
Certain authors have pointed to the utter disorder in melange terranes as explainable in terms of
gravity sliding. The jumble of variously sized sandstone, carbonate, mafic, and ultramafic (i.e.,
oceanic crustal) blocks within a matrix of shales that show signs of intense shearing seems to
indicate the type of plastic flow and general loss of cohesion that we should expect in a mass that
has moved relatively rapidly downslope. Many melanges are also divided by thrusts that appear

to separate individual "slide" masses, a phenomenon that is predicted by models such as that in
Figure 1 .
More recent interpretations, however, consider melanges to be the result of deformation at the
leading edge of overriding plates at subduction zones (Karig 1974; Seely and Dickenson 1977).
As shown in Figure 2 (Creation of melange.

Figure 2

Shallow structure and process as inferred from seismic profiles and field work in the eSumatra
region. Gravity sliding takes place down the trench slope break, into the trench ) this explanation
proposes the tectonic mixing of lithified sediment riding the subducting crust, as well as splinters
of that crust, into the fine-grained, semiconsolidated material deposited in the trench. Gravity
sliding in the form of submarine slumps down into the rise of the trench slope break are also

thought to occur, but the principal mechanism of melange creation is the "scraping off" and
incorporation of subducting material from the downgoing plate.
The specific character of a melange, in this scheme, will depend largely on the nature and
amount of this material, its initial structural features, the rate of subduction, and the dip of the
subducting lithosphere. Given the variety in each of these parameters throughout the world,
melanges might be expected to show greater similarity in their overall style of deformation than in
their specific characteristics.
Many geologists now believe that the generation of certain foreland fold and thrust belts is a
result of passive margin clastic-carbonate wedges becoming involved in accretionary prism-type
diastrophism during the early stages of continent-continent collision. If this be the case, the
direction of tectonic transport would be toward the subducting plate (see Figure 3 (b) The
megasuture is due to continental collision and thus represents a successor from B-subduction).

Figure 3

In relation to normal deposition along passive, divergent margins, so-called thin-skinned gravity
sliding is seen to occur in close association with growth faulting and shale diapirism (Jackson and
Galloway 1984). Two proposed versions of the structural evolution which might be responsible for
this close interrelationship are shown in Figure 4 ,

Figure 4

Figure 5

Figure 5

and Figure 6 ( Figure 5: A proposed evolution for gravity structures interpreted to exist in Zone

4a, Figure 6 : Proposed evolution for gravity structures interpreted to exist in Zone 4b.

Figure 6

) These have been derived from detailed study of the Mexican Ridge area in the western Gulf of
Mexico. ( Figure 7 , Tectonic map of southern Mexican ridges area. Note convex-outward pattern

of structural trends.) shows the distribution of the relevant structures.

Figure 7

Note how these vary between zones 4a and 4b, and how the two proposed evolutionary schemes
attempt to explain this change. Zone 4b, which is characterized by growth faulting, has a slightly
steeper slope.
In the view of most researchers, each glide sheet in this type of environment has a growth fault at
its head, a decollement along its principal length, and a thrust or faulted fold at its terminus. To
some degree, these sheets can be likened to giant slumps, with growth "scarps" marking the
upper planes of rotation. We should notice the ponding of sediments between and over the
ridges. The loading from this accumulation is thought to effectively stop gravity gliding and help
initiate shale diapirism.
Figure 8 (Rollover due to gravity gliding in Texas Gulf coast.

Figure 8

Trapping is due both to rollover and fault sealing.) The irregularity of the glide plane ( Figure 9 ,
Large -scale cross section showing regional structure that includes the features shown in Figure 8
), which is partly due to shale diapirism, acts to localize the degree of bed rotation.

Figure 9

Movement along the basal detachment is assumed to be by continuous gravity creep. Note how
the top of the Vicksburg formation becomes conformable in the direction of gliding.
As with shale structures, these gravity-induced features are important to petroleum exploration
mainly for their relation to growth faulting. The anticlines caused by the type of downslope
movement shown in Figure 4 , Figure 5 and Figure 6 faulting should be considered potential
hydrocarbon traps in deeper-water areas off passive margins.

Compressional Styles - Basement Involved


The two principal elements of basement-related structural styles are compressional fault blocks
and their bounding basement thrusts, which may range in inclination from near-vertical to less
than 30. Both high and low angle faults occur along a single thrust front and are now thought to
represent the end members of a continuous spectrum of basement-involved deformation.
Reverse faults involving basement generally occur along convergent plate margins, primarily in
foreland regions characterized by A-type subduction, and in the inner trench slope and outer high

portions of B-type subduction zones. Of these, the foreland regions are by far the most important
to petroleum exploration.
Basement-involved forelands primarily develop in two major settings: between a volcanic arc and
a craton (i.e., in a "backarc" setting), and in front of the arc during continental collision. The first of
these is particularly characteristic of the North and South American cordillera, while the latter
typifies the great Alpine-Himalayan orogenic zone.
Compressional basement block faulting is particularly well known and well studied in the Rocky
Mountain foreland region of the western United States
( Figure 1 , Principal basement uplifts and associated basins of the Rocky Mountain foreland

region Cross section lines refer to Fig 3).

Figure 1

In this area, large, rigid masses of Precambrian crystalline rock have been forced up and, to
some degree, laterally outward many thousands of feet. The larger uplifted blocks appear to be
intimately associated with deep foreland basins: in general, each block has its related basin(s),
whose timing of subsidence seems directly related to uplift history. Most often, these basins are
strongly asymmetric, with their structural axes running close to the thrust front and their back
flanks forming relatively gentle basement slopes away from the uplift.
These basins have served as the sites for considerable sediment accumulation and subsequent
petroleum generation and entrapment. Structural relief between the present crest of basement
uplifts and their associated basins is frequently on the order of 10 km or more. Thus, the vertical
component to faulting is undoubtedly considerable. The overlying sedimentary cover on these
blocks has usually acted in a passive manner, accommodating basement uplift by drape folding
and brittle fracture. Blocks are typically bounded by faults on both sides and are tilted at various
angles.
Figure 2 (Basement faulting in the Rocky Mountain foreland region) shows four styles of faulting
that have been proposed to explain the geometries observed in the field and in seismic profiles.

Figure 2

To some degree, the differences between these reflect the fact that various investigators have
concentrated their efforts in different parts of the region, along different major fault zones. Such
zones display variable geometries, which range from reverse listric planes (steepening with
depth) to low-dipping thrusts of 30 or less. ( Figure 3 , Two interpreted seismic profiles from

Wyoming showing expression of rigid basement uplifts.

Figure 3

(a) Southwest Wind river fault; (b) Casper Arch thrust. Note how these "end member" styles of
faulting correspond to (b) and (d) of Fig.2) Single faults apparently show both these geometries
along their strike. At present, the proposed styles of faulting reflect two major schools of thought.
These agree that compression is involved to some degree in deformation, but part company on
the question of whether displacement is dominantly vertical, "upthrusting" of Figure 2 parts (a)
and (b), or horizontal "overthrusting" of parts (c) and (d). The seismic profiles given in Figure 3
illustrate how apparently irreconcilable these two schools are.
Many of the principal faults have also apparently suffered a degree of ancillary strike slip
displacement. For these reasons, some geologists now think that at least several of the major
blocks, such as the Wind River, have been uplifted in a type of scissorlike rotational pattern, with
the inclination of the fault plane decreasing and the amount of lateral movement increasing with
distance from a 'nodal" zone of almost total vertical uplift. This has, in turn, been related to the

clockwise rotation of the entire Colorado Plateau during Laramide time (see, for example, Gries
1983).
With respect to those blocks bounded by near-vertical faults, Harding and Lowell (1979) describe
three basic structural levels from basement up into the sedimentary cover:

a tilted fault block of basement and immediately overlying units.


an intermediate level where the principal fault becomes a zone of steep drag folding.
an uppermost level of gentle drape or monoclinal folding.
The structure of fault segments can be complex, involving normal faulting along the crest of the
block and drape fold, tear faulting of the cover and basement, and more minor reverse faulting in
the foot-wall of the main fault.
Some geologists have pointed to older zones of Precambrian shearing within the basement itself
as a probable control on the location and possibly geometry of faults. Certainly, in any region
where basement faulting is a dominant structural style, the geologist should understand the
preexisting structural character of the rocks involved. Metamorphic rocks are, in general, the
products of orogenesis and very often contain important planes of weakness, such as shear
zones and major lithologic boundaries, that will have some amount of influence on all later
deformation.
Figure 4 (Regional east-west cross section through the Rocky Mountain foreland, from

northeastern Wyoming to southeastern Idaho.

Figure 4

The deep crustal faulting shown for the Wind River Mountains is based on recent COCORP
seismic reflection profiles. Note that faults are interpreted to reach to the crust - mantle boundary.
Western end of section shows decollement thrusting of the western overthrust belt ) shows
basement faults as both relatively planar and as listric to the crust-mantle boundary. Compressive
decoupling is, therefore, proposed to occur at or near the Moho. At present, this interpretation is
highly speculative. It does, however, take account of the prevailing hypothesis of intracontinental
underthrusting (a form of A-type subduction).
The most generally accepted idea about the genesis of these basement foreland uplifts is that the
principal compressional stresses are in some way directly related to B-subduction, as appears to
be the case in the South American cordillera ( Figure 5 , Basic tectonic setting and regional cross

section through northwest Columbia, showing three principal tectonic divisions.

