Anda di halaman 1dari 92

1 2 HYDRAULIC FRACTURING

2 3 HYDRAULIC FRACTURING

2-115

3 HYDRAULIC FRACTURING
Fracturing is used in order to break down a producing formation hydraulically
with a sand carrying fluid, the sand being used to prop the resulting fracture.
Hydraulic fracturing is a well mechanical stimulation technique aimed to
improve the matrix permeability surrounding a wellbore and create deep
penetrating fractures that provide high capacity channels for the flow from
deep within the producing formation to the wellbore.
A fracture treatment should be used in order to increase the original
permeability of the reservoir near the wellbore. The injected fluid is pumped at
a rate above the fracture pressure of the reservoir in order to create cracks or
fractures within the rock, thus crating new conductive channels through which
oil or gas may flow easily to the wellbore. Once he pressure is released after
the treatment the induced fracture will tend to close, so a proppant of sand or
small beads will normally be mixed with the treating fluid in order to keep the
fracture open after the treatments finished.

3.1.1 Basics of Rock Mechanics


Rocks generally obey the same laws of mechanics as other materials such as
metals, but because of discontinuities and inhomogeneities, they are
sufficiently different that a relatively new branch of engineering, rock
mechanics, has developed. Rock mechanics and the related rock stress
condition are important in many areas of oil and gas production operations.
In drilling, penetration rate, lost circulation, abnormal pressures, hole problems
and hole eccentricity are related to rock mechanics. In primary cementing, the
maximum displacement rate, cement column height and lost circulation are
related to rock mechanics. In sand control, formation strength and gravel
placement are related to rock mechanics. In hydraulic fracturing, fracture
initiation and propagation pressures, fracture geometry (length, width and
height), required proppant strength and fracture conductivity are related to rock
mechanics. In fracture acidizing and matrix acidizing, rock mechanics play a
similar role. In reservoir engineering, even the basic concepts of porosity and
permeability are related to rock mechanics.
Various rock properties and elastic constants can be measured on cores in the
laboratory simulating downhole conditions. Downhole logging measurements
present an approach to in-situ determination of rock properties. In an ideal
relaxed geologic area, rock stress conditions could be approximated from
simple relations using these measurements. Since earth tectonics also influence
in-situ stress conditions, simple calculations per se are usually insufficient to
fully characterize actual stress conditions. Actual rock stress conditions can be
determined more directly by step-rate or flowback measurements during
3-116

hydraulic fracturing or other well treatments involving pressures in excess of


the fracture pressure.
The following discussion defines some rock mechanics parameters and the
relations between them, and briefly describes how they can be determined.
3.1.2 MECHANICAL PROPERTIES OF ROCKS
Rocks are inhomogeneous composite materials containing different crystals.
Discontinuities and micro-cracks are randomly oriented throughout this
material. Since rocks have a very low tensile strength and since they are
normally under compressive stresses in the earth, most measurements of rock
properties are made by applying compressive loads. As compressive load is
applied to a rock, the micro-cracks, particularly those that are perpendicular to
the load, begin to close. Much of this initial strain is not related to the
deformation of the crystals themselves.
Rock Strength
Rock strength can be specified in terms of tensile strength, compressive
strength, shear strength or impact strength. In the context of fracture gradient,
only the tensile strength of rock is of importance. The tensile strength of rock is
defined as the pulling force required to rupture a rock sample divided by the
samples cross-sectional area. The tensile strength of rock is very small and is
of the order of 0.1 of the compressive strength. Thus, a rock is more likely to
fail in tension than in compression.
3.1.3 Stress-Strain Relations
Any material subject to tensile or compressive load will deform (Figure 3-1).
Load or force per unit area is called stress, ( ). Under stress, the deformation
or change in length, ( ), compared to the original length, (l), is called strain,
( ). At low stress levels, a plot of stress versus strain is a straight line for
linear elastic materials such as metals. Figure 3-2 shows this relationship.
Obviously, there is a maximum stress that a material can sustain before failure.
Ductile materials, such as steel, exhibit a plastic, or yielding, behavior such that
as the yield strength is exceeded, strain increases disproportionately with stress
until ultimate failure results. With brittle materials such as rock, this failure
occurs suddenly with little additional strain. The compressive stress required to
cause failure is called the uniaxial compressive strength, (Co) of the material.

3-117

Figure 3-1: unaixal and triaxial tests

3-118

Figure 3-2: Illustrates the stress versus strain in rocks

3-119

Stress
Is an applied external force on a solid body that causes internal resulting forces
to exist within the body whose resultant force will be equal in magnitude but
opposite in direction to the applied force. A stress could be tension if it tends to
elongate the subjected body or compression if it tends to contract the subjected
body.
=

F
A

3-1

Where
= Stress
F = applied force
A = cross sectional area
A stress is called normal () if it the fractured area is perpendicular (normal) to
the direction of the applied force. A stress is called shear stress ( ) if the
fractured area is parallel to the direction of the applied force.
Strain
Is the resultant deformation of a body, as a function of its original dimensions,
caused by an applied force (stress).
Tensile strain =

Lo L L
=
Lo
Lo

3-2

Where
Lo = original length
L = elongated length
L = the longitudinal strain
Shearing strain = tan

3-3

Where
= The deformation or fracture angle
Volume strain =

V
Vo

Where
V = Volume increase
Vo = Original volume

3-120

3-4

Figures 3-3 shows a bar whose natural length is Lo and which elongates to a
length L when equal and opposite pulls are exerted at its ends. Figure 3-4
illustrates the direction of the applied stress on a steel bar with the
corresponding axial and lateral strains. The resultant shear deformation which
causes change in shape of a body is illustrated in figure 3-5.
HOOKE'S LAW
When a body is subjected to a specific stress, it will undergo a specific strain. If
the body returns to its original dimension upon removal of the stress, the action
is said to be elastic. However, if upon the removal of the stress the body does
not return to its original dimensions, and there is a residual strain, the action is
said to be inelastic. Hooks observed that stress is directly proportional to the
strain. Figures 3-6 and 3-7 show a typical stress-strain curve for a rock.
The relationship between stress and the strain can be represented by the
following equation:
= E

3-5

Where
= stress
= strain
E = constant of proportionality
YOUNGS MODULUS
The amount of strain caused by a given stress is a function of the stiffness of a
material. Stiffness can be represented by the slope of the axial stress-strain plot
and is termed the Youngs modulus (E). E is defined as the young's modulus or
the modulus of elasticity and it is represented by the following equation:
E=

Tensile stress Compressive stress lb / in 2


2
=
= lb / in

Tensile strain Compressive strain in / in


E=

F/A
L/Lo

3-6

E=

stress
=
strain

3-7

3-121

Figure 3-3: Illustrates the longitudinal strain due to


elongation of steel bar.

Figure 3-4: Lateral strain vs. axial strain in a body.

3-122

Figure 3-5: Change in shape of a body in shear.

Figure 3-8: Definition of shear modulus

3-123

For mild steel, the Youngs modulus or modulus of elasticity is 30 x 106 psi.
For rock, E values range from 0.5 to 12 x 106psi. Common values for average
youngs modulus for different rock hardness are illustrated in table 3.1
SHEAR MODULUS (MODULUS OF RIGIDITY)
Shear stress applied to a particular plane surface in a block of material causes
that plane to move with respect to a second parallel plane some perpendicular
distance away as shown in Figure 3-8. The ratio of the applied shear stress to
the resulting angle of deformation is a measure of the rigidity of the material.
This ratio is termed the shear modulus or modulus of rigidity (G). For a fluid,
G = 0; for a solid, G is a finite number. The ratio of a shearing stress to the
corresponding shearing strain is called the shear modulus of a material and will
be represented by G. It is also called modulus of rigidity or the torsion
modulus.
G=

G=

Shearing stress
Shearing strain

3-8

shear stress
F / A lb / in 2
=

radian
resultant angle of

deformation in radians

The shear modulus G can also be determined from sonic log reading by the
following equation:
G = 1.34 1010

b
t s

3-9

Where
b = bulk density read from density log, gm/cc (Figure 3-9)
ts = shear sonic wave transmit time log, sec (Figure 3-10)
The shear modulus is related to the youngs modulus, E as follows:
E = 2 G (1 + )

G=

E
2 (1 + )

3-124

3-10

Figure 3-9: Illustrates the density log reading versus depth.

3-125

Figure 3-10: Shows the compressional and the shear sonic transit wave
time for a typical reservoir rock.

3-126

PLAIN STRAIN MODULUS


The plain strain modulus can be defined as follows:
E' =

3-11

12

BULK MODULUS
Compressive load applied on all sides of a block of material, as occurs in a
hydrostatic condition, causes a reduction in total bulk volume. The ratio of
stress applied (force per unit surface area) to the change in volume per unit of
original volume is termed the bulk modulus (Kb).
lb / in 2
force / surface area
Kb =

change in volume / original volume in 3 / in 3


Kb =

F/A
v/v

3-12

The modulus relating an increase in hydrostatic pressure to the corresponding


fractional decrease in volume is called the bulk modulus and it is represented
by Kb.

Kb = -

p
V/Vo

3-13

The Bulk Modulus related to the Youngs Modulus by the following equation:
K=

E
3 (1 2)

The bulk modulus can also be calculated using the sonic log data as described
in the following equation:

1
4
K b = b 2
x 1.34 x 1010
2
t c 3t s
Where:
tc
ts
b

= sonic compressional wave transit time log, sec


= sonic shear wave transit time log, sec
= bulk density read from density log, gm/cc

3-127

3-14

The reciprocal of the bulk modulus is called the rock bulk compressibility with
porosity, Cb.

