You are on page 1of 27

Asst Prof. Dr.

Mayssaa Al-Bidry

Chapter three
Fracture

A fracture is any planar or sub-planar discontinuity that is very narrow in one dimension compared
to the other two and forms as a result of external (tectonic) or internal (thermal or residual) stress.
Fractures are discontinuities in displacement and mechanical properties where rocks or minerals
are broken, and reduction or loss of cohesion characterizes most fractures. They are often described
as surfaces, but at some scale there is always a thickness involved. Fractures can be separated into
shear fractures (slip surfaces) and opening or extension.

FIGURE 1 Types of brittle deformation. (a) Orientation of the remote principal stress directions with respect
to an intact rock body. (b) A tensile crack, forming parallel to σ1 and perpendicular to σ3 (which may be
tensile). (c) A shear fracture, forming at an angle of about 30° to the σ1 direction. (d) A tensile crack that
has been reoriented with respect to the remote stresses and becomes a fault by undergoing frictional sliding.
(e) A tensile crack which has been reactivated as a cataclastic shear zone. (f) A shear fracture that has evolved
into a fault. (g) A shear fracture that has evolved into a cataclastic shear zone

A shear fracture or slip surface is a fracture along which the relative movement is parallel to the
fracture. The term shear fracture is used for fractures with small (mm- to dm-scale) displacements,
while the term fault is more commonly restricted to discontinuities with larger offset. The term
slip surface is used for fractures with fracture-parallel movements regardless of the amount of
displacement and is consistent with the traditional use of the term fault.

1
Asst Prof. Dr. Mayssaa Al-Bidry

Extension fractures (tensile fractures) are fractures that show extension perpendicular to the
walls. Joints have little or no macroscopically detectable displacement, but close examination
reveals that most joints have a minute extensional displacement across the joint surfaces, and
therefore they are classified as true extension fractures.
Extension fractures are filled with gas, fluids, magma or minerals. When filled with air or fluid we
use the term fissure. Mineral-filled extension fractures are called veins, while magma-filled
fractures are classified as dikes. Joints, veins and fissures are all referred to as extension fractures.
Contractional planar features (anti cracks) have contractional displacements across them and are
filled with immobile residue from the host rock. Stylolites are compactional structures characterized
by very irregular, rather than planar, surfaces. Some geologists now regard stylolites as contraction
fractures or closing fractures, as they nicely define one of three end-members in a complete
kinematic fracture framework together with shear and extension fractures. Such structures are
known as anti-cracks in the engineering-oriented literature.
Rock mechanics experiments carried out at various differential stresses and confining pressures set
a convenient stage for studying aspects of fracture formation.

Figure 2 Experimental deformation structures that develop under extension and contraction. Initial
elastic deformation is seen for all cases, while ductility increases with temperature (T) and confining
pressure (Pc). YP, yield point.

2
Asst Prof. Dr. Mayssaa Al-Bidry

Joints
Tensile fractures
Joints (also termed extensional fractures) are planes of separation on which no shear displacement
has taken place. The two walls of the resulting opening typically remain in tight (matching) contact.
Joints may result from regional tectonics (i.e. the compressive stresses in front of a mountain belt),
folding (due to curvature of bedding), faulting, or internal stress release during uplift or cooling.
They often form under high fluid pressure ( low effective stress), perpendicular to the smallest
principal stress.

Tensile cracking
Stress Concentration
A crack in rock on the atomic scale. One way to create such a crack would be for all the chemical
bonds across the crack surface to break at once. In this case, the tensile stress necessary for this to
occur is equal to the strength of each chemical bond multiplied by all the bonds that had once
crossed the area of the crack.

FIGURE Stress concentration adjacent to a hole in an elastic sheet. If the sheet is subjected to a remote
tensile stress at its ends (σr), then stress magnitudes at the sides of the holes are equal to Cσr, where C, the
stress concentration factor, is (2b/a) + 1. (a) For a circular hole, C = 3. (b) For an elliptical hole, C > 3

3
Asst Prof. Dr. Mayssaa Al-Bidry

If you know the strength of a single chemical bond, then you can calculate the stress necessary to
break all the bonds simply by multiplying the bond strength by the number of bonds. Using realistic
values for the elasticity (Young’s modulus, E) and small strain (<10%), Equation 5.3 results in a
theoretical strength of rock that is thousands of megapascals. Measurement of rock strength in the
Earth’s crust shows that tensile cracking occurs at crack-normal tensile stresses of less than about
10 MPa, when the confining pressure is low,1 a value that is hundreds of times less than the
theoretical strength of rock.