Figure 5

Note the extensive involvement of continental basement in faulting of the eastern cordillera ).
Here, an A-B subduction couple seems responsible for the opposing directions of tectonic
transport.
Broadly speaking, the vertical rise of large, rigid tectonic blocks has generated a good deal of
local variability, and thus a considerable diversity of structural traps should be expected ( Figure 6
, General hydrocarbon trapping possibilities associated with basement block uplifts).

Figure 6

The Elk basin oil field in the Big horn basin of northern Wyoming is an example of one of the
larger traps discovered thus far. To date, it has produced over 500 million bbl out of faultcontrolled closures such as those shown in Figure 7 (Uninterpreted and interpreted seismic

sections through South Elk basin producing area (northeast Big Horn basin, Wyoming ), showing
anticlinal folding over basement thrusts.

Figure 7

Note how deformation in this section appears to correspond with style (c) in Fig 2) and Figure 8
(Cross section of Elk Basin field approximately 10 miles north of the seismic profile shown in Fig

7 ).

Figure 8

In more recent years, some companies have attempted to penetrate several of the lower-dipping
Precambrian thrusts in order to explore the sedimentary section beneath, certain formations of
which are known to be productive in nearby domes and folds. To date, only a few of these
expensive wells have been successful. Oil and gas leaks within the thrust zones, however, have
been reported.
More generally, in addition to providing drape and fault-related closures, basement uplift can also
influence petroleum generation and trapping in more subtle ways. For example, in the Hardeman
basin of northeast Texas, high-angle splinter faults related to late Paleozoic uplift of the Red
River-Matador Arch and the Wichita Mountains created local avenues for the invasion of
dolomitizing solutions into the cores and flanks of Mississippian bioherms. As a result, these
became excellent, though highly localized, reservoirs.

Another possibility involves the inversion of rift-related block faults during later compressional
tectonism. Convergence of older passive plate margins can lead to the rejuvenation of originally
normal faults into high-angle reverse faults or thrusts. Such reactivation, since it reverses the
sense of displacement. is said to create "inverted" basins. This can be crucial to understand in
certain regions, since the sediments normally assumed to characterize rift provinces will become
involved in anticlinal, thrust, and compressional-wrench tectonism. In such cases, apparent horstand-graben morphology may be retained; however, the more recent uplift will have substantially
altered the original structural configuration.

Compressional Styles - Basement Not Involved


The primary structural style in this category is the decollement foreland thrust-fold belt. Due to the
thorough work of such authors as Price and Mountjoy (1970), Dahlstrom (1970), and Bally,
Gordy, and Steward (1966), the Canadian cordillera has come to be generally treated as a type
locality for such deformation. Figure 1 (shows basal decollement and telescoped nature of

regional thrust deformation.

Figure 1

Note the interpreted involvement of matamorphism in thrusting toward the west, i.e., closest to
the hot, mobile core of the orogen ) is the classic cross section by Price and Mountjoy (1970)
through the southern portion of this structural belt and shows most of the principal features we
have come to expect in this style. The majority of hydrocarbon production exists in the foothills
region, where faulting is especially complex. Figure 2 is a cross section through the
Jumpingpound gas field which amply shows this complexity.

Figure 2

Note the degree of imbrication above the ramp anticline, where closure is best: this is the primary
trap in this and many other fields of the foothills area.
Thrusted anticlines appear to be the most common traps in productive foreland regions
throughout the world, but with regard to the size and specific geometry of traps, it should be
emphasized that considerable variety exists within and between these regions. Over the span of
Proterozoic time, differences in the size and morphology of plates, in their marginal sedimentary
character, inherited structural features, and specific motions have all ensured a high degree of
structural variation in foreland orogenic zones.
Like compressive basement faulting, decollement thrusting is most often related to processes
occurring along convergent plate boundaries, both in the mentioned foreland belts and along the
leading edges of subduction zones. We have already discussed some of the details of B-type
subduction zones. Figure 3 displays the interpreted structure on a seismic section across the

active subduction complex offshore of northern California.

Figure 3

This section shows the decollement faulting in the accretionary prism quite well.
Note the change from the chaotically deformed accretionary prism sediments into the more
conformable forearc basin at the far right. The deformed complex is broken into thrust slices
about 400-500 m (1300-1600 ft) thick, whose bounding faults display variable curvature. This is
probably due to a degree of stratigraphic control and ramping of fault development. We can also
see that where compressional features dominate the accretionary prism, extensional faulting
characterizes the deep sea sedimentary section west of the trench. This has been observed in a
number of active subduction zones. Often, these faults have developed in the underlying ocean
crust and have been interpreted as being the result of lithospheric bending and consequent
tensional rupture.

Collision-related thrust belts verge toward the subducting ("on-coming") plate and represent the
succession of B-subduction by A-subduction. ( Figure 4 , a map showing distribution of the major

collision-related fold and thrust belts and related foredeeps (regional foreland basins.

Figure 4

) Many of the world's foredeeps (Appalachian, Canadian Rocky Mountain, Arabian, Uralian) are

highly productive of oil and gas. Cross section lines refer to Figure 5.) A transition to basementinvolved thrusting often occurs in these settings behind the thrust front. Figure 5

Figure 5

and Figure 6 (Generalized cross section through four of the world's major collision-generated

mountain systems at comparable scales) show how this involvement is usually interpreted to
increase into the mobilized,

Figure 6

metamorphic "core"' of an orogen, such as the Canadian cordillera. In the case of trench
deformation, the situation is less well understood. Two forms of basement-involved faulting are
presumed: that which incorporates "slivers" of ocean crust (known as ophiolites) into the
accretionary prism, and that which cuts continental crust. For both trench and foreland thrust
zones, then, overlap occurs between detached and "connected" structural styles.
Most thrust belts presently exposed at the earth's surface occur as sinuous belts up to thousands
of kilometers long. Their width is not uniform, but is instead characterized by sharply recessed
and more gently extended portions known, respectively, as "reentrants" and "salients" (see Figure
4 ). The cause for this type of variation along tectonic strike is not well understood, but often
thought to be related to preexisting basement features.
Figure 5 and Figure 6 compare the overall features of several major collision-generated foreland
regions. The Alps appear to have resulted from a series of collisional episodes between the

Eurasian continent and various arclike continental fragments that once bordered it to the south.
The Zagros orogen, as we have noted (see Figure 7 , Generalized map and cross section

showing continental breakup along the Red Sea rift and collision in the Zagros region of
southeastern Iran.

Figure 7

Numbers indicate total estimated separation (in km) between Africa and Arabia), is the continuing
result of Eurasia colliding with the much smaller Arabian continent. Both the extensive Himalayan
Mountain belt and the Appalachian-Ouchita-Marathon system are interpreted as the result of the
impact between larger continental masses-in the first case, India and Eurasia, in the second,
those Paleozoic continents (proto-North America, South America, Africa, and Europe) whose
collision marked the early formation of Pangaea. In both the Alps and the Himalayas, great
thicknesses of crystalline basement rocks are involved in the deformation and have exerted
considerable control on resulting structures. Though decollement thrusting is evident in the Jura

Mountains of western Switzerland and eastern France, it does not dominate the regional
structural style of the Alps as it does the Appalachian, Zagros, or Canadian forelands.
At the same time, however, we can also think of the Alps and the Himalayas as two opposite endmembers with respect to the general style of deformation The the sedimentary cover. The central
and southern European Alps, with their spectacular development of thrust and fold nappe
structures, represent the most highly contorted foreland region in the world. Broadly speaking,
they show the effects of very rapid, collision-related diastrophism on a thick, only semicompacted
sedimentary pile. Folding of a highly ductile nature is common; recumbent and overturned nappes
are piled up on top of each other like rumpled carpets to form the higher ranges (see Figure 8 ,

Alpine nappe structures). Sediments appear to have been squeezed up and out of forearc and
backarc basins, as the various island arc fragments were progressively welded to the Eurasian
continent.

Figure 8

The Himalayas present a very different case. Here, great thicknesses of well-indurated Paleozoic
and Early Mesozoic sandstone and carbonate rocks were involved in the deformation. This
generally created less contorted, more widely spaced, and far more massive structures.
As in the Canadian Rocky Mountains and the western overthrust belt of the United States,
ramping is well developed in the Appalachian foreland. This has been interpreted as being the
result of both ductility contrast related to stratigraphic variation and preexisting block faulting in
the basement.

The well-exposed, dramatic structures of the Zagros foreland appear to resemble those of the
Appalachians more than those of the Alps or Himalayas. Some of the larger folds stretch for as
much as 160 km (100 miles) along strike before plunging beneath the surface. The existence of
salt layers in the lithologic section has resulted in a high degree of complex, local decollement. In
places, for example, slip has occurred along and within salt intervals. This has apparently created
shallow anticlines that have subsequently been pierced by the tectonically mobilized salt.
In discussing foreland thrust belts, there are a number of general aspects that are of direct
importance to hydrocarbon exploration:
1. Anticlines occur primarily in the hangingwall, and are asymmetric toward the direction
of tectonic transport.
2. The geometry of thrusting is much dependent on the competence of the sedimentary
units involved. The number of thrusts, as well as the intensity of folding, decreases in
distinct proportion to increases in the amount of thick, competent units (e.g., sandstone,
carbonates).
3. Sedimentary thicknesses decrease and deposits change character to more shallowwater clastics in the direction of tectonic transport, i.e., away from the metamorphic core,
or, in stratigraphic terms, toward the basin margin. This usually results in an increase in
the overall density of faulting in the same direction.
4. Structures become progressively younger in the direction of tectonic transport.
5. Rocks involved in thrusting also become younger in this direction.
6. Advancing thrust sheets act to load and depress the crust into local fore-land basins
(sometimes called "molasse" basins). These fill with coarse marine and nonmarine
detritus, which then also becomes involved in deformation. Such basins are often rich in
plant-derived organic matter.
7. In many cases, a regional foreland basin, relatively rich in petroleum, will exist
immediately out in front of the thrust belt, presumably created by crustal loading on a
more massive scale.
Thus, deformation appears to begin in the deeper, thicker portions of the sedimentary wedge and
progress upward and on-to the craton. In a broad sense, this has meant that hydrocarbons have

had the best chance to accumulate and remain undisturbed near the youngest, leading portions
of thrust belts and in the regional foreland basins out in front of them.
As mentioned, traps in thrust belts are mainly associated with asymmetric anticlines ( Figure 9 ,

Seismic profile through the eastern Po plain in northeastern Italy (approximately 50 km southwest
of Venice), showing thrust structure of the Apennines.