1
1 V
=Kb
p Vo

Cb =
Cb =

3-15

1
1
4
Pb 2
x 1.34 x 1010
2

t
3

t
s
c

The ratio V/Vo is the fractional change in volume. Hence the compressibility
of a substance is defined as its fractional change in volume per unit increase in
pressure. The rock grain compressibility with zero porosity can be defined as:
Cr =

change in matrix volume


hydrostatic pressure

The rock grain compressibility with zero porosity is calculated from the
following equation:
Cr =

1
1
4

Pb
2
2
t ma 3t sma

3-16

x 1.34x 1010

Where
tma = matrix compressional sonic wave transit time, sec
tsma = matrix shear sonic wave transit time, sec
PORE ELASTICITY CONSTANT
To account for the anisotropically of the stressed formations, the poroelasticity
() should be considered and it is defined a follows: The constant is called
the poroelastic constant or the Biot elastic constant and is defined as follows:
=1

Bulk Modulus of Dry Rock


Bulk Modulus of Skeleton Material
= 1

Cr
Cb

1 > 0

3-17

Use of this model is reasonable for low porosity, low permeability sandstones,
shales and carbonates. Parameter is known as the Biots poroelastic

3-128

parameter which approaches the greater limit for a compliant rock and less for
a stiff low-porosity rock.
POISSON'S RATIO:
Compressive stress applied to a block of material along a particular axis causes
it to shorten along that axis but also to expand in all directions perpendicular to
that axis as illustrated in Figure 3-11. The ratio of the strain perpendicular to
the applied stress, to strain along the axis of applied stress, is termed Poissons
ratio ( ). The Poissons ratio is also defined as the ratio of the absolute value
of strain in the lateral direction to the strain in the axial direction.
=

lateral strain
axial strain

in / in

in / in

A material that under stress deforms laterally as much as it does axially would
have a Poissons ratio of 0.5. A material that does not deform laterally under
axial load would have a Poissons ratio of 0.9. Mild steel has a Poissons ratio
of about 0.3. In general, limestone, sandstone, shale and salt, exhibit Poissons
ratios of approximately 0.15, 0.25, 0.4 and 0.5 respectively. Figure 3-11
illustrates the Poissons ratio measurement of solid body. The Poissons ratio
is defined as:
=

x
z

3-18

The Poissons ratio can be estimated from sonic log readings as:
2
1 t s
1

2 t
= c
2
t s

t 1
c

3-19

Where
tc =sonic compressional wave transit time, sec
ts = sonic shear wave transit time, sec
Common values for Poissons ratio and Youngs modulus for different rock
compressive strength are presented in table 3.1.
Young's modulus and Poissons ratio are functions of the rock's hardness and
elasticity and are used in the fracture hardness. These two elastic rock
properties are usually determined by a compressive strength measurement in
the laboratory in which loads and deformations are measured. However, the
3-129

properties are sometimes deduced from sonic log determinations. Figure 3-12
illustrates the relationship between the sonic transit travel time and the youngs
modulus for some sedimentary rocks.
Table 3-1: Variation of Youngs Modulus and Poissons Ratio with rock type
Rock type
Limestone
Unconsolidated sandstone
Consolidated sandstone
Siltstone
shale
coal

Youngs modulus (106


psi)
5 - 13
0.2 - 1.3
1-3
4-5
1-5
0.1 - 1.0

Poissons ratio
0.3 - 0.35
0.25 - 0.35
0.15 - 0.30
0.20 - 0.30
0.25 - 0.45
0.35 - 0.45

3.1.4 MECHANICS OF FRACTURING


The mechanics of fracture initiation and extension and the resulting fracture
geometry are related to the stress condition near the borehole and in the
surrounding rock, the properties of the rock, the characteristics of the fracturing
fluid and the manner in which the fluid is injected. Hubbert and Willis
presented a simplified fracture mechanics theory which seems to explain many
of the events observed in field operations during squeeze cementing, gravel
packing, hydraulic fracturing and some instances of lost circulation during
drilling. Refinements of the basic theory are being developed, but knowledge
of the Hubbert and Willis theory provides a basis for understanding formation
fracturing. The work of Howard and Fast, Perkins and Kern, Nordgren,
Khristanovich, Geertsma and de Klerk, Daneshy and Cleary form the basis for
analysis of fracture extension and fracture geometry.
3.1.5 REGIONAL ROCK STRESSES (THE IN SITU PRINCIPAL
STRESSES)
Subsurface rocks are normally in a state of compressive stress due to the
weight of the overburden. This overburden weight creates stresses in both the
vertical and horizontal directions. Sedimentary rocks have little inherent tensile
strength, rather are held together by compressive stresses. A fracture is
extended when sufficient differential hydraulic pressure is applied to overcome
these compressive stresses.
At any point below the earths surface, three mutually perpendicular stresses
exist as shown in Figure 3.13. The maximum principal stress, V, is normally
vertical and is equal to the overburden stress in vertical holes. The intermediate
and minimum total principal stresses (H and h, respectively) are horizontal
and directly influence the fracturing of rock. In theory, the fluid pressure
3-130

required to rupture a borehole should be greater than or equal to the minimum


principal stress. However, the creation of a borehole within the earths surface
produces a magnification of stresses around the borehole wall such that the
resulting stresses are several times larger than the least principal stress.
Rocks are fractured when the applied forces are greater than the underground
stresses. The stresses that are exerted on a subsurface formation can be
represented by components in three directions. These forces that act on the
rocks are shown in figures 3-13 and 3-14.
Rocks buried deep in the ground are subjected to high stresses. These usually
differ in different directions since they originate from many different sources.
For instance, the vertical stress at a particular depth will be due, essentially, to
the weight of overlying formations. Hence, this is also known as the
overburden stress. The effect of this overburden stress will tend to spread or
expand the underlying rocks in the horizontal lateral directions (i.e. due to the
Poisson effect). However, as outlined earlier, this tendency for lateral
movement will be restrained by the presence of adjacent material, and therefore
horizontal lateral stresses which confine the rock will result. Regional tectonic
stresses, such as might cause earthquakes or mountain-building, contribute
further to these horizontal stresses. Temperature increases or reductions lead to
thermal expansion or contraction, the effects of which also contribute to the
stresses in the ground. The result is that, in the undisturbed state before any
engineering activity, the state of stress in the rock (termed the far field stresses,
the in situ stresses or the virgin stresses) will generally be compressive and can
be simplified and approximated to:

Vertical overburden stress, V

Maximum horizontal stress, H

Minimum horizontal stress, h

The vertical or overburden stress V, which has the greatest magnitude and
is therefore known as the maximum in situ stress
The maximum horizontal stress H, which has a lower magnitude and is
termed the intermediate in situ stress and which is aligned in the general
direction NW-SE
The minimum horizontal stress h, which has the lowest magnitude and is
thus known as the minimum in situ stress and which is aligned at right
angles to H (Figures 3-15 and 3-16).

3-131

Figure 3-11: Illustrates the Poissons ratio measurement


of solid body.

Figure 3-12: Acoustic time youngs modulus chart

3-132

Of course, local stresses may be modified by the presence of faults, or by the


intrusion of salt domes into shallower formations. Also, in some formations,
tendency for the rock to creep or flow over geological time periods (i.e salt or
soft mud rocks) can result in the horizontal stresses equilibrating towards the
magnitude of the overburden stress, such that all the stresses become equal and
are isotropic (i.e. pressure).
Also, as will be explained later, any reduction of fluid pressure throughout a
reservoir can lead to changes in the magnitudes of the horizontal in situ stresses
acting through the reservoir rock. Again, as most importantly, it should be
remembered that stresses acting in different directions act independently (i.e
they are tensor quantities) and as such they cannot be added together or
resolved to obtain a single equivalent vale (i.e. as one can do with vector
quantities).
3.1.6 OVERBURDEN STRESS
Overburden stress (v) is defined as the stress arising from the weight of rock
overlying the zone under consideration. In geologically relaxed areas having
little tectonic activity, the overburden gradient (=stress/depth) is taken as 1.0
psi/ft. In tectonically active areas, as in sedimentary basins which are still
undergoing compaction or in highly faulted areas, the overburden gradient
varies with depth, and an average value of 0.8 psi/ft is normally taken as being
representative of the overburden gradient.
In general, the overburden gradient varies from field to field and increases with
depth, owing to rock compaction. For a given field, accurate values of
overburden gradient can be obtained by averaging density logs from several
wells drilled in the area. The density-depth graph can be converted to an
overburden gradient-depth graph by the use of the relation:
Overburden stress = bulk density x depth
X acceleration due to gravity
The overburden vertical stress can also be represented by the following the
equation and it can be calculated from an integration of the density log:
v = g 0D bdD

3-20

If an average formation density is used the overburden stress is calculated from


the following equation:
v = 0.433 b x D

3-133

3-21

Figure 3-14: Stress components and preferred


plane of fracture

Figure 3-16: The direction of the stresses on a rock body

3-134

Where
v
b
g
D

= Vertical compressive tress, psi


= average rock density read from density log, gm/cc
= acceleration due to gravity
= depth, ft

Rock densities vary from 125 to 200 lb/ft3; 144 lb/ft3 is a reasonable average
and is the basis for the rule of thumb that the total vertical stress due to the
overburden is 1.0 psi/ft.
3.1.7 FORMATION PORE PRESSURE
Formation pore pressure is defined as the pressure exerted by the formation
fluids on the walls of the rock pores. The pore pressure supports part of the
weight of the overburden, while the other part is supported by the grains of the
rock. The terms pore pressure; formation pressure and fluid pressure are
synonymous, referring to formation pore pressure). Formations are classified
according to the magnitude of their pore pressure gradients. In general, two
types of formation pressure are recognized:
(1)

Normal pore pressure (or hydropressure). A formation is said to be


normally pressured when its pore pressure is equal to the hydrostatic
pressure of a full column of formation water. Normal pore pressure is
usually of the order of 0.433 0.465 psi/ft depending on the water
salinity.

(2)

Abnormal formation pressure (or geopressure). This type exists in zones


which are not in direct communication with its adjacent strata. The
boundaries of the abnormally pressured zone are impermeable,
preventing fluid communication and making the trapped fluid support a
larger proportion of the overburden stress.