σ=E⋅e Eq. 5.3


where E is a constant of proportionality called Young’s modulus The unit of this elastic constant is
Pascal (Pa = kg/m ⋅ s–2), Typical values of E for crustal rocks are on the order of –10 11 Pa.

We can also write elastic behavior in terms of the shear stress, σs:
σs = G ⋅ γ Eq. 5.4
where G is another constant of proportionality, called the shear modulus or the rigidity, and γ is the
shear strain. The corresponding constant of proportionality in volume change (dilation) is called the
bulk modulus, K:
σ = K ⋅ [(V – Vo)/Vo] Eq. 5.5
Perhaps more intuitive than the bulk modulus is its inverse, 1/K, which is the compressibility of a
material. Representative values for bulk and shear moduli are listed in Table 5.2.

It is quite common to use an alternative to the bulk modulus that expresses the relationship between
volume change and stress, called Poisson’s ratio, represented by the symbol ν. This elastic constant
4
Asst Prof. Dr. Mayssaa Al-Bidry

is defined as the ratio of the elongation perpendicular to the compressive stress and the elongation
parallel to the compressive stress:
ν = eperpendicular/eparallel Eq. 5.6

Poisson’s ratio describes the ability of a material to shorten parallel to the compression direction
without corresponding thickening in a perpendicular direction. Therefore, the ratio ranges from 0
to 0.5, for fully compressible to fully incompressible materials, respectively.
Incompressible materials maintain constant volume irrespective of the stress. A sponge has a very
low Poisson’s ratio, while a metal cylinder has a relatively high value. A low Poisson’s ratio also
implies that a lot of potential energy is stored when a material is compressed; indeed, if we remove
the stress from a sponge it will jump right back to its original shape. Values for Poisson’s ratio in
natural rocks typically lie in the range 0.25–0.35 (Table 5.3).

5
Asst Prof. Dr. Mayssaa Al-Bidry

A central characteristic of elastic behavior is its reversibility: once you remove the stress, the
material returns to its original shape. Reversibility implies that the energy introduced remains
available for returning the system to its original state. This energy, which is a form of potential
energy, is called the internal strain energy. Because the material is undistorted after the stress is
removed, we therefore say that strain is recoverable. Thus, elastic behavior is characterized by
recoverable strain. A second characteristic of elastic behavior is the instantaneous response to
stress: finite strain is achieved immediately. Releasing the stress results in an instantaneous return
to a state of no strain

Tensile Crack Development


Stretch a rock cylinder along its axis under a relatively low confining pressure (Figure 6.8a), a
process called axial stretching. As soon as the remote tensile stress is applied, preexisting
microcracks in the sample open slightly, and the remote stress is magnified to create larger local
stresses at the crack tips. Eventually, the stress at the tip of a crack exceeds the strength of the rock
and the crack begins to grow. If the remote tensile stress stays the same after the crack begins to
propagate, then the crack continues to grow, and may eventually reach the sample’s margins. When
this happens, the sample fails, meaning it separates into two pieces that are no longer connected
(Figure 6.8b).

6
Asst Prof. Dr. Mayssaa Al-Bidry

FIGURE 6.8 Development of a throughgoing crack in a block under tension. (a) When tensile stress (σt) is
applied, Griffith cracks open up. (b) The largest, properly oriented cracks propagate to form a throughgoing
crack.

We can also induce tensile fracturing by subjecting a rock cylinder to axial compression, under
conditions of low confining pressure. Under such stress conditions, mesoscopic tensile fractures
develop parallel to the cylinder axis (Figure 6.9a), a process known as longitudinal splitting.
Longitudinal splitting is similar to tensile cracking except that, in uniaxial compression, the cracks
that are not parallel to the σ1 direction are closed, whereas cracks that are parallel to the
compression direction can open up.
In rocks, as the compressive stress increases, the tensile stress at the tips of cracks exceeds the
strength of the rock, and the crack propagates parallel to the compressive stress direction.
In the compressive stress environment illustrated in Figure 6.9, the confining pressure required is
very small; but tensile cracks can also be generated in a rock cylinder when the remote stress is
compressive under higher confining pressure when adding fluid pressure in pores and cracks of the
sample (i.e., the pore pressure; Figure 6.10). The uniform, outward push of a fluid in a microcrack
can have the effect of creating a local tensile stress at crack tips, and thus can cause a crack to
propagate. We call this process hydraulic fracturing. As soon as the crack begins to grow, the
volume of the crack increases, so if no additional fluid enters the crack, the fluid pressure decreases.
Crack propagation ceases when pore pressure drops below the value necessary to create a
sufficiently large tensile stress at the crack tip, and does not begin again until the pore pressure
builds up sufficiently.