Figure 9

This profile reveals the thrust belt at its widest point. Gas pools exist in Pliocene sandstones that
wedge out against the rising structures. Oil is produced from complex structural traps in
underlying Mesozoic carbonates) and, to a lesser extent, fault truncations. Closures are most
often at less than about 3000 m depth (Harding and Lowell 1979). The actual size of individual
traps, and their height of closure can vary a great deal, depending on the spacing of thrusts and
the degree of asymmetry in folds.

Substantial-even giant-accumulations have been discovered in a number of the world's foreland


regions. Perhaps the most impressive example of production from thrust-fold structures is offered
by the oil fields in the Zagros Mountains. Here, reservoir quality is due to an extensive,
interconnected fracture system generated by the folding. Production is from the Asmari
Limestone, which, by itself, yields more than 75% of all petroleum currently being recovered from
traps in foreland belts. In addition, several recent major discoveries along the Idaho-Wyoming
thrust belt have encouraged continued exploration and have caused geologists to take another,
more detailed look at the petroleum potential of other foreland regions, such as the AppalachianOuchita system.
With regard to active subduction zones, several general statements can be made concerning
overall hydrocarbon potential. In the forearc, for example, both source and seal can be more than
adequate, but the large amount of volcanogenic material generally makes for rather poor
reservoir quality. Not only is the sediment matrix often fine-grained, but both primary and
secondary pore-plugging and swelling clays-most notably illite and montmorillonite-are abundant.
Moreover, low heat flow is characteristic of the forearc; this is even more the case in the forearc
basins that develop over the subduction complex itself. Structural traps, however, abound and
consist of anticlinal and thrust closures and relatively shallow drape folds above thrusts.
The problems mentioned for forearc regions appear to characterize the Makran subduction
complex, which, because of its unique setting, may otherwise appear to offer relatively strong
hydrocarbon potential ( Figure 10 ,

Figure 10

Interpreted structure of the Makran accretionary prism, Gulf of Oman,see Figure 4 for
approximate location) This very large accretionary prism is basically the continuation of the
Zagros collision zone into a B-type subduction zone that stretches nearly 900 km eastward from
the southwestern coast of Iran to Pakistan. It shows the development of many thick, coherent
thrust-fold structures whose amplitudes are unusually large and whose petroleum potential might
therefore seem to be substantially greater than that in most other forearc arc systems. As much
as 7 km of mostly late Tertiary abyssal plain sediments (most likely resulting from high erosion
rates that began with the India-Eurasia collisional event to the north and west) have been
deformed into an imbricated stack.
Despite this structurally attractive setting, potential appears moderate at best due to relatively low
geothermal gradients, which average 1 per 30 m (Harms et al. 1983). Gas seeps and traces of
heavy hydrocarbons, however, have been found. At the same time, wells drilled in coastal
Makran have encountered very high pressures. Because of the young age of the sediments

penetrated (Pliocene at total depth), their rapid accumulation (about 300 m per my), and their
subsequent deformation, we can expect pore pressures to be quite high. As a whole, then, the
combination of remote access, potential drilling problems, and low-to-moderate maturation
potential makes most active forearc systems relatively high exploration risks at present.

Compressional Styles - Basement Not Involved


The primary structural style in this category is the decollement foreland thrust-fold belt. Due to the
thorough work of such authors as Price and Mountjoy (1970), Dahlstrom (1970), and Bally,
Gordy, and Steward (1966), the Canadian cordillera has come to be generally treated as a type
locality for such deformation. Figure 1 (shows basal decollement and telescoped nature of

regional thrust deformation.

Figure 1

Note the interpreted involvement of matamorphism in thrusting toward the west, i.e., closest to
the hot, mobile core of the orogen ) is the classic cross section by Price and Mountjoy (1970)
through the southern portion of this structural belt and shows most of the principal features we
have come to expect in this style. The majority of hydrocarbon production exists in the foothills
region, where faulting is especially complex. Figure 2 is a cross section through the
Jumpingpound gas field which amply shows this complexity.

Figure 2

Note the degree of imbrication above the ramp anticline, where closure is best: this is the primary
trap in this and many other fields of the foothills area.
Thrusted anticlines appear to be the most common traps in productive foreland regions
throughout the world, but with regard to the size and specific geometry of traps, it should be
emphasized that considerable variety exists within and between these regions. Over the span of
Proterozoic time, differences in the size and morphology of plates, in their marginal sedimentary
character, inherited structural features, and specific motions have all ensured a high degree of
structural variation in foreland orogenic zones.
Like compressive basement faulting, decollement thrusting is most often related to processes
occurring along convergent plate boundaries, both in the mentioned foreland belts and along the
leading edges of subduction zones. We have already discussed some of the details of B-type
subduction zones. Figure 3 displays the interpreted structure on a seismic section across the

active subduction complex offshore of northern California.

Figure 3

This section shows the decollement faulting in the accretionary prism quite well.
Note the change from the chaotically deformed accretionary prism sediments into the more
conformable forearc basin at the far right. The deformed complex is broken into thrust slices
about 400-500 m (1300-1600 ft) thick, whose bounding faults display variable curvature. This is
probably due to a degree of stratigraphic control and ramping of fault development. We can also
see that where compressional features dominate the accretionary prism, extensional faulting
characterizes the deep sea sedimentary section west of the trench. This has been observed in a
number of active subduction zones. Often, these faults have developed in the underlying ocean
crust and have been interpreted as being the result of lithospheric bending and consequent
tensional rupture.

Collision-related thrust belts verge toward the subducting ("on-coming") plate and represent the
succession of B-subduction by A-subduction. ( Figure 4 , a map showing distribution of the major

collision-related fold and thrust belts and related foredeeps (regional foreland basins.

Figure 4

) Many of the world's foredeeps (Appalachian, Canadian Rocky Mountain, Arabian, Uralian) are

highly productive of oil and gas. Cross section lines refer to Figure 5.) A transition to basementinvolved thrusting often occurs in these settings behind the thrust front. Figure 5

Figure 5

and Figure 6 (Generalized cross section through four of the world's major collision-generated

mountain systems at comparable scales) show how this involvement is usually interpreted to
increase into the mobilized,

Figure 6

metamorphic "core"' of an orogen, such as the Canadian cordillera. In the case of trench
deformation, the situation is less well understood. Two forms of basement-involved faulting are
presumed: that which incorporates "slivers" of ocean crust (known as ophiolites) into the
accretionary prism, and that which cuts continental crust. For both trench and foreland thrust
zones, then, overlap occurs between detached and "connected" structural styles.
Most thrust belts presently exposed at the earth's surface occur as sinuous belts up to thousands
of kilometers long. Their width is not uniform, but is instead characterized by sharply recessed
and more gently extended portions known, respectively, as "reentrants" and "salients" (see Figure
4 ). The cause for this type of variation along tectonic strike is not well understood, but often
thought to be related to preexisting basement features.
Figure 5 and Figure 6 compare the overall features of several major collision-generated foreland
regions. The Alps appear to have resulted from a series of collisional episodes between the

Eurasian continent and various arclike continental fragments that once bordered it to the south.
The Zagros orogen, as we have noted (see Figure 7 , Generalized map and cross section

showing continental breakup along the Red Sea rift and collision in the Zagros region of
southeastern Iran.

Figure 7

Numbers indicate total estimated separation (in km) between Africa and Arabia), is the continuing
result of Eurasia colliding with the much smaller Arabian continent. Both the extensive Himalayan
Mountain belt and the Appalachian-Ouchita-Marathon system are interpreted as the result of the
impact between larger continental masses-in the first case, India and Eurasia, in the second,
those Paleozoic continents (proto-North America, South America, Africa, and Europe) whose
collision marked the early formation of Pangaea. In both the Alps and the Himalayas, great
thicknesses of crystalline basement rocks are involved in the deformation and have exerted
considerable control on resulting structures. Though decollement thrusting is evident in the Jura

Mountains of western Switzerland and eastern France, it does not dominate the regional
structural style of the Alps as it does the Appalachian, Zagros, or Canadian forelands.
At the same time, however, we can also think of the Alps and the Himalayas as two opposite endmembers with respect to the general style of deformation The the sedimentary cover. The central
and southern European Alps, with their spectacular development of thrust and fold nappe
structures, represent the most highly contorted foreland region in the world. Broadly speaking,
they show the effects of very rapid, collision-related diastrophism on a thick, only semicompacted
sedimentary pile. Folding of a highly ductile nature is common; recumbent and overturned nappes
are piled up on top of each other like rumpled carpets to form the higher ranges (see Figure 8 ,

Alpine nappe structures). Sediments appear to have been squeezed up and out of forearc and
backarc basins, as the various island arc fragments were progressively welded to the Eurasian
continent.