The maximum value of abnormal formation pressure is 1psi/ft. for tectonically


relaxed areas and 0.8 psi/ft. for active areas. Exceptions to theses values were
found in certain parts of Iran and Russia in which the abnormal formation
pressure is in excess of the overburden gradient.
Formation pressures (normal and abnormal) can be detected by geophysical
and logging methods. Geophysical methods provide prediction of formation
pressure before the well is drilled, while logging methods provide information
after the well or section of well has been drilled. Logging tools are run on a
wire line inside the well. They include electrical, sonic, neutron, bulk density
and lithology logs. In porous formations, the overburden stress, v, is supported
jointly by the rock matrix stress g and the formation pore pressure, Pp. Thus,
v = g + Pp
3-135

3-22

Where
g = matrix stress or grain to grain stress
Pp = formation pore pressure

3.1.8 HORIZONTAL COMPRESSIVE STRESSES IN THE ROCK


The horizontal compressive stresses x and y, which are induced when the
vertical compressive stress (overburden) z is applied, try to prevent the lateral
expansion of the rock (Figure 3-18), therefore, the horizontal strain xh and y
equal to zero. According to Hooke's law, the horizontal strain in the absence of
pore pressure is expressed as:
y
x

z
x =
3-23
E
E
E

x z
E
E
E
y
z

x
z =
E
E
E

y =

3-24
3-25

Where
E

= young's modulus, psi


= Poissons ratio

For rocks in compression, x = y, is essentially zero and since the lateral


compressive stress x equals the lateral effective stress y, In a basin not
subjected to tectonic deformation, the horizontal stress components, within a
specific lithology, will be equal in every direction. The above equations can be
solved simultaneously to incorporate the total compressive stress component
as:

x = y = h =

1 z

3-26

Where
h

= Principal horizontal matrix stress, psi

It is assumed that in a tectonically inactive area and where rocks act as elastic
materials, horizontal matrix stress is about one-third to one-half the vertical
stress depending on the Poissons ratio of the particular rock. With a value of
Poissons ratio for adjacent zones, it should be possible to predict which zone
would fracture preferentially. In soft shales or unconsolidated sands, horizontal
matrix stress should be relatively higher. Rigid materials, such as dolomite or
limestone should fracture at lower pressures.

3-136

In salt zones where Poissons ratio may be 0.5, horizontal matrix stress may be
equal to vertical matrix stress, thus high fracture pressures and perhaps
horizontal fractures should result. Over geologic time, relaxation or creep may
affect horizontal rock stress causing it to be greater than would be calculated
from elastic theory. This is particularly true in the cases of shales.
3.1.9 Pore Pressure and Effective Stress
Pore fluids in the reservoir rock play an important role because they support a
portion of the total applied stress. Hence, only a portion of the total stress,
namely the effective stress component, is carried by the rock matrix. In
addition, the presence of a freely moving fluid in a porous rock introduces a
time-dependent character to the mechanical response of a rock. The rock will
react differently, depending on whether the rate of loading is slow or fast
compared to the characteristic time that governs the process of pore-pressure
diffusion (itself governed by the rock deformation) (Detournay et al., 1986). In
other words, to rigorously take into account the effects of the presence of the
pore pressure, one needs to introduce and differentiate between drained and
undrained properties.
The vertical matrix compressive stress is reduced where the formation has
porosity and contains fluid. Part of the overburden load is supported by the
pressured fluid (Figure 3-17). Effective vertical compressive stress in the rock
matrix, :

v = v Pp

3-27

= 0.433b D Pp

3-28

or

= effective vertical compressive stress, psi

Pp = formation pore pressure, psi


b = formation rock bulk density read from density log, gm/cc
As this equation indicates, vertical matrix stress is influenced by pore pressure.
Matrix stress is increased by declining reservoir pressures. Abnormal pressures
reduce matrix stress thus measurement of shale density (or something related)
is a useful indicator of abnormal formation pressure zones in drilling operations
because the higher pore pressure reduces shale compaction.
Biot proposed a correlation factor for the pore pressure term in order to account
for the shalinesss of the rock because the cementation existing between the
grains prevented the full magnitude of the pore pressure from counteracting the
applied load and the above equation can be written as:

3-137

Figure 3-17: Effective stress on rock

3-138

~ = 0.433 D P

b
p

3-29

Where
= poroelastic constant

0.7 for in-situ conditions


1.0 for fracturing
The horizontal compressive stress equation represents the classical relationship
between normal stresses in the absence of formation pore pressure. In the
presence of formation pore pressure, the effective average horizontal
compressive stress in the rock matrix h is defined as:

h =

z Pp
1

3-30

Pore pressure also affects horizontal matrix stress, since, as was previously
shown, pore pressure affects vertical matrix stress and horizontal matrix stress
is a function of vertical matrix stress. In permeable rocks, pore pressure is a
dynamic parameter influenced by production, injection or leakoff from the
fracture during a fracturing job.
The above equation can be written to include the Biot constant to represent

the minimum horizontal stress h (min) to become Eq. 3.34. To account for

the anisotropically of the stressed formations, the poroelasticity () should be


included in the equation below, substituting the value of the effective
horizontal stress, the equation becomes:


h min =
v Pp + Pp
1

3-31

Where
v

= Vertical compressive stress


= Constant 0.751.0

The above equation is known as the Terzaghi fracture model and it is used to
calculate the minimum horizontal stress in the rock.
Pore pressure also affects horizontal matrix stress, since, as was previously
shown, pore pressure affects vertical matrix stress and horizontal matrix stress
is a function of vertical matrix stress. In permeable rocks, pore pressure is a

3-139

dynamic parameter influenced by production, injection or leakoff from the


fracture during a fracturing job.
The tectonic forces resulting from large crustal movements introduce an
additional directional component which can be vectorially added to the stress
components already described (Figure 3-19). The influence of such tectonic
forces leads to a condition where the two horizontal stress components are
unequal. This tectonic regime also contributes to the sharp contrast in stresses
experienced between adjacent lithologies. If the above case exists, the
maximum horizontal stress , h min , can be calculated from the following:

h max = h min + tect

3-32

In a tectonically inactive area and where rocks act as elastic materials,


horizontal matrix stress is about one-third to one-half the vertical stress
depending on the Poissons ratio of the particular rock. With a value of
Poissons ratio for adjacent zones, it should be possible to predict which zone
would fracture preferentially. In soft shales or unconsolidated sands, horizontal
matrix stress should be relatively higher. Rigid materials, such as dolomite or
limestone should fracture at lower pressures.
3.1.10 CONCEPT OF ROCK FAILURE:
3.1.11 FRACTURE INITIATION OR THE BREAKDOWN PRESSURE
Fracture Initiation
The formation breakdown pressure is the pressure required to overcome the
wellbore stresses in order to fracture the formation in the immediate vicinity of
the wellbore. In oil well drilling, the fracture gradient may be defined as the
minimum total in situ stress divided by the depth. Knowledge of fracture
gradient is essential to the selection of proper casing seats, for the prevention of
lost circulation and to the planning of hydraulic fracturing for the purpose of
increasing the well productivity in zone of low permeability.
A hydraulic fracture treatment is accomplished by pumping a suitable fluid into
the formation at a rate faster than the fluid can leak off into the rock. Fluid
pressure (or stress) is built up sufficient to overcome the earth compressive
stress holding the rock material together. The rock then parts or fractures along
a plane perpendicular to the maximum compressive stress in the formation
matrix.
The occurrence of breakdown is often seen at the surface as a pressure peak.
Once the pressure peak is surmounted, fluid can be injected into the formation
3-140

at lower pressures. Breakdown effect is caused by the concentration of


compressive stress in the formation close to the borehole. This stress
concentration results when a portion of the rock is removed (by drilling the
hole) while the regional rock matrix load is unchanged; thus the rock at the
borehole accepts greater compressive stress.
To initiate a fracture sufficient hydraulic pressure must be applied to overcome
this increased stress level at the wellbore. Use of a penetrating fluid, wherein
fluid pressure tends to support some of the regional rock matrix load, reduces
the required fluid pressure to initiate breakdown.
As applied to sedimentary formation, the word fracture is sometimes thought to
be an irreparable occurrence somewhat the same as breaking a piece of glass.
This is not true. In creating a fracture, the formation matrix stress is
temporarily overcome using fluid pressure.
As soon as the fluid pressure is relaxed, the fracture closes back with little if
any increase in conductivity along the fracture, unless propped open by sand, or
in the case of a high pressure squeeze job, by cement filter cake.
Accurate knowledge of the fracture gradient is of paramount importance in
areas where selective production and injection is practiced. In such areas, the
adjacent reservoirs consist of several sequences of dense and porous zones such
that, if a fracture is initiated (during drilling or stimulation), it can propagate,
establishing communication between hydrocarbon reservoirs and can extend
down to a water-bearing zone.
The fracture gradient is dependent upon several factors, including type of rock,
degree of anisotropy, formation pore pressure, magnitude of overburden and
degree of tectonics within the area. It follows that any analytical prediction
method will have to incorporate all of the above factors in order to yield
realistic values of the fracture gradient.
HORIZONTAL FRACTURE
Assuming vertical components of force are exerted against the formation, the
condition necessary for horizontal fracture initiation is that the wellbore
pressure must exceed the vertical stress plus the vertical tensile strength of the
rock:

Pb = + S v + Pp

3-33

Where
(Pb
Sv
PP

= borehole pressure required to initiate horizontal fracture


= vertical tensile strength of rock (minimal)
= formation pore pressure

3-141

v = 0.433 b D PP

3-34

Where
b

= Density of formation rock, gm/cc

VERTICAL FRACTURE
Hubbert and Willis method
The Hubbert and Willis method is based on the premise that fracturing occurs
when the applied fluid pressure exceeds the sum of the minimum effective
stress and formation pressure. The fracture plane is assumed to be always
perpendicular to the minimum principal stress. In the case of a non-penetrating
fluid in openhole, Hubbert & Willis presented equations to calculate fracture
initiation pressure.