7
Asst Prof. Dr. Mayssaa Al-Bidry

FIGURE 6.9 (a) A cross section showing a rock cylinder with mesoscopic cracks formed by the process of
longitudinal splitting. (b) An “envelope” model of longitudinal splitting. If you push down on the top of an
envelope (whose ends have been cut off), the sides of the envelope will move apart.

FIGURE 6.10 (a) Cross-sectional sketch illustrating a rock cylinder in a triaxial loading experiment. Fluid
has access to the rock cylinder and fills the cracks. (b) A fluid-filled crack that is being pushed apart from
within by pore-fluid pressure.

8
Asst Prof. Dr. Mayssaa Al-Bidry

Modes of Crack-Surface Displacement


In the above mentioned, rock cracks are perpendicular to the stress.
But IF rock cracks in other orientations there are three configurations of crack loading. These
configurations result in three different modes of crack-surface displacement (Figure 6.11).
During Mode I displacement, a crack opens very slightly in the direction perpendicular to the crack
surface, so Mode I cracks are tensile cracks. They form parallel to the principal plane of stress that
is perpendicular to the σ3 direction and can grow in their plane without changing orientation. During
Mode II displacement (the sliding mode), rock on one side of the crack surface moves very slightly
in the direction parallel to the fracture surface and perpendicular to the fracture front. During Mode
III displacement (the tearing mode), rock on one side of the crack slides very slightly parallel to
the crack but in a direction parallel to the fracture front.

FIGURE 6.11 Block diagrams illustrating the three modes of crack surface displacement: (a) Mode I, (b)
Mode II, (c) Mode III. Mode I is a tensile crack, and Mode II and Mode III are shear-mode cracks

Joints in rocks
Joint growth is controlled by the mechanical layer thickness of the deforming rock. The
aperture of a joint is the space between its two walls measured perpendicularly to the mean
plane.
Apertures can be open (resulting in permeability enhancement) or filled by mineral cement
(resulting in permeability reduction).
A joint with a large aperture (> few mm) is a fissure. If present in sufficient number, open
joints may provide adequate porosity and permeability such that an otherwise impermeable
rock may become a productive fractured reservoir.

9
Asst Prof. Dr. Mayssaa Al-Bidry

Joint patterns
There are five main arrangements:
- Parallel sets are curved or straight
- Fans sets along fold or intrusion crests
- Radiate sets around intrusion centers
-Concentric sets around intrusion and collapse centers (cone, ring, cylindrical).
- Polygonal sets as columnar or prismatic.

Master joints
Joints that have dimensions ranging from tens of centimetres to hundreds of meters and repeat
distances of several centimetres to tens of meters are called master joints. In addition, most rocks
contain numerous inconspicuous joints of smaller size and closer spacing, some of them, the micro
joints or microfractures, visible only in thin section under the microscope.

Surface morphology of joints


1- Plumose Structure
Plumose structures form at a range of scales, depending on the grain size of the host rock. In very
fine-grained coal, for example, components of plumose structure tend to be much smaller than in
relatively coarser siltstone.
Some of the best examples of plumose structure form in fine-grained rocks like shale, siltstone, and
basalt, but you might not see obvious plumose structure on joints in very coarse-grained rocks like
granite .
A plumose structure spreads outward from the joint origin, which represents the point at which the
joint started to grow. Joint origins typically look like small dimples in the fracture plane. Several
distinct morphologies surround the joint origin.
In the mirror zone, which lies closest to the origin, the joint surface is very smooth. Further from
the origin, the mirror zone merges with the mist zone, in which the joint surface slightly roughens.
Mirror and mist zones, while they are well developed in joints formed in glassy rocks, are
difficult to recognize in coarser rocks. Continuing outward, the mist zone merges with the hackle
zone, in which the joint surface is even rougher.
The plume defines the local direction of joint propagation. The median line may be straight and
distinct, or it may be wavy and diffuse. On some joint surfaces, concentric ridges known as arrest
lines form on the joint surface at a distance from the origin. These ridges represent, as the name
indicates, breaks in the growth of the joint.

10
Asst Prof. Dr. Mayssaa Al-Bidry

11
Asst Prof. Dr. Mayssaa Al-Bidry

2-Twist Hackle
Features such as bedding planes and pre-existing fractures locally modify the orientation of
principal stresses because they approximate free surfaces. If a growing joint enters a region where
it no longer parallels a principal plane of stress (for example, as occurs when the crack tip of a joint
in a sedimentary bed approaches the bedding plane), the crack axes to a new orientation.
(Hackles diverge sharply at angles of about 30° from the central axis, gradually curving to
angles of about 70° near the margins of the joint surface. The scale of plumose patterns seems
to depend on the grain size of the rock).
the joint splits into a series of small en echelon joints, because a joint surface cannot twist and still
remain a single continuous surface. The resulting array of fractures is called twist hackle, and the
edge of the fracture plane where twist hackle occurs is called the hackle fringe (Figure 7.3a).