Figure 8

The Himalayas present a very different case. Here, great thicknesses of well-indurated Paleozoic
and Early Mesozoic sandstone and carbonate rocks were involved in the deformation. This
generally created less contorted, more widely spaced, and far more massive structures.
As in the Canadian Rocky Mountains and the western overthrust belt of the United States,
ramping is well developed in the Appalachian foreland. This has been interpreted as being the
result of both ductility contrast related to stratigraphic variation and preexisting block faulting in
the basement.

The well-exposed, dramatic structures of the Zagros foreland appear to resemble those of the
Appalachians more than those of the Alps or Himalayas. Some of the larger folds stretch for as
much as 160 km (100 miles) along strike before plunging beneath the surface. The existence of
salt layers in the lithologic section has resulted in a high degree of complex, local decollement. In
places, for example, slip has occurred along and within salt intervals. This has apparently created
shallow anticlines that have subsequently been pierced by the tectonically mobilized salt.
In discussing foreland thrust belts, there are a number of general aspects that are of direct
importance to hydrocarbon exploration:
1. Anticlines occur primarily in the hangingwall, and are asymmetric toward the direction
of tectonic transport.
2. The geometry of thrusting is much dependent on the competence of the sedimentary
units involved. The number of thrusts, as well as the intensity of folding, decreases in
distinct proportion to increases in the amount of thick, competent units (e.g., sandstone,
carbonates).
3. Sedimentary thicknesses decrease and deposits change character to more shallowwater clastics in the direction of tectonic transport, i.e., away from the metamorphic core,
or, in stratigraphic terms, toward the basin margin. This usually results in an increase in
the overall density of faulting in the same direction.
4. Structures become progressively younger in the direction of tectonic transport.
5. Rocks involved in thrusting also become younger in this direction.
6. Advancing thrust sheets act to load and depress the crust into local fore-land basins
(sometimes called "molasse" basins). These fill with coarse marine and nonmarine
detritus, which then also becomes involved in deformation. Such basins are often rich in
plant-derived organic matter.
7. In many cases, a regional foreland basin, relatively rich in petroleum, will exist
immediately out in front of the thrust belt, presumably created by crustal loading on a
more massive scale.
Thus, deformation appears to begin in the deeper, thicker portions of the sedimentary wedge and
progress upward and on-to the craton. In a broad sense, this has meant that hydrocarbons have

had the best chance to accumulate and remain undisturbed near the youngest, leading portions
of thrust belts and in the regional foreland basins out in front of them.
As mentioned, traps in thrust belts are mainly associated with asymmetric anticlines ( Figure 9 ,

Seismic profile through the eastern Po plain in northeastern Italy (approximately 50 km southwest
of Venice), showing thrust structure of the Apennines.

Figure 9

This profile reveals the thrust belt at its widest point. Gas pools exist in Pliocene sandstones that
wedge out against the rising structures. Oil is produced from complex structural traps in
underlying Mesozoic carbonates) and, to a lesser extent, fault truncations. Closures are most
often at less than about 3000 m depth (Harding and Lowell 1979). The actual size of individual
traps, and their height of closure can vary a great deal, depending on the spacing of thrusts and
the degree of asymmetry in folds.

Substantial-even giant-accumulations have been discovered in a number of the world's foreland


regions. Perhaps the most impressive example of production from thrust-fold structures is offered
by the oil fields in the Zagros Mountains. Here, reservoir quality is due to an extensive,
interconnected fracture system generated by the folding. Production is from the Asmari
Limestone, which, by itself, yields more than 75% of all petroleum currently being recovered from
traps in foreland belts. In addition, several recent major discoveries along the Idaho-Wyoming
thrust belt have encouraged continued exploration and have caused geologists to take another,
more detailed look at the petroleum potential of other foreland regions, such as the AppalachianOuchita system.
With regard to active subduction zones, several general statements can be made concerning
overall hydrocarbon potential. In the forearc, for example, both source and seal can be more than
adequate, but the large amount of volcanogenic material generally makes for rather poor
reservoir quality. Not only is the sediment matrix often fine-grained, but both primary and
secondary pore-plugging and swelling clays-most notably illite and montmorillonite-are abundant.
Moreover, low heat flow is characteristic of the forearc; this is even more the case in the forearc
basins that develop over the subduction complex itself. Structural traps, however, abound and
consist of anticlinal and thrust closures and relatively shallow drape folds above thrusts.
The problems mentioned for forearc regions appear to characterize the Makran subduction
complex, which, because of its unique setting, may otherwise appear to offer relatively strong
hydrocarbon potential ( Figure 10 ,

Figure 10

Interpreted structure of the Makran accretionary prism, Gulf of Oman,see Figure 4 for
approximate location) This very large accretionary prism is basically the continuation of the
Zagros collision zone into a B-type subduction zone that stretches nearly 900 km eastward from
the southwestern coast of Iran to Pakistan. It shows the development of many thick, coherent
thrust-fold structures whose amplitudes are unusually large and whose petroleum potential might
therefore seem to be substantially greater than that in most other forearc arc systems. As much
as 7 km of mostly late Tertiary abyssal plain sediments (most likely resulting from high erosion
rates that began with the India-Eurasia collisional event to the north and west) have been
deformed into an imbricated stack.
Despite this structurally attractive setting, potential appears moderate at best due to relatively low
geothermal gradients, which average 1 per 30 m (Harms et al. 1983). Gas seeps and traces of
heavy hydrocarbons, however, have been found. At the same time, wells drilled in coastal
Makran have encountered very high pressures. Because of the young age of the sediments

penetrated (Pliocene at total depth), their rapid accumulation (about 300 m per my), and their
subsequent deformation, we can expect pore pressures to be quite high. As a whole, then, the
combination of remote access, potential drilling problems, and low-to-moderate maturation
potential makes most active forearc systems relatively high exploration risks at present.

Strike Slip Tectonics: Wrench Faults


For most exploration purposes, strike slip, oblique slip, and wrench fault systems can be
considered as roughly equivalent, though their specific plate tectonic settings are variable. In all
cases, they are assumed to involve basement. These fault systems can be pure strike slip or they
may include components of compression (sometimes called ""transpression") or extension
("transtension") ( Figure 1 , Diagram illustrating the evolution of various structures associated with

major wrench faults. Arrows labeled "C" and "E" indicate compressional and extensional
components).

Figure 1

Major wrench faults are usually associated with transform plate margins. In our general
discussion of faulting, we have looked at several major examples of these structures ( Figure 2
and Figure 3 , (Examples of major strike slip faults in various parts of the world.

Figure 2

Dots mark the site of active volcanoes.) which can occur as single shear planes or systems of
parallel faults (Harding 1974,1983, 1985).

Figure 3

For petroleum exploration purposes, the geologist is most interested in the structures associated
with wrench faults, for it is these that form potential structural traps.
To a large extent, the components of compression and extension are determined by the degree
of obliqueness involved in plate convergence. As convergence swings from "head on" to lower
and lower angles of incidence, the relative amount of strike slip motion increases. Theoretically,
associated compression should decrease.
Wrench fault structures are extremely varied, as mentioned by Harding (1985), and include many
of the features seen in other styles. Figure 4 (Diagram illustrating the evolution of various

structures associated with major wrench faults.

Figure 4

Arrows labeled "C" and "E" indicate compressional and extensional components) shows how they
can be resolved according to the complex stress fields that result from such large-scale shearing.
The development of specific structures and their relative importance is naturally dependent on
how much compression or extension might be involved in plate convergence.
An example of a productive backarc wrench fault setting in Sumatra is given in Figure 5 (Map of

Sumatra showing location of the Barisan wrench fault and associated folds and oil fields.

Figure 5

Large arrows indicate relative direction of magnitude of plate convergence in the Java Trench.
Box at lower right indicates area of seismic line shown in Fig. 6.) and Figure 6 (Inverted basin
structure - a case of superposed tectonism.

Figure 6

Note the differences in offset indicated by upper reflectors versus lower reflectors. Reactivation of
basement faulting in compression is presumed to be responsible for the structural inversion
observed. The Rambutan field produces from fold closures in the hangingwall of an inverted
fault.) In this case, wrenching has created a series of large, secondary anticlines and inverted
basins. Reservoirs are in Late Tertiary sands derived from a crystalline source. Notice that for
several of the major faults shown in Figure 7 (Interpreted seismic profile across a wrench zone in

the Andaman Sea, showing a negative flower-type structure.

Figure 7

Note that seismic data is relatively chaotic immediately east of the interpreted fault. Letters "A"
and "T" refer to the blocks whose relative motion is away from or toward the viewer ), the original
sense of offset is preserved at depth, while the upper layers reveal the latest displacement.
Both folds and reverse faults are commonly associated with wrench faults characterized by
transpression. The strike of these secondary features is commonly at low angles (not parallel) to
that of the main fault plane. They also frequently occur in echelon patterns (as diagrammed in
Figure 4 ). Extensional structures are primarily normal faults. As we have seen in Figure 5 and
Figure 6 , transpression can also cause basin inversion and thus relatively shallow closures.
The term flower structure is often used to describe the upward-branching form that many wrench
faults have been interpreted to show on seismic profiles. The flower itself is denoted "positive" or
"negative" on the basis of whether the units within it are arched by reverse components of

displacement or dropped by normal separations. Figure 7 gives an interpreted example of a


negative flower.
Figure 8 (Structure contour map of Los Angeles basin, showing location of Whittier and Norwalk

faults.