Pb = 3 h min h max + S h Pp

3-35

. Haimson and Fairhurst


Conditions for vertical fracture initiation depend on the relative strength of the
two principal horizontal compressive stresses. To cause formation breakdown,
the pressure in the borehole must be somewhat greater than the minimum stress
at the borehole, and must also overcome the tensile strength of the rock.
Consider an uncased vertical wellbore (or an open hole) under the action of

horizontal in-situ stresses min and max . The hydraulically induced fracture
is a vertical fracture and the fracture plane is perpendicular to the minimum
horizontal in-situ stress min as shown. Note that the above equation is
independent of the hole size and the elastic moduli of the rock medium. By
applying the theory of elasticity, and assuming the rock is an elastic medium
and has a tensile failure stress sh, expression for the fracture initiation pressure
for non-permeable formations where no fracturing fluid invades or penetrates
the formation can be expressed as:
Pb = 3h min h max + Sh + PP

3-36

Where
Sh = horizontal tensile strength of the formation
The rock has a tensile failure stress on the order of 500 to 1500 psi. Equation ()
clearly shows that the rock tensile failure stress T has a small effect on the
magnitude of breakdown pressure and the hole breakdown pressure is mainly
to overcome the compressive circumferential hole stress produced by the insitu stresses.
3-142

Haimson and Fairhurst, Medlin and Masse, Schmidt and Zoback modified the
above fracture initiation equation to include the formation pore pressure and
the poroelastic constant , for a formation impermeable to fracturing fluid as:
Pb = 3h min h max + Sh + PP

3-37

Substituting the following equation into the above equation:


h =
v Pp
1

3-38

Assuming horizontal stresses are equal (isotropic), h min = h max , the above
equation can also be written for a formation permeable to fracture fluid. The
breakdown fracturing pressure equation becomes:

2
Pb =
v Pp + S h + Pp
1

3-39

The above equation is only valid when no fluid invades or penetrates the
formation. In porous and permeable rocks the drilling mud normally penetrates
the formation, thereby changing the magnitude of the stress concentrations
around the borehole. The effect of fluid penetration is to create a force radially
outward which reduces the stress concentrations at the walls of the hole,
thereby making it more easy to fracture. Haimson and Fairhurst modified the
above equation to take into account the effects of fluid penetration, to obtain
formation breakdown pressure

Pb =

Pb =

3 3 2 + To
+ Pp
1 2v
2

1 v

3-40

2 3 + To
+ Pp
1 2v
2

1 v

3-41

Substitution of the h min equation into the above equation, the initiation
fracturing pressure equation becomes:

3-143


2
v Pp + To
1

Pb =
+ Pp
1 2
2

3-42

Where
Pb

h max

=
=
=

borehole pressure required to initiate vertical


fracture
vertical compressive stress
maximum principal horizontal matrix stress
minimum principal horizontal matrix stress

To
Pp

=
=

horizontal tensile strength of rock (minimal)


formation pore pressure

h min

3.2 FRACTURE EXTENSION


As injection of frac fluid continues, the fracture tends to grow in width as fluid
pressure in the fracture, exerted on the fracture face, works against the
elasticity of the rock material. After gradient frac fluid pad has been injected
to open the fracture wide enough to accept proppant, sand (the most common
proppant) is added to the frac fluid and is carried into the fracture to hold it
open after the job.
A vertical fracture grown in length upward, downward and outward. The
growth upward or downward may be stopped by a barrier formation;
downward growth may also be stopped by fallout of sand to the bottom of the
fracture. The growth outward away from the wellbore, (as well as upward or
downward) will be stopped when the rate of frac fluid leakoff through the face
of the fracture into the formation equals the rate of fluid injection into the
fracture at the wellbore.
Eaton method
The Eaton method is the most widely used in the oil industry. It is basically a
modification of the Hubbert and Willis method, in which both overburden
stress and Poissons ratio are assumed to be variable. Poissons ratio is a rock
property which describes the effect produced in one direction as stress is
applied in a perpendicular direction. Thus, for a two-dimensional case, if y is a
stress applied in the y direction and x is the resulting stress in the x direction,
then Poissons ratio. Most rocks tested under laboratory conditions produce a
Poissons ration of 0.25-0.3. Under field conditions, however, the rock is
subjected to a much greater degree of confinement and Poissons ratio can vary
from 0.25 (or less) to a maximum value of 0.5. Fracture reopening pressure or

3-144

Figure 3-20: Plot of leak-off test data

3-145

extension pressure is calculated form the following minimum horizontal matrix


stress equation (Eaton's equation).

FG =

v pf
D

Pp

+
1

3-43

Where
FG = Fracture pressure gradient
Pp = Formation pore pressure
3.2.1 FIELD DETERMINATION OF FRACTURE GRADIENT
Two methods are used for determining fracture gradient: direct and indirect.
The direct method relies on determining the pressure required to fracture the
rock and the pressure required to propagate the resulting fracture. The indirect
method uses stress analysis to predict fracture gradient.
It is a common practice to pressure test each new casing seat in field
applications to determine the exact minimum fracture gradient. The primary
reason for this practice is due to the inability of any theoretical procedure to
account for al possible formation characteristics. As an example, several
authors have noted wells that exhibited lower than expected fracture gradients
due to abnormally low bulk densities in the rock.
The most common procedure used for the field determination of fracture
gradients is the leakoff test (often called the pressure integrity test). The
procedure used in the test is to close the blowout preventers and then gradually
apply pressure to the shut-in system until the formation initially accepts fluid.
These results of the test would be similar to those shown in Figure 2-21. The
following example illustrates the procedure.
The direct method uses mud to pressurize the well until the formation fractures.
The value of the surface pressure at fracture is noted and is added to the
hydrostatic pressure of mud inside the hole to determine the total pressure
required to fracture the formation. This pressure is described as the formation
breakdown pressure. The test can be made in the open hole section below
surface or intermediate casings. The hole is first filled with fresh mud and the
annular preventors are closed. A pumping unit having accurate pressure gauges
is used to pump small increments of mud, - bbl. After each increment, the
shut-in pressure is observed and plotted against time or volume of mud pumped
in. Figure 2.20 gives a simplified schematic drawing of pressure plotted against
time during a fracturing test.
The pressure at the top of the curve is described as the formation breakdown
pressure. Continued pressurization will then merely extend the fractures created
3-146

by the breakdown pressure. The pressure required to extend such fractures is


described as the fracture propagation pressure; its magnitude is much lower
than the breakdown pressure. The fracture propagation pressure is normally
taken to be equal to the minimum principal stress, h. on the assumption that a
fracture outside the vicinity of the hole can be made when the value of h is
exceeded. The last portion of the curve represents the instantaneous shut-in
pressure after the pump is stopped. This pressure is known as the fracture
closure pressure, being the pressure required to keep the fracture from just
closing.
Example
Casing was set at 10,000 ft. in a well. The operator wishes to perform a leakoff
test to determine the fracture gradient at 10,000 ft. If the mud weight was 11.2
ppg, what is the fracture gradient at the casing seat?
Solution:
1.
2.

Close the blowout preventers and rig up a low volume output pump.
Apply pressure to the well and record the results as follows:
Volume pumped, bbl
0
1
1
2
2
3
3
4
4
5
5
6

Pressure, psi
0
45
125
230
350
470
590
710
830
950
990
1010

3-147

Figure 3-21: Typical results from a leakoff tst

3-148

3.
4.

The results are plotted in Figure ( ).


From these results, it appears that the formation will begin to fracture
when 950 psi is applied.
Fracture gradient
=
[(11.2ppg) (0.052) (10,000 ft) + 950 psi]/
10,000 ft.
=
6,774 psi/10,000 ft = 0.6774 psi/ft
=
13.02 ppg equivalent

FRACTURING TREATMENT PROCESS


After fracturing initiation, continual pumping would result in the fracture
propagating in a plane parallel to the maximum stress and perpendicular to the
minimum stress. Fracture extension pressure is lower than the reopening
pressure, but must exceed the minimum horizontal stress. It is a function of the
minimum horizontal stress, pump rate, hydraulic fluid characteristics, leak off
due to micro fissures, and matrix permeability.
The hydraulically induced fracture propagates from the wellbore into the
reservoir as pumping continues. A typically downhole pressure record (i.e., the
pressure measured inside the hole near the opening of hydraulic fracture) is
sketched in Figure 3-32. It is clear that the applied wellbore pressure first
balances the reservoir pressure (or pore pressure), then overcomes the
compressive circumferential hole stress, causing a tensile stress on the hole
surface. A fracture is initiated when this surface stress reaches the tensile
failure stress of the rock medium. The hydraulically induced fracture
propagates in the reservoir as pumping continues, and at the same time the fracfluid leaks off from the fracture surface into the surrounding rock medium. It
is important to observe that the opening of the fracture is maintained by the
differential between the net pressure (fluid pressure minus reservoir pressure)
and the minimum in-situ stress, while the rate of fluid leak-off from the fracture
surface is caused by the net pressure alone.
Referring to Figure 3-32 again, the maximum pressure is the initial breakdown
pressure, pb. The pressure drops, but not always in the field, when a fracture is
initiated at the hole surface. The near constant portion of the pressure curve is
the propagation pressure pprog. This is the pressure that causes the propagation
of hydraulic fracture into the reservoir. When pumping stops, the pressure
drops suddenly to a lower value but continues to decrease slowly to the
reservoir pressure due to fluid leaking off from the fracture as shown in the
figure. The transition point is called the shut-in pressure, psi (or the
instantaneous shut-in pressure, ISIP). At this point, the fluid flow inside the
fracture has ceased, and there is no friction loss due to fluid flow inside the
fracture. However, fluid continues to leakoff from fracture surface and the
fracture opening width continues to decrease. The fluid pressure inside the

3-149

fracture eventually reaches to equilibrium with the minimum in-situ stress min
and at this point the hydraulic fracture closes.
The fracture closure pressure, which can be determined from the pressure
decline analysis, is taken as a measure of the minimum in-situ stress. Although
the instantaneous shut-in pressure (ISIP) is somewhat higher then the fracture
closure pressure, the ISIP can be easily identified from the measured pressuretime curve. The ISIP is often used to estimate the magnitude of the minimum
horizontal in-situ stress by field engineers. Unfortunately, the situation is
somewhat more complicated in field conditions. The underlying control factors
for this pressure drip are discussed by McLennan and Roegier.
Most wellbores that need fracturing are cased wellbores. To fracture a cased
wellbore, the wellbore is first perforated with shaped charges to form a series
of perforated holes spiralling along the wellbore surface.
The perforations are typically made at spacings of 4 to 6 inches and at a phase
angle of 60 to 120 degrees as shown in the figure. When the wellbore is
pressurized, the perforated holes in (or near) the direction of maximum
horizontal in-situ stress ( max ) will be fractured first. Figure 3-32 illustrates the
fracture operational sequences during the fracturing treatment.