Joint arrays
1- Systematic versus Non-systematic Joints
Systematic joints are planar joints that comprise a family in which all the joints are parallel or sub
parallel to one another, and maintain roughly the same average spacing over the region of
observation. Systematic joints may cut through many layers of strata, or be confined to a single
layer.
Non-systematic joints have an irregular spatial distribution, they do not parallel neighbouring
joints, and they tend to be non-planar. Non-systematic joints may terminate at other joints. You will
often find both systematic and non-systematic joints in the same outcrop.

12
Asst Prof. Dr. Mayssaa Al-Bidry

2- Joint Sets and Joint Systems


A joint set is a group of systematic joints. Two or more joint sets that intersect at fairly constant
angles comprise a joint system, and the angle between two joint sets in a joint system is the
dihedral angle.
If the two sets in a system are mutually perpendicular (i.e., the dihedral angle is 90°), we call the
pair an orthogonal system (Figure 7.6a), and if the two sets intersect with a dihedral angle
significantly less than 90° (e.g., a dihedral angle of 30° to 60°), we call the pair a conjugate system
(Figure 7.6a).
Many geologists use the terms “orthogonal” or “conjugate” to imply that the pair of joint sets
formed at the same time (non-parallel joint sets typically form at different times. (So, we use the
terms merely to denote geometry, not a mode or timing of origin).
In joint systems where one set consists of relatively long joints that cut across the outcrop whereas
the other set consists of relatively short joints that terminate at the long joints, the through going
joints are master joints, and the short joints that occur between the continuous joints are cross
joints.
Both orthogonal and conjugate systems occur in such gently folded strata. The joint sets of an
orthogonal system in folded sedimentary rocks commonly have a spatial relationship to folds of
the region.

13
Asst Prof. Dr. Mayssaa Al-Bidry

Joint Spacing in Sedimentary Rocks


Joint spacing as the average distance between adjacent members of a joint set, measured
perpendicular to the surface of the joint. Informally, geologists refer to joints as being “closely
spaced” or “widely spaced” in a relative sense, but to be precise, you should describe joint spacing
in units of length (e.g., 5 cm).

14
Asst Prof. Dr. Mayssaa Al-Bidry

Experimental work suggests that joints form in sequence; that is, first joint 1, then joint 2, then joint
3, and so on. When a new joint forms, it is at some distance greater than a minimum distance from
a preexisting joint. Formation of a joint relieves tensile stress for a critical distance .
The zone on either side of a joint in which there has been a decrease in tensile stress is called the
joint stress shadow. Stresses sufficient to create the next joint are only achieved outside of this
shadow and are created by traction between the bed and beds above and below it, as well as by
stress transmitted within the bed beyond the fracture front of the preexisting joint. The spacing
between joints is determined by the width of the joint stress shadow; so, because the shadow is
about the same width for all joints in the bed, the spacing ends up being constant.
Joint spacing depends on four parameters: bed thickness, stiffness, tensile strength, and strain. All
other parameters being equal, joints are more closely spaced in thinner beds, and are more widely
spaced in thicker beds. The relationship is a reflection of joint-stress shadow width, because the
greater the length of the joint (i.e., length of the joint trace in a plane perpendicular to bedding and
joint), the wider the stress shadow (Figure 7.9b and c).

15
Asst Prof. Dr. Mayssaa Al-Bidry

Origin and interpretation of joints


1- Joints Related to Uplift and Unroofing
Lithostatic pressure due to the weight of overlying rock compresses rock at depth. Also, because of
the Earth’s geothermal gradient, rock at depth is warmer than rock closer to the surface. Subsequent
regional uplift leads to erosion of the overburden and the unroofing of buried rock.
As the burial depth of rocks decreases, they cool and shrink. The rock can shrink in a vertical
direction without difficulty, because the Earth’s surface is a free surface. But, because the rock is
embedded in the earth, it is not free to shrink elastically in the horizontal direction as much as if it
were unconfined, so horizontal tensile stress develops in the rock.

16
Asst Prof. Dr. Mayssaa Al-Bidry

Furthermore, as the overburden diminishes, rock expands (very slightly) in the vertical direction.
Therefore, because of the Poisson effect (Poisson's ratio is a measure of the Poisson effect, the
phenomenon in which a material tends to expand in directions perpendicular to the direction of
compression), it contracts in the horizontal direction.