Figure 8

These are compressional structures, related to the San Andreas fault system. As shown, oil fields
in this area are associated with both fold and fault closures ) and Figure 9 (Cross section through
Whittier oil field, showing complex structural relationships.

Figure 9

Whittier fault zone is highly mylonitized and acts as an updip seal. Note apparent positive flower
structural style ) together show an example of the complex but productive structures associated
with wrench faulting in the Los Angeles basin of southern California. We should note in the cross
section that the Whittier fault, with its highly sheared mylonite zone, acts as an updip seal. The
general appearance of the structure is a positive flower, with several generations of faulting
evident. This type of complexity is common along wrench fault zones, and can create a variety of
potential traps ( Figure 10 , Trapping possibilities associated with wrench faults). In addition to
anticlinal and fault-trap closures, shearing on such a scale may be expected to generate fractures
in surrounding competent lithologies (Harding 1985).

Figure 10

Despite the relatively wide use of flower structure terminology, questions remain as to its value.
Often, seismic data in the crucial core areas of supposed flowers is inconclusive and open to
alternate interpretations. Some structures that have been called positive flowers (e.g., in the
Ardmore basin of southern Oklahoma; see Harding and Lowell 1979) have also been interpreted
as thrusts. Those with apparent negative "bloom" have also been analyzed as rift structures. In
regions such as the North Sea or the Red Sea, the problem is most likely compounded by the fact
that components of both strike slip and normal displacement are responsible for the fault
geometries observed.

Vertical Tectonics: Basement Warps


The deep interior of every major craton is characterized by broad regional arches and open
circular basins that can contain more than 14 km of sedimentary fill. Figure 1 (Regional NW-SE

cross section through Gulf of Mexico) shows the distribution of the world's principal intracratonal
basins.

Figure 1

These structures have proved to be the sites where vast quantities of hydrocarbons have
accumulated in both structural and stratigraphic traps. Such traps are often far greater in
continuous extent than are those in more tectonically disturbed regions. To date, plate tectonic
theory has been unable to adequately account for the existence and behavior of such major
intracratonal warps. It would seem at this point that accurate explanation of them is contingent
upon better understanding of the lower crust and upper mantle and the variations that may
characterize the boundary between them.
A majority of the world's intracratonal basins have persisted through Phanerozoic time. Their
activity, as documented by the details of their sedimentary fill, has been intermittent and of

variable rate. Along with the regional arches that often form their margins, they are the dominant
structural style of continental interiors. Their relative concentration of deeper-water sediments
during periods of transgression, the great thicknesses of total deposits within them, and their
considerable depth into the crust all combine to make such basins nearly ideal provinces for the
generation of petroleum.
Hydrocarbons are often trapped by secondary faults and closures within these basins. ( Figure 2 ,

Various trapping possibilities associated with intracratonal basins and domes.

Figure 2

) The patterns of such smaller-scale structures are completely singular to each basin, but usually
display pronounced trends that appear to be related to large-scale tectonic patterns in the
underlying basement that have been identified from gravity and magnetic data. Generally
speaking, then, basement structure is often an essential key to exploration of the potential in
overlying sediments. Drilling, however, rarely penetrates the entire depth of sedimentary fill, and
thus direct study of basement rocks is usually limited to the shallower margins of a basin.
Faults are most often high-angle and show either normal or strike slip displacement.
Compressional structures are usually open folds that may, as in the case of the Michigan basin
be consistent in orientation. ( Figure 3 , Structural contour map showing top of Precambrian, with

four major faults. All of these have strong strike slip components of displacement.

Figure 3

Figure 4 , Regional cross section.

Figure 4

Note buried late Precambrian Keweenawan rift and Grenville front. Figure 5 , Maps showing
basement provinces and the trend of productive anticlines in the overlying sedimentary basin fill.
The overall NW-SE trend in fold strike is only suggested by the basement map.

Figure 5

The fight-hand figure shows locations of the wells productive in the deep Cambro-Ordovician as
of mid-1984.)
A large number of traps in basins of this type, however, are stratigraphic and can be very subtle.
During periods of active subsidence, downwarps act to localize certain depositional patterns.
Unconformities; pinchouts due to sedimentary onlap; and reef, evaporite, and alluvial plain
sedimentation are among the potential trap-generating characteristics that result.
Figure 3 , Figure 4 , and Figure 5 show the basic structural setting of the Michigan basin. The
cross section indicates that, far from being a simple crustal sag, the floor of this basin is ruptured
by the failed Keweenawan rift and the leading edge of the Grenville front, a probable megasuture
of Late Precambrian age. Moreover, the "glove" of Michigan in general marks the enigmatic
juncture of several principal shield provinces of North America.

Thus, it would seem that the crust in this specific location was potentially thinned and vulnerable
to subsidence. Recent seismic surveys across Hudson Bay in Canada also show distinct
basement block faulting. Though not a model for all intracratonal basins, this type of setting may
indicate that deep-seated inhomogeneities in crustal character are an essential factor in the
localization of such profound structures. Overall, simplicity of form disguises an apparent
complexity of origin.
Other proposed ideas for intracratonal downwarping include the following: thermal contraction in
the mantle; subcrustal flowage of material; mineral phase changes (causing local increases in
crustal density); subcrustal erosion; differential cooling of the lithosphere. Crustal loading due to
continued sedimentation is an obvious contributing factor, once, of course, subsidence has
begun.
For exploration, the structural configuration of a specific basin will provide a broad context for
locating potential traps within it (compare Figure 5 part (b) and ( Figure 6 , Distribution of principal

oil and gas fields in Michigan.Traps are associated with several types of features: in the north and
west, a Siluraian pannacle reef trend).

Figure 6

Rather than identifying structures only, seismic data may be of greater use in mapping certain
productive intervals and their stratigraphic relationships. In Michigan, most of the older fields (now
mostly in the stripper-well stage) were discovered in a trend of gentle anticlines in the central
portion of the basin. The only giant producing area-the AlbionScipio trend (see Fig. 3 and Fig. 6)
is associated with fracturing in Ordovician carbonates along a local strike slip fault. More recent
production, however, has come from a Silurian pinnacle reef trend that rims the basin to the north
and west. Though often prolific, these reservoirs are quite small and unrelated to structure,
except in the most general terms of stratigraphic localization along the basin margin.
As a general rule, the structures of one intracratonal basin cannot be used as a specific guide to
exploring other such provinces. Each basin is unique. For this reason, the explorationist is forced
to become more of a specialist than might otherwise be the case. As an overall approach,
petroleum geologists in Michigan have found it useful to extrapolate certain stratigraphic
relationships observed near the margins of the basin into its center. This, however, will be risky in

other provinces, given the facies changes that can occur. Thus, the range of traps shown in
Figure 7 (Various trapping possibilities associated with intracratonal basins and domes ) is only a
guide.

Figure 7

Generally, however, with only a few exceptions, such as the Anadarko basin of central Oklahoma,
most intracratonal basins have been explored only to relatively shallow levels. To some degree,
therefore, a deep frontier remains in many heavily drilled areas. Recent gas discoveries below
15,000 ft (4500 m) in the Cambro- Ordovician sandstones of the Michigan basin and below
20,000 ft (6000 m) in the Ordovician Ellenberger group of the Permian basin in western Texas
strongly support this idea.
a. Interpret the dipmeter log in Figure 1 and the structural style it reveals.

Figure 1

Sketch the cross section along an E-W line.


b. Where in the section do you find closure?
c. Briefly describe the tectonic regime(s) in which you expect to find this structural style.

References
Allen, c. R. 1962. Circum-Pacific faulting in the Philippines-Taiwan region. Jour. Geophys. Res.
67:4795-4812.
Allen, J. R. L. 1970. Physical processes of sedimentation. Earth Science Series No. 1, ed. J.
Sutton and J. V. Watson. George Allen & Unwin., Ltd.
Albright, W. A., W. L. Turner, and K. R. Williamson. 1980. Ninian Field, U.K. Sector, North Sea. In
Giant oil and gas fields of the decade 1968-1978, ed. M. Halbouty. AAPG Memoir no. 30:173195.

Arthur, K. R., D. R. Cole, G. L. Henderson, and D. W. Kushnir. 1983. Geology of the Hibernia
discovery. AAPG Memoir no. 32:181-197.
Atwater, T. 1970. Implications of plate tectonics for the Cenozoic tectonic evolution of western
North America. Geol. Soc. Amer. Bull. 81:3513-3536.
Badgley, P. C. 1962. The analysis of structural patterns in bedrock. SME/AIME Transcript 225.
381-399.
_______ 1965. Structural and tectonic principles. New York: Harper and Row.
Badley, M. E. 1985. Practical seismic interpretation. Boston: IHRDC.
Bally, A. W. 1975. A geodynamic scenario for hydrocarbon occurrences. Proc., 9th World Petrol.
Congr., Tokyo, v. 2 (Geology), Essex, England: Applied Sci. Pub., Ltd. 33-44.
Bally, A. W., ed. 1983. Seismic expression of structural styles: A picture and work atlas: AAPG
Studies in Geology Series, no. 15, 3 vols.
Bally, A. W., P. L. Gordy, and G. A. Steward. 1966. Structure, seismic data, and orogenic
evolution of southern Canadian Rocky Mountains. Can. Petrol. Geol. Bull. 14:337-381.
Bally, A. W., and J. S. Oldow. 1983. Plate tectonics, structural styles, and the evolution of
sedimentary basins. Houston Geological Society, short course manual.
Bally, A. W., A. B. Watts, J. A. Grow, W. Manspeizer, D. Bernoulli, c. Schreiber, and J. M. Hunt.
1981. Geology of passive continental margins. AAPG Education Course Note Series, no. 19.
Biddle, K. T., and D. R. Seely. 1983. Structure of a subduction complex. In Seismic expression of
structural styles: A picture and work atlas, ed. A. W. Bally. AAPG Studies in Geology Series, no.
15. 3:3.4.2-129-135.
Billings, M. P. 1972. Structural geology, 3d ed. Englewood Cliffs, NJ: Prentice-Hall.
Bishop, R. S. 1978. Mechanism for emplacement of piercement diapirs. AAPG Bull 62:15611583.
Blatt, H., G. Middleton, and R. Murray. 1972. Origin of sedimentary rocks. Englewood Cliffs, NJ:
Prentice-Hall.