Figure 3-32: illustrates the field operation during a fracture


treatment
3-150

Sand-Out May Stop Treatment Prematurely


The width of the fracture is related to the new fracture pressure (pressure in
excess of fracture closure pressure) working against the elasticity of the
formation. As sand enters the fracture and is deposited, more fluid pressure is
required to create greater stress against the fracture face to increase the frac
width. If the required fluid pressure cannot be applied due to equipment or
casing limitations, fluid injection rate slows, sand drops out of the fluid at a
more rapid rate, and a sand-out in the fracture occurs.
Providing sufficient fracture length and width has been generated, a sand-out
within the fracture is desirable from the standpoint of well productivity. A
sand-out in the casing can occur due to the fact that insufficient fracture width
has been generated to accept the size sand carried in the fluid or due to drop
out of sand inside the casing closing off the perforated section. Sand-out in the
casing usually occurs early in the treatment and is obviously undesirable.
3.2.2 HYDRAULIC FRACTURE GEOMETRY ANALYSIS
Hydraulic stimulation has proven to be a dominant factor in the success of
marginal wells in low-permeability, low-porosity, dense rocks. Twenty percent
or more of the total well cost can be involved in fracturing; proper treatment
design is a must if low-production wells are ever to reach payout. The
treatment design is critical. Too small a fracture treatment may result in such
inadequate drainage of the reservoir that the well remains unprofitable.
Conversely, too large a treatment can be an unnecessary waste of completion
funds and render the well unprofitable; worse, the fracture may migrate into a
nearby aquifer.
Design and Implementation
A cost effective fracture stimulation treatment must consider: reservoir rock
deliverability, fracture mechanics of the productive zone and the boundary
zones, frac fluid characteristics, proppant and proppant transport, operational
constraints, the surface and downhole well flow system, and the resulting frac
job cost and producing economics. Frac job design and implementation
involves many compromises and must deal with many factors that are at best
approximations. This not withstanding, tremendous progress has been made in
the last twenty years in quantifying design and implementation procedures.
3.2.3 FRACTURE GEOMETRY MODELS:
The prediction of fracture geometry is one of the central issues in the
engineering design of the stimulation treatment. Over the years, various models
have been developed to determine the relationship between injection rate, fluid

3-151

leak off behavior, fracture dimensions (width and length), and total volume of
fluid pumped into the formation.
3.2.4 FRACTURE HEIGHT
When pressure is increased in the borehole, rupture occurs in the plane that is
perpendicular to the direction of least compressive stress (x or y). The
pressure required to induce this fracture is called the initial or breakdown
pressure. Once a fracture has been initiated, the pressure necessary to hold the
fracture open (in the case of a vertical fracture) will be equal to the minimum
total horizontal stress. This stress is often referred to as closure stress. In
tectonically relaxed areas, the least principal stress is generally horizontal.
Fracturing should therefore occur along vertical planes.
Hydraulic fracture design depends on two sets of variables: the distribution and
magnitude of in-situ minimum horizontal stress in the producing and
surrounding formations, and the flow behavior of the fracturing fluid. These
variables determine:
The direction and geometry (height, length and width) of the created
fracture,
Whether multiple zones should be fractured one at a time, in groups or
simultaneously,
Design parameters of hydraulic fracturing, such as horsepower, pumping
pressure and proppant transport, and
The fracturing fluid flow behavior and efficiency.
3.2.5 HYDRAULIC FRACTURE LENGTH MODELS
ONE DIMENSIONAL FRACTURE LENGTH MODEL
In 1957, Howard and Fast presented a mathematical formula for determining
the fracture area of a newly opened fracture on the basis of treating conditions.
During the fracturing process, the fracture fluid is injected at the wellhead at a
constant injection rate, qi. In the fracture this injection rate is split up into two
components as shown in figure 2-21. Part of the liquid, ql enters the formation
as a result of the differential pressure (Pf - Pe) between the fracture and the
external boundary and the remainder, qf, increase the fracture area, i.e., it
increases the volume for the fracture. An expression for the fracture area at any
time may be derived by using this basic concept and the following
assumptions:
1. The fracture is of uniform width.
2. The flow of fracturing fluid into the formation is linear and the direction
of flow is perpendicular to the fracture face.

3-152

3. The velocity of flow into the formation at any pint on the fracture face is
a function of the time of exposure of the point to flow.
4. The velocity function v = f (t) is the same for every point in the
formation, but the zero time for any point is defined as the instant that
fracturing fluid first reaches it.
5. The pressure in the fracture is equal to the sand face injection pressure,
which is constant.
The basis of fracture geometry prediction is a material balance expression:

Vi = Vf + Vl

3-44

Where
Vi = the total fluid volume injected
Vf = the fracture volume
Vl = the fluid loss volume
The above equation can be written in terms the flow rate as:

qi = q l + qf

3-45

Where
qi = constant injection rate during extension ft/min
ql = volume rate of fluid loss (leak off) to the formation, ft/min
qf = volume rate of fracturing fluid remaining inside the fracture,
ft/min
The rate at which the fracture fluid leaks from the two fracture faces into the
formation is related to the velocity, v, and the fracture area A of one face, by:

q l ( t ) = 2 0A ( t ) v ( t ) dA

3-46

Where
qi = volume rate of fluid loss to the formation.
v = velocity of flow perpendicular to the fracture plane
A = area of the fracture face.
Since the extent of the fracture increases with time, the fracture area is a
function of time. For a given element of area dA formed at time , the velocity
of flow into the element is v (t - ), correcting for the zero time in the fourth
assumption. Since A is also a function of time,

dA
dA =
d
d
3-153

3-47

The above leak off equation can therefore be written as:


t
dA
q L ( t ) = 2 v ( t )
d
0
d

3-48

The rate of increase of fracture volume is:

dA
dt

qf = W

3-49

Where
W = fracture width
By substitution, the above equation becomes:

dA
dA
t
q i = 2 v(t )
d + W

0
dt
d

Where

3-50

= Constant injection rate during extension, ft3/min.


= Velocity of fluid flowing into the formation at a given
point, ft/min
t
= Total pumping time, min.

= Time required for the fluid to reach a given point, min.


(t - ) = Time interval during which the fluid has leaked from
any points, min.
A
=
Total area of one face of the fracture at any time during
injection, ft2.
W
= Average fracture width, ft.

qi
v(t-)

The value of v as a function of time will be determined for each of the fluid
loss mechanisms. In all the cases the value of v (t) is:
v( t ) =

3-51

Or
v( t ) =

C
t

3-52

Where
C = Leak off coefficient, which is different for each mechanism.

3-154

Substitution of the above equation into Eq. 3.52 yields:


t c dA
dA
q i = 2
d + W

dt
o t 8 d

3-53

The above equation is the mass conservation equation representing the linear
propagation on a vertical fracture for the case of no spurt loss. The above
equation may be solved for the fracture area at any time A (t) by means of the
Laplace transformation, provided that qi is a constant and the equation for v (t)
is known. In 1957, Carter derived the basic solution of the above equation for
estimating the extent of the fracture area taking into account the effect of fluid
leaking into the formation and substituting the leak off velocity equation into
the general mass balance equation. Carter equation indicating the generated
fracture area during a fracturing treatment as a function of fracture length, a
constant fracture width. Normally, the solution to the above equation will be
handled by computer programs, although it can be solved by graphical
solutions.
The following equation which is Carters solution to the above mass
conservation equation is used to determine the created fracture area, assuming
that the fracture width is constant and there is no spurt loss.

2C t

qi w f w
A(t ) =
(e
4 C2

.erfc 2C t
w

4C t

+
1)

3-54

or
A(t ) =

e x .erfc( x ) + 2 x 1

4 C 2

qi w f

3-55

Where
x=

2C
wf

3-56

Where
A(t)
qi
t
wf

= total area of one face of the fracture, ft


= constant injection rate, fts/min
= total pumping time, min
= constant fracture width, ft
3-155

C
= fluid-loss coefficient, ft/min1/2
erfc(x) = complementary error function of x
The total area of the created one face of the fracture according to the Carter
model is:
A = 2L h f

3-57

Where
L = fracture half length (wing length)
hf = fracture height
And the volume of the two-sided fracture is:
V = Wf A

3-58

Fractures normally have two wings, one on either side of the well, which are
assumed to be roughly equal in length and geometry (the length of one wing is
termed L) (Figure 3-22). The pad volume does the work of opening up the
fracture tip and creating fracture width; and of sealing the fracture face to
control fluid loss. Figure 2.22 illustrates the fracture shape with the pad volume
inside the fracture. The fracture wing length L as a function of time in Carters
model is then obtained by dividing the area by twice the fracture height.
L=

x2
e . erfc( x ) + 2x 1

8 h f C2

qi w f

3-59

TWO DIMENSIONAL FRACTURE LENGTH MODEL


Geertsma and de Klerk derived a vertical fracture propagation model including
the effect of fluid leakoff which is similar to Carters fracture propagation
model except that the fracture volume differs from Carters because of the
elliptic shape in the horizontal cross sections. Fluid loss is incorporated by
assuming that it has no effect on fracture shape or pressure distribution.
And the above mass conservation equation is rewritten to include the effect of
spurt loss for the linear propagation of a vertical fracture as:
VI = Vf + VL + VSP

3-60

Performing a derivation procedure similar to that previously presented by


Howard and Fast, the governing mass conservation equation including the
effect of spurt loss is given by:

3-156

dA
f =q
t v (t ) dA d V
4

sp
i
0
dt
d
dt

dV

3-61

The fracture volume, Vf , and fracture area, A for a two-sided two- wing
symmetric elliptical fracture shape, at any time for the linear propagating KGD
or PKN -type fracture model are defined respectively as follows:
Vf =

h f Lw w
2

3-62

A f = 4Lh f

3-63

Substituting the above fracture area and volume equations and the above
leakoff velocity function equation into the mass conservation equation leads to

C
d dL

dL
w w + 4VSP
+ 4 t

h 0 t d
2
h dt

qi =

3-64

Performing a volume balance and solution procedure similar to that of Carter


using the La Place transformation, and including the effect of spurt loss VSP,
they obtained the following equation, which is written in field units, for the
determination of the one sided fracture wing length, L, including the effect of
fluid leak off and spurt loss for the KGD or the PKN models:

L=

w + 8 V 2 + e 2 erfc 1

sp
w

32 h f C 2

qi

3-65

And

8C

3-66

ww + 8V sp

The above equation can be written in field units to become:

L=

q i 5.615

V
+
w
8
w
sp
2
32 h C 12
f

+ e erfc

3-157

3-67

Figure 3-21: Fracture dimensions in Carter and Howard


and Fast model

Figure 3-22: Showing fracture dimensions including pad volume

3-158

And

8C

w w + 8Vsp
12

3-68

Where
qi =
hf =
G =
=
=
ww =
Vsp =
C =

Injection rate, bbl/min


Fracture height, ft
Shear modulus, psi
Poissons ratio
Injection fluid viscosity, cp
Fracture width at the wellbore, in
Spurt loss, ft3/ft2
Fracturing fluid coefficient, ft/min1/2

If an average fracture width, wf, is assumed to be constant and if SP = 0, the


solution to the above equations would be Carter's (1957) solution. In order to
solve for the half fracture length, L, and fracture width, w, taking into
consideration the effect of spurt loss and realizing large fracture heights, an
iterative technique between fracture width and length equations should be
performed until a constant solution was found and then the net fracture pressure
also found.
3.2.6 FRACTURING FLUIDS:
The fluid used in fracturing jobs is there to create fracture volume, width,
length and height and transport the propping agent into the fracture. An
important factor in creating a fracture is the leak off rate into the formation. If
the leak off rate is high, then little fluid remains within the fracture to
propagate it even further into the formation. The ability of the fracturing fluid
to carry the proppant is dependent on the fluid viscosity, the more viscous the
fluid, the better its carrying capacity.
3.2.7 Types of fracturing fluids:
1.
2.
3.
4.

