Sei sulla pagina 1di 14

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 90º.
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 90º, 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.

Potrebbero piacerti anche