Bloom, A. L. 1978. Geomorphology. Englewood Cliffs, NJ: Prentice-Hall.


Brewer, J. A., S. B. Smithson, J. E. Oliver, S. Kaufman, and L. D. Brown. 1980. The Laramide
orogeny: Evidence from COCORP deep crustal profiles in the Wind River Mountains, Wyoming.
Tectonophysics. 62:165-189.
Brown, A. R. 1985. The role of horizontal seismic sections in stratigraphic interpretation. AAPG
Memoir no. 39:37-47.
Brown, L. F., Jr. 1979. Deltaic sandstone facies of the mid-continent. In Pennsylvanian
sandstones of the mid-continent. Tulsa Geol. Soc. Special Pub., no. 1:35-63.
Brown, L. F., Jr., and W. L. Fisher. 1977. Seismic-stratigraphic interpretation of depositional
systems: Examples from Brazilian rift and pull-apart basins. In Seismic stratigraphy-applications
to hydrocarbon exploration, ed. C. E. Payton. AAPG Mem. 26. 213-248.
_______ 1980. Geology and geometry of depositional systems. In Seismic stratigraphic
interpretation and petroleum exploration. AAPG Education Course Notes Series, no. 16:56.
Brown, W. G. 1982. A short course in structural geology. AAPG southwest sectional meeting,
Wichita Falls, Texas.
Bruce, C. H. 1973. Pressure shale and related sediment deformation- mechanisms for
development of regional contemporaneous faults. AAPG Bull. 57:878-886.
_______ 1983. Shale tectonics, Texas coastal area growth faults. In Seismic expressions of
structure and styles. AAPG Studies in Geology Series, No. 15, 3 vols. 2.3.1-1-2.3.1-6.
Bullard, T. 1973. Eugene Island block 126 field, offshore St. Mary Parish, Louisiana: Offshore
Louisiana oil and gas fields. Lafayette Geol. Soc. p.111
Burchfield, B. C., and G. A. Davis. 1975. Nature and controls of cordilleran orogenesis, western
United States; extensions of an earlier synthesis. Am. j Sci. 275-A:363-396.
Casella, C. J. 1964. Geologic evolution of the Bear-tooth Mountains, Montana and Wyoming, Part
4: Relationship between Precambrian and Laramide structures in the Line Creek area. Geol. Soc.
Amer. Bull. 75:969-986.

Chapin, C. E., and S. M. Cather. 1983. Eocene tectonics and sedimentation-Colorado Plateau,
Rocky Mountain Area. In Rocky Mountain foreland basins and uplifts, ed. J. D. Lowell, 33-57.
Rocky Mt. Assn. of Geologists.
Christensen, C. F. 1983. An example of a major syndepositional listric fault. In Seismic
expression of structural styles: A picture and work atlas, ed. A. W. Bally. AAPG Studies in
Geology Series, no. 15. 2:2.3.1-36-40.
Cloos, E. 1947. Oolite deformation in the South Mountain Fold, Maryland. Geol. Soc. Amer. Bull.
58:843-918.
Cloos, H. 1939. Conversations with the earth. Munich: Piper & Co.
_______ 1948. Gang und gehwerk einer falte. Zeitschrift Deutsche Geol. Geshichte 100:290-303.
Cook, F. 1982. The COCORP seismic reflection traverse across the southern Appalachians.
AAPG Studies in Geology Series, no. 14. 61 p.
Cook, F., et al. 1979. Thin-skinned tectonics in the crystal line southern Appalachians; COCORP
seismic reflection profiling of the Blue Ridge and Piedmont: Geology 7:563-567.
Currelle, R., and R. Marco. 1983. Reflection profiles across the Aquitaine Basin (salt tectonics). In
Seismic expression of structural styles: A picture and work atlas, ed. A. W. Bally. AAPG Studies
in Geology Series, no. 15. 2:2.3.2-11-18.
Dahlstrom, C. D. A. 1970. Structural geology in the eastern margin of the Canadian Rocky
Mountains. Can. Petrol. Geol. Bull. 18:332-406.
Dailly, G. C. 1976. A possible mechanism relating pro-gradation, growth faulting, clay diapirism,
and over-thrusting in a regressive sequence of sediments. Can. Petrol Geol. Bull. 24:92-116.
Davis, G. H. 1984. Structural geology of rocks and regions. New York: John Wiley & Sons.
De Jong, K. A., and R. Scholten, eds. 1973. Gravity and tectonics. New York: John Wiley & Sons.
Dennis, J. G. 1972. Structural geology. New York: John Wiley & Sons.
De Sitter, L. U. 1964. Structural geology. New York: McGraw-Hill.

Dickinson, W. R. 1974. Plate tectonics and sedimentation. Soc. Econ. Paleon. and Mineral. Spec.
Pub. no. 22:1-27.
Dieterich, J. H. 1969. Origin of cleavage in folded rocks. Am. Jour. Sci 267:155-165.
Dillon, W. P., C. K. Paull, R. T. Buffler, and J. P. Fail. 1979. Structure and development of the
southeast Georgia embayment and northern Blake plateau: Preliminary analysis. In Geological
and Geophysical Investigations of Continental Margins, ed. J. S. Watkins, L. Montadert, and P.
W. Dickerson. AAPG Memoir no. 29:27-43.
Douglas, R. J. W. 1950. Callum Creek, Langford Creek, and Gap map-areas, Alberta. Geol
Survey of Canada. Memoir no. 255.
Edwards, D. 1984. Mechanical properties evaluation and its applications. Report prepared for
Schlumberger Technical Services, Inc., European Unit, 62 p.
Elliot, D. 1976. The motion of thrust sheets. Jour. Geophys Res. 81:949-963.
_______ 1977. Some aspects of the geometry and mechanics of thrust belts. Seminar notes, 8th
Ann. Mtg. Can. Soc. Petrol Geol Parts I and II.
Elter, P., and L. Trevisan. 1973. Olistostromes in the tectonic evolution of the Northern
Apennines. In Gravity and tectonics, ed. K. A. DeJouge and R. Scholten. New York: John Wiley &
Sons.
Erxleben, A. W., and G. Carnahan. 1983. Slick Ranch Area, Starr County, Texas. In Seismic
expression of structural styles: A picture and work atlas, ed. A. W. Bally. AAPG Studies in
Geology Series, no. 15. 2:2.3.1-22-27.
Fischer, A. G. 1969. Geologic time-distance rates: The Bubnoff unit. Geol. Soc. Amer. Bull.
80:549-551.
Flint, R. F. 1 971. Glacial and qua ternary geology. New York: John Wiley & Sons.
Fox, F. G. 1959. Structure and accumulation of hydrocarbons in southern foothills, Alberta,
Canada. AAPG Bull. 43:992-1023.
Gansser, A. 1964. Geology of the Himalayas. New York: Interscience.

Gibson, I. L., and G. P. L. Walker. 1964. Some composite rhyolite-basalt lavas and related
composite dykes in eastern Iceland. Proc. Geol. Assoc. London 74:301-318.
Gilreath, J. A. and R. W. Stephins. 1975. Interpretation of log responses in a deltaic environment.
Paper presented at AAPG Marine Geology Workshop, Dallas.
Goguel, J. 1962. Tectonics. San Francisco: W. H. Freeman & Co.
Gries, R. 1983. North-south compression of Rocky Mountain foreland structures. In Rocky
Mountain foreland basins and uplifts, ed. J. D. Lowell. Rocky Mt. Assn. of Geologists 9-33.
Griggs, D. T., and J. Handin. 1960. Observations on fracture and a hypothesis of earthquakes. In
Rock deformation-A symposium, ed. D. T. Griggs and J. Handin. Geol. Soc. Amer. Memoir no.
79:347-373.
Grow, J. A., R. E. Mattick, and J. S. Schlee. 1979. Multichannel seismic depth sections and
interval velocities over outer continental shelf and upper continental slope between Cape
Hatteras and Cape Cod. In Geological and geophysical investigations of continental margins, ed.
J. S. Watkins, L. Montadert, and P. W. Dickerson. AAPG Memoir no. 28:65-85.
Grow, J. A., D. R. Hutchinson, K. D. Klitgord, W. P. Dillon, and J. S. Schlee. 1983. Representative
multichannel seismic profiles over the U.S. Atlantic Margin. In Seismic expression of structural
styles: A picture and work atlas, ed. A. W. Bally. AAPG Studies in Geology Series, no. 15.
2:2.2.3-1-20.
Halbouty, M. T. 1979. Salt domes, Gulf region, United States and Mexico, 2nd ed. Houston: Gulf
Publishing Co.
Handin, J. W., R. V. Hager, M. Friedman, and J. M. Feather. 1963. Experimental deformation of
sedimentary rocks under confining pressure: Pore pressure tests. AAPG Bull. 47:717-755.
Harding, T. P. 1974. Petroleum traps associated with wrench faults: AAPG Bull. 58:1290-1304.
_______ 1976. Tectonic significance and hydrocarbon trapping consequences of sequential
folding synchronous with San Andreas faulting, San Joaquin Valley, California. AAPG Bull.
60:356-378.