Oil based fluids:


Used for low injection rates, shallow to medium fracture depths.
Water based fluids:
Have high density and low friction loss in deeper wells, and have
high injection rates.
Emulsions:
Either oil in water or water in oil. Low fluid loss.
Foams and gases:
Used for very low permeable zones.
3-159

3.2.8 Fluid Loss Control:


There are three different leak-off mechanisms responsible for controlling the
amount of the fracturing fluid that leaks off into the formation during the
fracturing operation (figure3- 23).
1. Viscosity control
Depends on treatment fluid leak off viscosity.
2. Compressibility control
Depends on reservoir fluid viscosity and compressibility.
3. Wall building control
Depends on the filter cake building of the treatment fluid.
The rate of fluid leak-off from the fracturing fluid into the formation is a
function of the total pumping time and it is expressed in the following
equation:
v(t ) =

CT
t

3-69

Where
CT = fracturing fluid coefficient.
The fracturing fluid coefficient, Ceff, depends on, the characteristics of the
fracturing, fluid, the reservoir fluid and the type of formation rock. The
fracturing fluid coefficient is related to the properties of the fracturing fluid. A
low fracturing fluid coefficient means low fluid loss properties and thus larger
fracturing area for a given fluid volume and injection rate. The fracturing fluid
which remains inside the fracture is necessary for the extension of the fracture.
Therefore, it is important to control the properties of the fracturing fluid in
order to have minimum fluid escape into the formation, ql.
Viscosity Control Coefficient:
The viscosity of the fracturing fluid is much greater than that of the formation
fluids. Therefore, the fluid loss into the formation is controlled by the viscosity
of the fracturing fluid. The fracturing fluid loss viscosity coefficient, c , is
described in the following eq.

3-160

KP
C v = 0.0469
frac

1/2

3-70

The differential pressure in the fracture face is calculated from the following
equation:
P = G f D Pr

3-71

Where
K
= relative permeability of formation to leakoff effluent, darcies

= formation porosity, fraction


frac = fracturing fluid viscosity, cp
P = differential pressure across the face of the fracture, psi
Gf = fracture gradient, psi/ft
Pr = reservoir pressure.
Compressibility Control Coefficient:
The rate of fluid loss into the formation is controlled by the viscosity and the
compressibility of the reservoir fluids if the fracturing fluid used is lease oil or
water with no fluid loss additives. This is because the physical properties of the
fracturing fluid are identical or nearly identical with those of the reservoir
fluids.
During fracturing, the formation to be fractured is completely saturated with
liquid, additional volumes of fracturing fluid should be injected in order for the
fracturing operation to proceed. Since the formation is completely saturated
with liquid any additional volumes of fracturing fluid cannot be taken by the
formation unless the fluids in the reservoir are compressed. The equation
describing the rate of fluid loss into the formation in this case is the diffusivity
equation in linear form. The reservoir fluids compressibility control coefficient
is given in the following eq.
12
kcf

Cc = 0.0374

res

3-72

Where
Cc
K

cf

= compressibility control coefficient


= formation permeability relative to mobile reservoir fluid,
Darcy's
= formation porosity, fraction
= total isothermal compressibility of the reservoir fluids, psi-1
3-161

res = viscosity of reservoir fluids, centipoises.


P = differential pressure across the face of the fracture, psi.
Wall Building Control Coefficient:
Fluid loss additives, such as asphaltic type material, synthetic gums and
insoluble solids are added to the fracturing fluids in order to form a temporary
filter cake on the walls of the fracture, which minimizes the amount of fluid
leaks into the formation. The filter cake forms a low permeability layer, which
controls the fluid loss from the fracturing fluid into the formation.
Because of this low permeability, most of the pressure drop will occur across
the filter cake, and therefore, the rate of flow of the fracturing fluid into the
formation is obtained by applying Darcy's law across the filter cake. The
fracturing fluid wall-building coefficient is described by the following
equation:
Cw =

Where
mact
Af

0.0164 mact
Af

3-73

= slope of water leak of loss curve, cc/ min1/2


= Area of the filter paper or core exposed to fluid, cm2

Correcting for the experimental pressure to simulate the reservoir pressure


1/2
Pact

mact = mlab

Plab

3-74

Where
mlab = slope of water loss curve measured in laboratory, cc/min
Figure 2-24 is obtained from the filtration experiment run on the fracturing
fluid in the laboratory. The figure illustrates the procedure for calculating m
(the slope the fluid loss curve in the laboratory). The spurt loss, Vsp , is
determined from the y-axis intercept b in the leak off static filtration curve (
figure 2-24).
VSP =

Where

24.4 b
A

3-75

Vsp= Fracturing fluid spurt loss, gal/100 ft2

3-162

All the three fracturing fluid leak off control mechanisms tending to retard the
fluid loss into the formation. The three mechanisms operate simultaneously in
controlling the fluid loss from the fracturing fluid into the formation.

1
CT

1
Cv

1
Cc

1
Cw

3-76

3.2.9 HYDRAULIC FRACTURING WIDTH MODELS


Among these models are three fundamentally different theories, KGD
(Khristianovitch and Zheltov, 1955; Geertsma and de Klerk, 1969) and PKN
(Perkins and Kern, 1961; Nordgren 1972), which are widely used to predict the
geometry of a constant-height vertical fracture with very different results.
3.2.10 TWO-DIMENSIONAL FRACTURE WIDTH MODELS
- Assumption of constant height, or penny-shaped circular fracture
geometry
3.2.11 1) PKN fracture model
PKN model is used to simulate the propagation of a vertical hydraulic fracture.
It is used for length height ratio of less than 1, this model assumes a plane
strain in vertical sections and longer, narrower fractures exist. This model
includes the variation in the flow rate along the fracture. The primary
assumption of these models is that the fracture length is greater than the height.
Assuming that there is no flow in the vertical direction, the pressure in a
vertical cross section of the fracture is constant, and the fracture has an
elliptical shape. Also, there is no elastic coupling between the planes, i.e. the
state of stress at a point X does not depend on the pressure distribution at other
locations along the fracture length. Hence, the fracture width can be expressed
as a function of the local pressure. The only coupling between the different
vertical cross sections is due to the fluid flow in the fracture. Figure 3-25 is a
representation of the PKN model.
Perkins and Kern (1961) assumed that a fixed height vertical fracture is
propagated in a well confined pay zone; i.e. the stresses in the layers above and
below the pay zone are sufficiently large to prevent fracture growth out of the
pay zone. They further assumed that the fracture cross section is elliptical with
the maximum width at a cross section proportional to the net pressure at that
point and independent of the width at any other point (vertical plane strain).
Although Perkins and Kern developed their solution for non-Newtonian fluids
and included turbulent flow, it is assumed here that the fluid flow rate is
governed by the basic equation for flow of a Newtonian fluid in an elliptical
section (Lamb 1932).
3-163

Figure 3-23: Show the invaded zone of the leak-off fracturing fluid

Figure 32-24: Illustrates the fracturing fluid wall building


control in the laboratory

3-164

dp
64 qu
=
dx
h f w 3

Where

3-77

q = Injection rate, ft3/min


hf = Fracture height, ft
P = Pressure inside the fracture, psi
x =Distance along the fracture, ft
= Injection fluid viscosity, cp
w =Fracture width, in

And the shape of the fracture is elliptical so that the average width is

w=

ww
4

E' =

3-78

Where

Where

3-79

12

E =Plane strain modulus

Integrating the above equation along the fracture half-length L and writing it in
terms of field units and setting P net =0 at the fracture tip.
Pnet

q E3
= 0.015 i 4
h f

1/ 4

3-80

Where
qi = injection rate, bbl/min
= Fracturing fluid viscosity, cp
P = Fracturing pressure, psi
The fracture width equation for the PKN model in field units is written as
follows:
ww

q L
= 0.357 i
E

1/ 4

3-81

Hydraulic fracture designs for obtaining the fracture width and length were
performed by iterating between the Carter technique to obtain, L , fracture

3-165

length as a function of time and the Perkins and Kern model to determine the
width, wf , until a consistent solution was found, and then the net fracturing
pressure Pnet is determined.
3.2.12 2) KGD fracture model
KGD model is more appropriate when the fracture length is smaller than the
height. It is used for length height ratio of greater than 3. The flow rates inside
the fracture and into the adjacent formations can be illustrated from figure 326. Khristianovich, Zheltov, Geertsma and de Klerk derived an equation for
the determination of the width of vertical fractures with a rectangular cross
section and assuming the fracture has an elliptical shape also it assumes plane
strain in horizontal sections and shorter, wider fracture to exist.
dP
12q
=
dx
hf w3