2- Natural Hydraulic Fracturing


Three principal stresses at depth in most of the continental lithosphere are compressive. Yet joints
form in these regions, and these joints may be decorated with plumose structure indicating that they
were driven by tensile stress.
How can joints form if all three principal stresses are compressive? The solution comes from
considering the effect of pore pressure on fracturing.
the increase in pore pressure in a pre-existing crack pushes outward and causes a tensile stress to
develop at the crack tip that eventually exceeds the magnitude of the least principal compressive
stress. If the pore pressure is sufficiently large, a tensile stress that exceeds the magnitude of σ3
develops at the tip the crack, even if the remote principal stresses are all compressive, and the crack
propagates, a process called hydraulic fracturing.
Oil well engineers commonly use hydraulic fracturing to create fractures and enhance
permeability in the rock surrounding an oil well. They create hydraulic fractures by
increasing the fluid pressure in a sealed segment of the well until the wall rock breaks.
Hydraulic fracturing also occurs in nature, due to the fluid pressure of water, oil, and gas in
rock and it is this natural hydraulic fracturing that causes some joints to form. The pores and
the crack are connected, the fluid pressure in the pores and the crack are the same. Fluid pressure
within the crack is pushing outwards, creating an opening stress, but at the same time, the fluid
pressure in the pores, as well as the stress in the rock, is pushing inwards, creating a closing stress.
If the closing stress exceeds the opening stress, the crack does not propagate. If the fluid pressure
increases, the opening stress increases at the same rate as the increase in fluid pressure, but the
closing stress increases at a slower rate.
Eventually, the opening stress exceeds the closing stress so that the crack propagates; effectively,
the outward push of the fluid in the crack creates a tensile stress at the crack tip.
Why does the closing stress increase at a slower rate than the fluid pressure and the opening
stress?
The reason is that grains in the rock are cemented to one another, so that the grains cannot move
freely in response to the increase in fluid pressure in the pores. The elasticity of the grains
themselves, therefore, takes up some of the push caused by the fluid pressure. Thus, the closing
stress acting on the fluid in the crack where it is in contact with a grain is less than where the fluid
in the crack is in contact with a pore, but the outward push of the fluid in the crack is the same
everywhere. As a result, the net outward push exceeds the net inward push, and tensile stress locally
develops. Once the crack propagates, the volume of open space between the walls of the crack
increases, so the fluid pressure in the crack decreases. As a consequence, the crack stops growing
until an increase in fluid pressure once again allows the stress intensity at the crack tip to drive the
tip into unfractured rock. Thus, the surfaces of joints formed by natural hydraulic fracturing tend to
have many arrest lines.

17
Asst Prof. Dr. Mayssaa Al-Bidry

Hydraulic Fracturing
Hydraulic fracturing (informally known as hydrofracking, fracking,bor hydrofracturing) is a
process that typically involves injecting water, sand, and chemicals under high pressure into a
bedrock formation via a well. This process is intended to create new fractures in the rock as well as
increase the size, extent, and connectivity of existing fractures in order to extract trapped oil and
gas.
Hydraulic fracturing is a well-stimulation technique used commonly in low-permeability rocks like
tight sandstone, shale, and some coal beds to increase oil and/or gas flow to a well from petroleum-
bearing rock formations.
A similar technique is used to create improved permeability in underground geothermal reservoirs.
Frac sand is a specialized type of sand that is added to fracking fluids that are injected into
unconventional oil and gas wells during hydraulic fracturing. Frac sand keep induced fractures open
and extend the time and flow rate of oil and gas from a well.

3 - Joints due to non-de compressional volume changes


Stylolitic joints have a characteristic saw-tooth. The interlocking ‘teeth’ are normal or oblique to
the joint surface. Stylolitic joints are surfaces along which relatively soluble rock material has been
removed by pressure-induced chemical dissolution to accommodate shortening.
Shortening is parallel to the teeth direction. Dissolution is triggered by stress concentration at the
contact between grains. Relatively insoluble residues (clay, iron oxides, etc) remained accumulated
in the joint. This deformation mechanism is called pressure solution.
Stylolites are particularly common in limestone. Amplitudes represent a minimum estimate of the
amount of shortening (compaction) that has occurred. Assuming that insoluble material was
initially evenly distributed in the rock and that there has been no contamination by circulating fluids,
the thickness of insoluble residue along a stylolitic joint would be proportional to the amount of
material dissolved and would therefore be proportional to the shortening displacement across the
stylolitic joint.