_______ 1983a, Divergent wrench fault and negative flower structure, Andaman Sea. In Seismic
expression of structure styles: A picture and work atlas, ed. A. W. Bally. AAPG Studies in
Geology Series, no. 15. 3:4.2-1-9.
_______ 1983b. Structural inversion at Rambutan oil field, South Sumatra basin. In Seismic
expression of structural styles, ed. A. W. Bally. AAPG Studies in Geology Series, no. 15, v. 3.
_______ 1984. Graben hydrocarbon occurrences and structural style. AAPG Bull. 68:333-362.
_______ 1985. Seismic characteristics and identification of negative flower structures, positive
flower structures, and positive structural inversion. AAPG Bull. 69:582-601.
Harding, T. P., and J. D. Lowell. 1979. Structural styles, their plate-tectonic habitats and
hydrocarbon traps in petroleum provinces. AAPG Pull 63:1016-1069.
Harms, J. C., H. N. Copel, D. C. Francis, and T. J. Shakelford. 1983. Summary of the geology of
the Makran Coast. In Seismic expression of structural styles: A picture and work atlas, ed. A. W.
Bally. AAPG Studies in Geology Series, no. 15. 3:3.4.2-173-178.
Hatcher, R. D., Jr. 1972. Developmental model of the southern Appalachians. Geo. Soc. Amer.
Bull. 83(9): 2735ff.
Heard, H. C. 1960. Transition from brittle to ductile flow in Solenhofen limestone as a function of
temperature, confining pressure, and interstitial fluid pressure: In Rock deformation-a symposium,
ed. D. T. Griggs and J. Handin. Geol. Soc. Amer. Memoir no. 79:193-226.
Heard, H. C., and C. B. Raleigh. 1972. Steady-state flow in marble at 500 C to 8000 C. Geol.
Soc. Amer. Bull. 83:935-956.
Hills, E. S. 1971. Elements of structural geology, 2nd ed. London: Methuen and Co., Ltd.
Hobbs, B. E. 1971. The analysis of strain in folded layers. Tectonophysics. 11:329-375.
Hobbs, B. E., W. D. Means, and P. F. Williams. 1976. An outline of structural geology. New York:
John Wiley & Sons.
Hoeppener, R. 1964. Zur Physikalischen Tektonik: Felsmechanik und Ingenieurgeologie. 2:22-44.

Holland, D. S., C. F. Sutley, R. F. Berlitz, and J. A. Gilreath. 1976. East Cameron Block 270, a
Pleistocene field. In North American oil and gas fields, ed. J. Braunstein. AAPG Memoir no.
24:205-228.
Holmes, A. 1978. Principles of physical geology, 3rd ed. New York: John Wiley & Sons.
Inoue, E. 1960. Land deformation in Japan. Geograph. Surv. Inst. (Japan) Bull. 6:73-134.
Jackson, J., and D. P. McKenzie. 1983. The geometric evolution of normal fault systems. Journal
of Structural Geology 5:471-482.
Jackson, M. P., and W. E. Galloway. 1984. Structural and depositional styles of Gulf Coast
tertiary continental margins: Application to hydrocarbon exploration. AAPG Continuing Education
Course Notes Series, no. 25.
Jamison, H. C., L. D. Brockett, and R. A. McIntosh. 1980. Prudhoe Bay-a ten-year perspective.
AAPG Memoir no. 30:289-314.
Jones, H. P., and R. G. Speers. 1976. Permo-Triassic Reservoirs of Prudhoe Bay Field, North
Slope, Alaska. In North American oil and gas fields, ed. J. Braunstein. AAPG Memoir no. 24:2351.
Karig, D. E. 1974. Evolution of arc systems in the western Pacific. Ann. Rev. Earth Planet. Sci.
2:51-75.
_______ 1977. Growth patterns on the upper trench slope. Amer. Geophys. Union Monograph, M.
Ewing Series. 1:175-186.
Karig, D. F., M. B. Lawrence, G. F. Moore, and J. R. Carvey. 1980. Structural framework of the
forearc basis, Northwest Sunatra. j Geol. Soc. London. 137:77-91.
Karig D. F., and G. F. Sharman. 1975. Subduction and accretion in trenches. Geol. Soc. Amer.
Bull. 86:377-389.
Karig, D. F., S. Suparka, G. F. Moore, and P. E. Hehanussa. 1979. Structure and evolution of the
Sunda Arc. In Geological and geophysical investigations of continental margins, ed. J. S. Watkins
et al. AAPG Memoir no. 29:223-239.
Kennedy, W. G. 1946. The Great Glen fault. Quart. j Geol. Soc., London. 102:41-72.

Kleinschmiede, W. F. J. 1960. Geology of the Valle de Aran (Central Pyrenees). Leidse Geol.
Meded. 25:129-245.
Kuenen, P. H. 1953. Significant features of graded bedding. AAPG Bull., Vol. 37,1044-1066.
_______ 1965. Value of experiments in geology. Geol. en Mijnenbouw 44:22-36.
Krumbein, W. C. 1942. Criteria for subsurface recognition of unconformities. AAPG Bull. 26:3662.
Lamerson, P. R. 1982. The fossil basin area and its relationship to the Absaroka thrust fault
system. In Geologic studies of the Cordilleran thrust belt, ed. R. B. Powers. Rocky Mt. Assn. of
Geologists. 279-340.
Laubscher, H. P. 1961. Die Fernschubhypothese der Jurafaltung. Eclog. Geol. Helvetica 54:221282.
Lehner, P., H. Doust, G. Bakker, P. Allenbach, and J. Gueneau. 1983. Active margins-Caribbean
margin of South America, Profiles C-1422, C-1412, and C-1 413. In Seismic expression of
structural styles: A picture and work atlas, ed. A. W. Bally. AAPG Studies in Geology Series, no.
15. 3:3.4.2-111-129.
LePichon, X., et al. 1982. Active tectonics in the Hellenic Trench. In Trench and forearc
sedimentation and tectonics, ed. J. K. Leggett. Geol. Soc. London Sp. Pub. 10:319-334.
Levorsen, A. I. 1967. Geology of petroleum. 2nd ed. San Francisco: W. H.. Freeman.
Ligget, M. A., and H. F. Ehrenspeck. 1974. Pahranagat shear system, Lincoln County, Nevada.
Argus Explor. Co., Rept. of Inv., NASA CR 1356388, E74-1 0206.
Lowell, J. D. 1985. Structural styles and petroleum exploration. Tulsa, OK: OGCI Publications.
Lowell, J. D., G. J. Genik, T. H. Nelson, and P. M. Tucker. 1975. Petroleum and plate tectonics of
the southern Red Sea. In Petroleum and global tectonics, ed. A. G. Fisher and S. Judson.
Princeton: Princeton Univ. Press. 129-153.
Mackin, J. H. 1950. The down-structure method of viewing geologic maps. j Geology. 58:55-72.

Maher, C. E. 1980. Piper oil field. In Giant oil and gas fields of the decade 1968-1978, ed. M.
Halbouty. AAPG Memoir no. 30:131-173.
Manspeizer, W. 1981. Early Mesozoic basins of the central Atlantic passive margins. In Geology
of passive continental margins AAPG Education Course Note Series, no. 19. 4-1-4-60.
Martin, R. G., and J. E. Case. 1975. Geophysical studies in the Gulf of Mexico. In The ocean
basins and margins, Vol. 3, ed. A. E. Nairn and F. G. Stelhi. New York: Plenum Press.
Maxwell, J. C. 1962. Origin of slaty and fracture cleavage in the Delaware Water Gap area, New
Jersey and Pennsylvania. In Petrologic studies, Buddington volume. Geol. Soc. Amer. 281-311.
Mescherikov, Y. A. 1968. Neotectonics. In Encyclopedia of geomorphology, ed. R. W. Fairbridge.
New York: Van Nostrand. 768-773.
Mitchum, R. M., P. R. Vail, and S. Thompson. 1977. The depositional sequence as a basic unit
for stratigraphic analysis. AAPG Memoir 26:53-62.
Montgomery, S. L. 1984. Michigan basin: Expanding the deep frontier. Petroleum Frontiers 1
(Spring). Denver: Petroleum Information Corp.
Morgan, L., and W. Dowdall. 1983. The Atlantic continental margin: In Seismic expression of
structural styles.' A picture and work atlas, ed. A. W. Bally. AAPG Studies in Geology Series, no.
15. 2:2.2.3-30-36.
Murray, G. H. 1968. Quantitative fracture study-Spanish Pool, McKenzie County, North Dakota.
AAPG Bull. 52:57-65.
Nadai, A. 1950. Theory of flow and fracture of solids. New York: McGraw-Hill.
Nenyon, M. K., and A. A. Fitch. 1985. Seismic reflection interpretation. Geoexploration
Monographs 1, no. 8. Berlin-Stuttgart: Gebruder Borntraeger. 312.
Nettleton, L. L. 1934. Fluid mechanics of salt domes. MPG Bull. 18:1175-1204.
Nickelsen, R. P., and V. N. Hough. 1967. Jointing in the Appalachian plateau of Pennsylvania.
Geol. Soc. Amer. Bull. 78:609-630.
Park, R. G. 1983. Foundations of structural geology. London: Chapman & Hall.