3-82

Solving the above differential equation, the KGD model pressure and width
equations in field units respectively are
1/ 4

q E3

Pnet = 0.03 i
h f L2

3-83

And the fracture width equation in field units becomes:

ww

q L2
= 0.345 i
E h f

1/ 4

3-84

Hydraulic fracture designs for obtaining the fracture width and length were
performed by iterating between the Carter technique to obtain, L , fracture
length as a function of time and the Khristianovich, Zheltov, Geertsma and de
Klerk model to determine the width, wf , until a consistent solution was found,
and then the net fracturing pressure Pnet is determined.
The PKN and KGD models, both of which are applicable only to fully confined
fractures (constant fracture height). Each differ from the other in one major
assumption which is the way in which they convert a three-dimensional (3D)
solid and fracture mechanics problem into a two-dimensional (2D) (i.e. plane
strain). KGD model assumes plane strain in the horizontal direction i.e. all
horizontal cross sections, act independently or equivalently and all sections are
identical which is assumed that the fracture width changes much more slowly
vertically along the fracture face from an y point on the fracture face than it

3-166

Figure 3-25: PKN fracture model

Figure 3-26: KGD Fracture model

3-167

does horizontally. This means that the fracture growth vertically is smaller than
the growth horizontally.
In practice, this is true if the fracture height is much greater than the length.
PKN model, on the other hand, assumed that each vertical cross section acts
independently which is equivalent to assuming that the pressure at any section
is dominated by the height of the section rather than the length of the fracture.
This is true if the length of the fracture is much greater than the height. This
difference is one basic assumption of the models lead to different ways of
solving the problem and can also lead to different fracture geometry
predictions.
3.2.13 3-PSEUDO 3-DIMENSIONAL
This model is the extension of the 2D models to include height growth. It is
goods model for engineering design and evaluation. This model assumes height
growth along fracture length.
3.2.14 SIMULTANEOUS SOLUTION OF WIDTH AND AREA
FUNCTIONS:
In all of these solutions for determining fluid efficiency, volume of fluid in the
fracture, fracture area and / or fracture length, it was necessary to obtain an
average fracture width. As has been shown, this width can be determined by
either of two methods, Perkins and Kern (PKN) or Christianovich (KGD).The
equations presented by Perkins and Kern allows calculation of width as a
function of an assumed height, fluid viscosity, injection rate, fracture length
and Young's Modulus. Similarly the equations presented by Christianovich
(KGD) allow calculation of width as a function of an assumed height, fluid
viscosity, injection rate, fracture length and Young's Modulus.
The fracture length based on the Howard and Fast (Carter) equation which then
was modified by Christianovich (KGD) depends upon fracture width. The
fracture width equations presented by Perkins and Kern (PKN) or
Christianovich (KGD) also depends upon fracture length. This then requires a
simultaneous solution of both equations numerically or graphically. This
graphical solution is done by preparing a log-log graph using both sets of
equations. The graphical procedure technique is described below:
Based on a selected treatment volume, injection rate and fluid loss coefficient,
values of can be determined for various assumed fracture widths. The values
of along with the corresponding error function values obtained from the
complementary error function (Table 3-2), can then be used to calculate the
created fracture length and the fracture area and the term AQ using the
appropriate equations for each of the assumed fracture width conditions.

3-168

The area values can be calculated directly by substituting x or values in the


corresponding fracture length or area equations. These areas should then be
used to calculate a term called AQ, which is area (A) divided by the injection
rate (2qi)
AQ =

A
2q i

3-85

Based on these calculations a plot of log (AQ) versus log wf (assumed fracture
width) can be obtained as shown in figure 2-27, for a specific pumping volume
and, hence, for selected pumping time. By selecting several time values, a grid
of time lines may be constructed.
Using the Perkins and Kern (PKN) or Christianovich (KGD) fracture width
equations and assuming the same fracture height and injection rate, values for
fracture width can be determined for selected values of fracture length. From
these selected lengths, values for the fracture area and then this term AQ can be
calculated from the above equations. The fracture area for the PKN or KGD
models is calculated from the following equation:
A = 4 Lh f

3-86

The resulting relationship between width and AQ can be plotted on the same
graph as the Howard and Fast or Christianovich (KGD) values, this giving
solution to the two equations which is illustrated as the solid line on figure 327. The intersections of these curves represent solutions that satisfy both
equations. Thus, the intersection on figure 3-27 of the 20 minute curve with the
corresponding width curve represents the solution for the width (wf) and value
of AQ that would result if a fluid of that viscosity and Ceff were pumped for 20
minutes. The value of AQ can then be used to find the created fracture length
from the appropriate fracture length area equation using the assumed fracture
height. The intersection of the two curves will give a fracture width for the
original desired treatment volume.
It should be understood that both of these solution methods (1) Perkins and
Kern equations with the Howard and Fast and Carter equations and (2) KGD
equations with the Howard and Fast equations, are only approximate solutions
to make hand calculations possible. These equations for fracture width and
length are quite difficult to solve without the aid of a computer. However, the
solution can be made using the graphical solution method.
This results from the fact that, in both cases, fluid loss effect on rate and
volume during the development of the fracture has been ignored in calculation
of fracture width. In other words, both the methods for solving fracture width
have assumed conditions that exist only at the end of the treatment. The true
3-169

Table 3-2: Complementary error function

3-170

Figure 3-27: Solution of fracture width equation with fracture


area equation.

3-171

method of solution would modify the injection rate by the efficiency, at each
increment of time, and determine the width based on the effective injection
rate.
This type of solution is not practical for hand calculations and should be
handled by a computer. The simultaneous solution of both width and length
equation yields the required fracture length, L and the fracture width ww needed
for selected fracture treatment. The created fracture volume during the
treatment can be obtained from the following equation:

Vf =

h f Lw w
2

3-87

3.2.15 FRACTURING FLUID EFFICIENCY:


If we define the fluid efficiency of a fracturing treatment as the volume of the
fracture divided by the volume of fluid injected, then:

Eff =

Vf
Vi

A
qi t

Wf

3-88

Where:
Eff = fracturing fluid efficiency
Vf = volume of fracture
Vi = volume of fluid injected.
3.2.16 PROPPANTS:
Proppants are used in order to hold a fracture open after the fracturing job is
complete, and provide conductivity. After a fracturing treatment, the formation
closes in around the proppant decreasing the width of the fracture. The
important factors that are needed to be considered in the design of the
proppants are:
1. Proppant size
Will it fit the fracture.
2. Permeability
Will it be conductive enough.
3. Strength
Will it crush due to closure pressure.
3.2.17 Proppant Transport:
The purpose of propping agents in a hydraulic fracturing treatment is to hold
the fracture open and provide a permeability path for the fluid flow into the
3-172

wellbore. The improvement of well productivity depends on final propped


geometry and fracture conductivity. Therefore, it is important to design a
treatment schedule such that maximum proppant coverage can be obtained at
the end of pumping.
Proppants slurry concentration
In addition to the length, the propped width of the fracture describes the
fracture geometry that controls post-treatment production. The fracture
conductivity is simply the product of the propped width and the proppant pack
permeability. The width in that expression is the propped width of the fracture.
As should be obvious from the last two sections, the relationship between
hydraulic width and propped width is indirect; it depends greatly on the fluid
efficiency and especially on the possible end-of-job proppant concentration.
Assuming that a mass of proppant, Wp, has been injected into a fracture of halflength xf, and height hf and the proppant is uniformly distributed, then:
Wp =

Lh f w p (1 p )p x 62.4
2

3-89

Where the product 2 Lh f w p (1 p ) represents the volume of the proppant

pack and is characteristic of the proppant type and size. The density, p , is also
a characteristic property of the proppant. A frequently used quantity is the
proppant concentration in the fracture, CP , defined as:
Cp =

Wp

3-90

4Lh f

3.2.18 Proppant Selection:


Placing the appropriate amount and type of proppant in the fracture is critical to
the success of a hydraulic fracturing treatment. Proppant concentration and
strength will determine the fracture conductivity over the producing life of the
well. Selection of a proppant is mainly governed by the fracture conductivity
needed for a desired production rate.
Factors such as proppant properties (strength, particle size, roundness, fines
content), reservoir closure stress, drawdown rate, embodiment, and resultant
propped fractured width will have an effect on fracture conductivity. When a
hydraulic fracture is created, the in-situ stresses must be overcome to open and
propagate the fracture. After the well is put on production, these same stresses
tend to close the fracture and act on the proppant. This is often referred as
closure stress and is represented by the equation If the proppant is not strong
3-173

enough to withstand the closure stress of the fracture, it will be crushed and
will plug the proppant pack, and the permeability of the pack will be drastically
reduced.
PAD VOLUME:
Is the volume of fluid leak off into the formation through the faces of the
fracture. The pad volume and the proppant concentration determine the final
propped fracture penetration and conductivity. Insufficient pad volume often
results in premature screen out of propping agents caused by early pad
depletion.
3.2.19 FLUID VOLUME REQUIREMENTS
A fracture execution consists of certain distinct fluid stages, each extended to
perform a significant and specific task.
Pad is fracturing fluid that does not carry proppant. It is intended to initiate and
propagate the fracture. During the fracture propagation, fluid leakoff, into the
reservoir and normal to the created fracture area, is controlled primarily
through the buildup of a wall filter cake. The volume of fluid leaking off is
proportional to the square root of the residence time within the fracture.
Therefore, pad, being the first fluid injected, acts as sacrificial to the following
proppant-carrying slurry.
After the pad injection, proppant slurry is added to the fracturing fluid in
increasing slurry concentrations until at the end of the treatment the slurry
concentration reaches a predetermined value. This value depends on the
proppant-transporting abilities of the fluid and/or the capacity the reservoir and
the created fracture can accommodate. An approximation of the relationship
between total fluid volume requirements, Vi, and the volume that is pad, Vpad,
based on the fluid efficiency, EFF, was given by Nolte (1986) and Meng and
Brown (1987):

1 EFF
Vpad Vi

1 + EFF

3-91

The required total fracturing fluid for the treatment (Vi = qi t) is equal to the
pad volume plus the proppant slurry volume. Since the total portion of the total
fluid volume that is pad was determined from the above equation, the onset of
proppant addition can be obtained from the following equation:
t pad =

Vpad
qi

3-92

3-174

3.2.20 FRACTURE PENETRATION AND CONDUCTIVITY:


In order for fracturing treatment to be successful, the fracture should penetrate
deep into the formation and provide good conductivity. The fracture
penetration is dependent on the fracturing treating pressure and the type of fluid
used for the fracturing treatment. The fracture conductivity is obtained by
packing the fracture with a proppant, which forms high permeability when it is
left in the fracture space. Figure 3-28 illustrates the fracture dimensions and
proppant placement technique.
The proppant is used in order to hold the fracture open after the withdrawal of
the fracturing fluid from the fracture. In order to increase flow capacity to the
wellbore, it is necessary to maximize Wf, rf and Kf. The propped fracture
permeability is determined from figure 3-29. Complete sets of data for most
common proppants (proppant type, size, concentrations) can be found in Penny
(1986) and Penny (1988).
The fracture conductivity can be calculated from the following equation:
FCD =

K f Wf
Ke L

3-93

The fracture conductivity ratio is calculated from the following equation:


Conductivity Ratio =

Kf Wf
Ke

40
Aw

3-94

Where:
Kf = Permeability of packed fracture
Wf = Width of fracture
Ke = Formation permeability
Aw = Well spacing, 4re2
The proppant fracture permeability can be calculated from the closure stress
fracture permeability charts or from the following Blake-Kozeny equation:
Kf =

Where

Kf
dp
f

=
=
=

d 2p3f

150(1 f )

fracture permeability, cm2


diameter of the proppant, cm
porosity of the packed fracture, fraction

3-175

3-95

Figure 3-28: illustrates the fracture dimensions conductivity.