18
Asst Prof. Dr. Mayssaa Al-Bidry

Mechanics of jointing
A genetic classification of joints has been based on the size of inferred, imperceptible displacement
related to the three principal stress axes of a region. If the total displacement is normal to the fracture
surface, it is an extension or dilatant joint (mode 1 fracture). If the shear component has some finite,
yet negligible value, the fracture called a shear joint is really a fault (modes 2 and 3 fracture),
keeping in mind that the shear component may have accumulated after formation of a former
extension (dilatant) joint.
1- Extension joints
Linear elastic fracture mechanics predicts that the orientation of extension joints in a relatively
isotropic rock is controlled by the remote stress field at the time of fracture propagation:
joints are gaping planes parallel to the maximum compressive stress σ1 and perpendicular to the
direction of the least principal stress σ3.
In other words, they form in the plane containing σ1 and σ2. Otherwise, there would be a shearing
stress and a corresponding finite shear displacement on the joint plane.
The pattern of extension joints is commonly T-shaped, the younger joint abutting the older one.
Given suitable anisotropy of the tensile strength, it is however possible to get joints normal to σ2
or even σ1.
2- Hybrid joints
Hybrid joints show components of both extension and shear components. They are interpreted as
failure surfaces initiated in the transition from tensile to shear failure.
3- Shear joints
This term is unfortunate and ambiguous because shear joints are small faults. Conjugate “shear
joints” generally define X, Y or V shapes. The acute bisector of these shapes is parallel to σ1, unless
these patterns represent unrelated crosscutting or abutting fractures.

19
Asst Prof. Dr. Mayssaa Al-Bidry

The Nature of Naturally Fractured Reservoirs


Fractures are the most abundant visible structural features in the Earth’s upper crust. They are
apparent at most rock ridges, and it is likely that most reservoirs contain some natural fractures.
Naturally fractured reservoirs are elusive systems to characterize and difficult to engineer and
predict. Nearly all hydrocarbon reservoirs are affected in some way by natural fractures.
In carbonate reservoirs, natural fractures help create secondary porosity and promote
communication between reservoir compartments. However, these high-permeability conduits
sometimes short-circuit fluid flow within a reservoir, leading to premature water or gas production
and making secondary-recovery efforts ineffective.
Natural fractures also occur in siliciclastic reservoirs of all types.
In addition, natural fractures are the main producibility factor in a wide range of less conventional
reservoirs, including shale-gas, basement rock and volcanic-rock reservoirs.
Natural fractures play a lesser role in high porosity, high-permeability reservoirs, they
commonly form barriers to flow, frustrating attempts to accurately calculate recoverable reserves
and predict production over time.
Ignoring the presence of fractures is not optimal reservoir management; eventually, fractures cannot
be ignored because the technical and economic performance of the reservoir degrades.
How naturally fractured reservoirs are formed
Unlike induced fractures, natural fractures are caused by stress in the formation usually from
tectonic forces such as folds and faults. Fractures occur in preferential directions, determined by
the direction of regional stress. This is usually parallel to the direction of nearby faults or folds, but
in the case of faults, they may be perpendicular to the fault or there may be two orthogonal
directions

20
Asst Prof. Dr. Mayssaa Al-Bidry

Open and healed fractures


A fracture is often a high permeability path in a low permeability rock, or it may be filled with a
cementing material, such as calcite, leaving the fracture with no permeability. It is important to
distinguish between open and healed fractures. The total volume of fractures is often small
compared to the total pore volume of the reservoir.
Size
Naturally fractured reservoirs are observed across a vast range of scale from microcracks to mile
long features. The vertical extent of fractures is often controlled by thin layers of plastic material,
such as shale beds or laminations, or by weak layers of rock, such as stylolites in carbonate
sequences. The width of these beds may be too small to be seen on logs, so fractures may seem to
start and stop.
Types
Naturally fractured reservoirs can be open, permeable pathways, or they can be permeability baffles
resulting from the presences if secondary mineralization or other fine-grained material filling the
gaps. Most natural fractures are more or less vertical. Horizontal fracture may exist for a short
distance, propped open by bridging of the irregular surfaces. Most horizontal fractures, however,
are sealed by overburden pressure. Both horizontal and semi-vertical fractures can be detected by
various logging tools.