Petersen, F. A. 1983. Foreland detachment structures. In Rocky Mountain foreland basins and
uplifts, ed. J. A. Lowell. Rocky Mt. Assn. of Geologists. 65-79.
Pettijohn, F. J., P. E. Potter, and R. Siever. 1972. Sand and sandstone. New York: SpringerVerlag.
Pew, E. 1983. Seismic structural analysis of deformation in the southern Mexican ridges. Master's
Thesis, University of Texas, Austin.
Pierce, W. G. 1957. Heart Mountain and South Fork detachment thrusts of Wyoming. MPG Bull.
41:591-626.
Pieri, M. 1983. Three seismic profiles through the Po Plain. In Seismic expression of structural
styles.' A picture and work atlas, ed. A. W. Bally. AAPG Studies in Geology Series, no. 15.
3:3.4.1-8-26.
Plafker, G. 1965. Tectonic deformation associated with the 1964 Alaskan earthquake. Science.
148:1675-1687.
Powell, J. B. 1968. Exploration history of Delhi Field, northeastern Louisiana. In Stratigraphic oil
and gas fields, ed. R. E. King. AAPG Memoir no. 16:548-559.
Price, R. A., and E. W. Mountjoy. 1970. Geologic structure of the Canadian Rocky Mountains
between Bow and Athabasca Rivers-a progress report. Geol. Assoc. Canada Spec. Paper 6, 25
p.
Quiring, H. 1939. "Uber Contravergente Transformation von Faltenzonen im Rheinischen
Gebirge. Zeitscher. Deutsche. Geol. Gesellsch. Bd. 91, Hft. 6. 421-432.
Ragan, D. M. 1984. Structural geology.' An introduction to geometrical techniques, 3rd ed. New
York: John Wiley & Sons.
Ramsay, J. G. 1962. The geometry and mechanics of formation of similar type folds. Jour. Geol.
70:309-327.
_______ 1967. Folding and fracturing of rocks. New York: McGraw-Hill.
Ramsay, J. G., and M. I. Huber. 1985. The techniques of modern structural geology. London:
Academic Press.

Reading, H. G. 1980. Characteristics and recognition of strike-slip fault systems. In Sedimentation


in oblique-slip mobile zones, ed. P. F. Ballance and H. G. Reading. Int. Assn. of Sedimentologists
Spec. Pub. 4:7-26.
Reineck, H. E., and I. B. Singh. 1975. Depositional sedimentary environments. New York:
Springer-Verlag.
Richter, C. F. 1952. Elementary seismology. New York: W. H. Freeman.
Roberts, W. H. 1982. Gulf Coast Magic. Trans. Gulf Coast Assoc. Geol. Societies. 32:205-214.
Robson, D. A. 1971. The structure of the Gulf of Suez rift, with special reference to the eastern
side. Jour. Geol. Soc. 127:247-276.
Roeder, D. H., and 0. E. Gilbert. 1977. Structure, kinematics, and hydrocarbon prospects of thrust
and fold belts. AAPG Structural Geology School Course Notes.
Rodgers, J. 1953. The folds and faults of the Appalachian Valley and Ridge province. Kentucky
Geol. Survey Spec. Pub. 1:150-166.
Roux, W. F., Jr. 1977. The development of growth fault structures. AAPG Structural Geology
School Course Notes, 33 p.
Royse, F., Jr., M. A. Warner, and D. L. Reese. 1975. Thrust belt structural geometry and related
stratigraphic problems, Wyoming-Idaho-Northern Utah. In Deep drilling frontiers in the Central
Rocky Mountains, ed. D. W. Boylyard. Rocky Mountain Association of Geologists Symposium:
41-55.
Rutten, M. G. 1969. The geology of western Europe. Amsterdam: Elsevier.
Sacrison, W. R. 1978. Seismic interpretation of basement block faults and associated
deformation. In Laramide folding associated with basement block faulting in the western United
States, ed. V. Mathews. Geol. Soc. Amer. Memoir no. 151:39-49.
Sanger, J. B., and J. M. Widmier. 1977. Seismic stratigraphy and global changes of sea level,
Part 9: Seismic interpretation of clastic depositional facies. AAPG Memoir no. 26:165-184.
Sbar, M. L., and L. R. Sykes. 1973. Contemporary compressive stress and seismicity in eastern
North America: An example of intra-plate tectonics. Geol. Soc. Amer. Bull. 84:1861-1881.

Seely, D. R., and W. R. Dickinson. 1977. Structure and stratigraphy of forearc regions. AAPG
Education Course Notes, no. 5.
Seni, S. J., and M. P. Jackson. 1984. Evolution of salt structures, East Texas diapir province,
Parts 1 and 2. AAPG Bull. 67(8): 1245-1274.
Spencer, E. 1977. Introduction to the structure of the earth, 2nd ed. New York: McGraw-Hill.
Spicher, A. 1980. Tectonics profile through Switzerland. In Geology of Switzerland, ed. R.
Trumpy. Swiss Geological Comm.
Spillers, J. P. 1965. Distribution of hydrocarbons in South Louisiana by type of traps and trendsfrio and younger sediments. Trans. Gulf Coast Assn. Geol. Societies. 15:37-39.
Sprague, E. L. 1983. Geology of the Tepee flats-Bullfrog fields, Natrona County, Wyoming. In
Rocky Mountain foreland basins and uplifts, ed. J. D. Lowell. Rocky Mountain Association of
Geologists. 339-345.
Stearns, D. W. 1967. Certain aspects of fracture in naturally deformed rocks, ed. R. E. Piercker.
In NSF advanced science seminar in rock mechanics. Bedford, MA: Air Force Cambridge
Research Lab Spec. Rept. 97-118.
_______ 1971. Mechanisms of drape folding in the Wyoming Province. Wyoming Geol. Assoc.
23rd Ann. Field Conf., Wyoming Tectonics Symp., Guidebook. 125-143.
_______ 1977. Reservoirs in fractured rock. In Fracture-controlled production. AAPG Reprint
Series, no. 21:174-199.
Stearns, D. W., and M. Friedman. 1972. Reservoirs in fractured rock. AAPG Memoir no. 16:82106.
Stone, D. S. 1983a. Seismic profile-South Elk basin. In Seismic expression of structural styles.' A
picture and work atlas, ed. A. W. Bally. AAPG Studies in Geology Series, no. 15. 3:3.4.1-8-26.
_______ 1983b. The Greybull sandstone pool (lower Cretaceous) on the Elk basin thrust-fold
complex, Wyoming and Montana. In Rocky Mountain foreland basins and uplifts, ed. J. D. Lowell.
Rocky Mt. Assn. of Geologists. 345-357.

Stude, G. R. 1978. Depositional environments of the Gulf of Mexico, South Timbealer Block 54:
Salt dome and salt-dome growth models. Trans. Gulf Coast Assn. of Geol. Societies. 28:627-624.
Sunwall, M. T., K. A. McQuillan, and C. J. Nick. 1983. Salt diapir-Gulf of Mexico. In Seismic
expression of structural styles.' A picture and work atlas, ed. A. W. Bally. AAPG Studies in
Geology Series, no. 15. 2:2.3.2-32-38.
Suppe, J. 1985. Structural geology. Englewood, NJ: Prentice-Hall.
Tixier, M. P., G. Loveless, and R. Anderson. 1973. Estimation of formation strength from the
mechanical properties log. Paper no. 4532, preprint prepared for Soc. Petrol. Eng. 48th Ann. Fall
Meeting, September 30.
Trusheim, F. 1960. Mechanism of salt migration in northern Germany. AAPG Bull. 44:1519-1541.
Turcotte, D., and G. Schubert. 1982. Geodynamics. New York: John Wiley & Sons.
Vail, P. K., R. M. Mitchum. R. G. Todd, J. M. Widmier, S. Thompson, J. N. Bubb, and W. G.
Hatfield. 1977. Seismic stratigraphy and global changes in sea level, Parts 1 and 2. AAPG
Memoir no. 26:49-213.
Vine, F. J. 1966. Spreading of the ocean floor; new evidence. Science 154:1405-1415.
Wallace, R. E. 1970. Earthquake recurrence intervals on the San Andreas fault. Geol. Soc. Amer.
Bull. 81:2875-2890.
Wellborn, R. E. 1977. Structural style in relation to oil and gas exploration in North Park-Middle
Park basin, Colorado.. In Exploration frontiers of the central and southern Rockies, ed. H. K. Veal,
Rocky Mt. Assn. of Geologists. 41-60.
Whitten, E. H. T. 1966. Structural geology of folded rocks. Skokie, IL: Rand McNally.
Wilson, G. W. 1981. Introduction to small-scale structures. London: George Allen & Unwin, Ltd.
Wilson, J. T. 1965. A new class of faults and their bearing on continental drift. Nature
207(4995):343-347.
Wood, D. S. 1974. Current views of the development of slaty cleavage. Ann. Rev. Earth Sci. no.
2:1-35.

Woodbury, H. 0., I. B. Murray, Jr., and R. E. Osborne. 1980. Diapirs and their relation to
hydrocarbon accumulation. In Facts and Principles of World Petroleum Occurrence, ed. A. D.
Miall. Canadian Soc. Petrol. Geol. Memoir no.6:119-143.
Zoback, M., D. Moos, L. Mastin, and R. Anderson. 1985. Well bore breakouts and in situ stress.
Jour. Geophys. Res. 90:5523-5530.

Vous aimerez peut-être aussi