Classification
Naturally fractured reservoirs can be classified in different types, depending on the storage
capacities or porosity and permeability of the matrix and the fractures. Different definitions for
these types can be found in literature.
Aguilera classified the naturally fractured reservoirs in types A, B and C (see Figure 1). In
reservoirs of type A most fluid is stored in the matrix; the fractures provide only a very small storage
capacity. Typically the matrix rock tends to have a low permeability, whereas the fractures exhibit
a much larger permeability. In type B reservoirs approximately half of the hydrocarbon storage is
in the matrix and half in the fractures. The fractures provide the storage capacity of type C
reservoirs, without contribution of the matrix

Figure 1: Porosity distribution in fractured rocks


Fractured reservoirs are classified by Nelson based on the interaction between the relative porosity
and permeability contributions from both the fracture and matrix systems (see Figure 2).

21
Asst Prof. Dr. Mayssaa Al-Bidry

In Type 1 reservoirs, fractures provide both the porosity and permeability elements.
Type 2 reservoirs have low porosity and low permeability in the matrix, and fractures provide the
essential permeability for productivity.
Type 3 reservoirs have high porosity and may produce without fractures, so fractures in these
reservoirs provide added permeability.
Type M reservoirs have high matrix porosity and permeability, so open fractures can enhance
permeability, but natural fractures often complicate fluid flow in these reservoirs by forming
barriers.
Fractures add no significant additional porosity and permeability to Type 4 reservoirs, but instead
are usually barriers to flow.
Another reservoir class, Type G, has been created for unconventional fractured gas reservoirs.

Figure 2. Naturally fractured reservoir classification according to Nelson

Type 1 reservoirs, with fractures providing both primary porosity and primary permeability,
typically have large drainage areas per well, and require fewer wells for development. These
reservoirs show high initial production rates. They are also subject to rapid production decline, early
water breakthrough and difficulties in determining reserves.
Type 2 reservoirs can have surprisingly good initial production rates for a low-permeability matrix
but can have difficulties during secondary recovery if the communication between the fracture and
the matrix is poor.
Type 3 reservoirs are typically more continuous and have good sustained production rates but can
have complex directional permeability relationships, leading to difficulties during the secondary-
recovery phase.
Type M reservoirs have impressive matrix qualities but are sometimes compartmentalized, causing
them to underperform compared with early producibility estimates, and making secondary-recovery
effectiveness variable within the same field.
Type 4 reservoirs would plot near the origin because the fracture contribution to permeability in
Type 4 reservoirs is negative

22
Asst Prof. Dr. Mayssaa Al-Bidry

Fractured Rocks Properties


Different fracture properties affect the reservoir performance of a naturally fractured reservoir. The
fluid flow properties of the fractures include fracture porosity, fracture permeability, or the fluid
saturation within the fracture system. Another important factor is the wettability of the rock and
possible wettability changes during the production time of the reservoir

1 Porosity been little mutual displacement along the


Porosity can be classified as primary or fracture.
secondary.
Primary porosity forms during
deposition of sediments and includes
interparticle and intraparticle porosities.
Secondary porosity forms after
deposition and develops during
diagenesis by dissolution, dolomitization
and through production of fractures in the
rock.
The matrix porosity, also often called
fabric porosity, can be both primary and
secondary.
The fracture porosity is always a
secondary one and generally refers to
porosity that occurs along breaks in a
sediment or rock body where there has

The fractures enable fluid movement and as a consequence solution of minerals.


Depending on the extent of solution, the resulting pores are classified as molds, solution
enlarged molds or vugs.
Vuggy porosity is a non-fabric selective porosity caused by selective removal (solution) of
grains in a rock. If vugs and molds are connected by fractures then their volume become
part of the fracture porosity.
In carbonate rocks, fracture porosity may originate from collapse related to solution, or
tectonic deformation.
Fractures can be observed on cores, and can be characterized as filled, semi-filled and open
fractures. Filled fractures do not contribute to the porosity. The fractures are described
by their orientation as horizontal, vertical or oblique fractures.
Let φf be the fracture porosity and φm the matrix porosity, then the storativity dimensionless
parameter ( storativity or the storage coefficient)

expresses the ratio between the storage capacity of the fracture network and the total
storage capacity.

23
Asst Prof. Dr. Mayssaa Al-Bidry

2. Permeability
The permeability of a porous rock is a measure of the ability to transmit fluids. A reservoir
can have primary and secondary permeability.
The primary permeability is referred to as matrix permeability; the secondary permeability
can be either called fracture permeability or solution vugs permeability.

Matrix- and fracture permeability are other important parameters that have to be known for
an estimate of the influence of the fractures on the overall reservoir performance.

Solution vug permeability refers to an increased permeability in matrix rocks (especially


in carbonate reservoirs) where the natural permeability of the matrix is increased by
percolation of acid waters that dissolve the matrix rock.
The permeability in these flow channels can be calculated by combining Darcy’s law for
fluid flow and Poiseuille’s law for capillary flow.

Open fractures in Naturally Fractured Reservoirs generally have a higher permeability than
the matrix, building the flow channels of the system. The flow rate through a narrow
cleavage can be calculated by Lamb’s law:

Where W is the effective fracture aperture (fracture width). The fracture cross section A is
the product of the fracture width W and the breadth b:

A=W⋅b (1-3)

μ is the viscosity, and dp/dx is the pressure gradient.


The flow rate can also be expressed by the Darcy equations:

Both Equation 1 and Equation 3 are valid for laminar flow.


So it is evident that the permeability of a single fracture is:

A fracture with 10-5 m width (i.e.: 0.1 mm) has a permeability of 844 Darcy. As a
consequence of Equation 1.2 and Equation 1.3, between two flat plates, the flow rate is
proportional to the cube of the aperture W.
This is naturally not valid for natural fractures because they are rough.

24
Asst Prof. Dr. Mayssaa Al-Bidry

Figure 2: Parallel fractures in flow direction

The effective permeability in a fractured solid cube, shown in Figure 2 is:

Where

is the fracture porosity. Inserting Equation 1.7 in Equation 1.6 results in

Note that as a consequence of the Equation 1.5 and Equation 1.6 the effective permeability
is proportional to the cube of the aperture W:

If the matrix is also permeable, then the overall effective permeability is:

The approximation is valid if ᶲf « 1 .


Note, that Equation 1.5 and Equation 1.6 cannot be used for real fractures in porous rocks,
because it is derived for steady state, isothermal, laminar flow between parallel glass plates.
Fracture permeability is, similarly to the fracture porosity, highly scale-dependent.
A fracture of width W expressed in inches has a permeability of:

The resultant intrinsic permeability of a fracture of 0.01 in. would be 5400 darcys. The
intrinsic permeability of Equation 1.11 is valid for a single point. The formulation can be
extended for the bulk properties of the system for one set of parallel fractures:

where D is the distance between the fractures.

25
Asst Prof. Dr. Mayssaa Al-Bidry

3. Compressibility
The stress on the reservoir rock is determined by the confining and the pore pressures. The
confining (or overburden) pressure, caused by the weight of overlying rock is partially
compensated by the pressure of the fluids in the pores.
The net confining pressure, pe, is the difference of the two pressures:

A number of investigations indicate that the effect of varying the confining and pore
pressure on porosity and permeability is mainly governed by the net confining pressure
and is not greatly dependent on the absolute values of either total confining pressure or
pore fluid pressure.

Figure 3 shows a typical stress-strain


curve manifesting three regions. The
linear region of elastic deformation exists
up to a stress called yield stress. Beyond
that, the material shows plastic behavior.
Increase in stress causes a non-linear
increase of strain and if the strain is
relaxed, the response curve does not
retrace the original load path but rather
follows an elastic path typical of a more
consolidated rock. Ultimately, if enough
stress is applied, the rock becomes fully
compacted and the stress/strain relation
regains linearity. For consolidated
sandstone the yield point may exceed
Figure3 Typical rock stress/strain curve
1380 bara (20,000 psia), while for soft showing three regions of behavior: elastic,
chalk it can be as low as 60-70 bara (800 plastic pore collapse, and compacted work-
to 1,000 psia). hardening

26
Asst. Prof. Dr. Mayssaa Al-Bidry

Note, that also if often pore collapse and compaction will be modelled by increased
compressibility factor, the following discussion is valid only for the elastic state of the
reservoir rocks. Dealing with deep carbonate reservoirs, the first step must always be
to estimate or better yet, to measure the yield point.
The isothermal compressibility factor, in general, is defined as the specific volume
change caused by change of pressure:

The volume V may refer to the bulk volume (Vb), the solid volume (Vs) or the fluid,
e.g. the oil volume (Vo).
where cs is the compressibility factor of the solid phase. The matrix block, tight or
porous, surrounded by fractures will expand towards the fractures, therefore the
compressibility of the fracture porosity is determined by the compressibility of the
matrix bulk volume:

If the matrix is tight, then

Reference:
1-Natural Fractured Reservoir Engineering. Zoltán E. and Georg Mittermeir.2014
2-The Nature of Naturally Fractured Reservoirs. Tom Bratton, Dao Viet Canh, Nguyen
Van Que , Nguyen V. Duc, Paul Gillespie, David Hunt, Bingjian Li, Richard Marcinew, Satyaki
Ray, Bernard Montaron, Ron Nelson, David Schoderbek and Lars Sonneland. Article in
Oilfield Review. 2014
3-Ben A. Van Der Pluijm and Stephen Marshak . Earth Structure, Second edition.
2004.

27

You might also like