Figure 3-29: Illustrates the relationship between fracture


conductivity and fracture closure stress.

3-176

3.2.21 WELLHEAD TREATING PRESSURE:

w = f + perf + fric hydro

3-96

The density of the fracturing fluid including proppants is

t =

8.34 + C p

3-97

1 + 0.0456 C p

The actual fracturing fluid injection rate including proppants is

qi =

Vi +

W
( f x 8.34)
t

3-98

The pressure drop through perforations can be calculated from the following:

qi

perf = 0.237 T
Cp N d 2

3-99

Where
Pperf

qi
Cp
N
d

= pressure drop across perforations, psi


= fluid density, lb/gal
= Pump rate, bbl/min
= perforation discharge coefficient
= Number of perforations
= diameter of perforations, in

The hydrostatic pressure at the bottom of the well can be calculated from the
following equation:

(Ph )l

1 + 0.0453 x Pc
= Ph x

1 + ABV x Pc

Where
Ph = Hydrostatic pressure gradient
PC
= Proppant concentration, lb/gal
ABV = Absolute volume of the proppant gal/lb
3-177

3-100

Tables 3-3 through 3-5 list the physical properties for various proppants.
Hydrostatic pressure for various fluids may be calculated, using the hydrostatic
pressure gradient figures 3-30 and 3-31. For proppants other than sand, the
above equation can be used to correct the hydrostatic pressure gradient
obtained from the graph for using different density proppants other than sand.
Phydro = (Ph )l x D

3-101

Pf - Friction pressure drop during pumping can be determined from


appropriate pipe friction pressure losses vs. flow rate equations or nomographs.
The rheological properties of fracturing fluids are particularly useful for the
estimation of the friction pressure drop. This is true not only for the calculation
of the treating pressure, it is especially necessary for the prediction of the
downhole fracture propagation (net) pressure, since measurements are usually
impractical. Downhole pressure gauges cannot be used unless they are
permanently installed and insulated from the proppant-carrying slurry or they
measure the pressure indirectly as in the annulus.
The real-time measurement or extrapolation of the net fracturing pressure is a
potent tool for the detection of the morphology of the created fracture. To
calculate the friction pressure drop for a power law fluid, at first the Reynolds
number must be estimated. The fracturing fluid consistency idex can calculated
from the following equation:

n = 3.32 log

600
300

3-102

The fracturing fluid consistency idex can calculated from the following
equation:

K=

510 300

3-103

511n

The fracturing fluid Reynolds, NRe , number can calculated from the following
equation:

N Re

89,100 p v
=
K

2 n

0.0416d

3 +1/ n

3-178

3-104

Table 3-3: Typical Proppants and Their Characteristics


Type
Northern White Sand

Texas Brown Sand

Curable Resin-coated
sand

Precured Resin-coated
sand

ISP
ISP-lightweight
sintered bauxite

Zirconium Oxide

Mesh Size

Particle Size
(in.)

Density
(lb/ft3)

Porosity

12/20

0.0496

165

0.38

16/30

0.0350

165

0.39

20/40

0.0248

165

0.40

12/20

0.0496

165

0.39

16/30

0.0350

165

0.40

20/40

0.0248

165

0.42

12/20

0.0496

160

0.43

16/30

0.0350

160

0.43

20/40

0.0248

160

0.41

12/20

0.0496

160

0.38

16/30

0.0350

160

0.37

20/40

0.0248

160

0.37

12/20

0.0496

198

0.42

20/40

0.0248

202

0.42

20/40

0.0248

170

0.40

16/20

0.0400

231

0.43

20/40

0.0248

231

0.42

40/70

0.0124

231

0.42

20/40

0.0248

197

0.42

3-179

Table 3-4: Properties for Various Size Proppants


Maximum Diameter
-in.

Average Proppant
Diameter PD)
in.

0.173

6/12**

0.132

0.099

8/12

0.093

0.087

8/16**

0.093

0.082

10/20

0.079

0.061

10/30

0.079

0.056

12/20*

0.067

0.054

16/20**

0.047

0.041

16/30**

0.047

0.039

18/20**

0.039

0.036

18/35

0.039

0.032

20/40*

0.0336

0.0272

20/50

0.0336

0.0218

30/50**

0.0237

0.0185

30/60

0.0237

0.0180

40/60

0.0168

0.014

40/70*
70/140**
(100 mesh)

0.016

0.013

0.0084

0.0099

Proppant Size
-mesh
4/8

* recognized by API as a primary proppant size


** recognized by API as an alternate proppant size
NOTE:

Average proppant diameters shown are based on typical size


distribution from screen analysis data, not the average of large and
small sieve openings.

3-180

Table 3-5: Specific Gravities and Absolute Volumes


Specific

Proppant

Gravity

Sand

Absolute
Volume
gal/lb

2.65

0.0453

SUPER PROP*

3.55-3.73

0.0338-0.0322

Resin-Coated
Sand

2.57-2.61

0.0467-0.0460

Z-PROP 126**

3.17

0.0379

INTERPROP
***

3.13

0.0384

2.73

0.0440

or
CARBOPROP***
PRO-FLO ****
*
**
***
****

High Strength Sintered Bauxite proppant


Zirconium Oxide proppant
Intermediate Strength Sintered Bauxite proppant
Intermediate Strength Low Density Sintered proppant

3-181

Figure 3-30: Shows the relationship between the API gravity


versus fluid density

Figure 3-31: Illustrates the hydrostatic pressure gradients for


various fluid densities

3-182

For laminar flow (NRe < 2100), the Fanning friction factor for smooth pipes can
be calculated from the well known expression:
n 3 + 1/ n

Kv

0.0416

=
dL
144,000d 1+ n

dp f

3-105

If the turbulent flow (NRe > 2100), then the Fanning friction factor is calculated
from figure 3-33 after calculating the Reynolds number from the above
equation. The friction pressure drop in oilfield units is given by the following
equation:
2

dp f f p v
=
dL
25.8 d
Where

=
u =
d =
K =
qi =

3-106

density in lb/ft3
velocity in ft/sec
diameter in in.
lbf-secn/ft2
the injection rate in, bpm

The calculation implied by the equations above would result in an upper limit
of the friction pressure drop. Polymeric solutions are inherently friction
pressure reducing agents and the actual Pf is likely to be smaller that the one
calculated The required pump hydraulic horsepower for the fracturing
treatment can be calculated from the following equation:
HHP = 0.0245 x qi x Pw
Where
qi = fracturing fluid injection rate, bpm
Pw = pump pressure, psi

3-183

3-107

Figure 3-33: Friction factors for power-law fluid model

3-184

3.2.22 PRODUCTIVITY RATIO:


It is defined as the fractured well productivity, and it is directly proportional to
the formation capacity (both expressed in mid-ft). The productivity ratio is the
ratio of the productivity index of the well after fracturing to that of the well
before fracturing, Jf/JO. Van Poollen, Tinsley and Saunders presented the
following equation for determining the ratio of well productivity before and
after stimulation by a horizontal fracture.

qf
qo

r
Log( f )
rw
r
Log( f )
r
rw
Log( e ) +
(
kh
)f
rf
[1 +
]
(kh )o

3-108

The productivity ratio can also be expressed as:


PR =

kavg

3-109

Where
kavg = average permeability of the fractured formation
k = the permeability of the unfractured formation
The average permeability of the fracture zone resulted from a horizontal
fracture is equal to the average permeability predicted for radial flow in parallel
beds. Figure 3-34 illustrates the configuration of the shape of the horizontal
fracture.
k fz =

k f w f + kh
h

3-110

Where
kfz = the average permeability of the fracture zone
kf = the permeability of the fracture
wf = the fracture width
The average permeability of the fractured formation, kavg , is equal to the
average permeability predicted for series beds in radial flow

3-185

kavg =

kk fz ln (re / rw )

3-111

k fz ln re / r f + k ln r f / rw

Substituting the average zone permeability equation into the average fracture
formation permeability equation yields:
k f w f + kh
ln (re / rw )
k

kavg =
k f w f + kh

ln re / r f + k ln r f / rw

3-112

Dyes et al also presented the following equation to calculate the productivity


ratio of a horizontal fracture.

kh
r

+ 1 ln

qf k fW f
k f w f rw

qo Kh k f W f r e r f

+ 1 ln
+ ln
r
r
kh
w
f

3-113

Where:
qf = well production rate after fracturing.
qo = well production rate before fracturing.
The productivity improvement resulting from a vertical hydraulically created
fracture can be predicted by the following equation which was introduced by
Prats:
PR =

kL
2re
2k f w f
ln
L

ln re / rw


+ 1

0.3
rw kL


ln 2k w
100
L
f f


+ 1

3-114

The productivity index before fracturing is defined as:

Jo =

qo
p p wf

3-186

3-115

Where
qo = production rate before fracturing
Pe = Reservoir average pressure
Pwf = bottom hole flowing pressure before fracturing.
The productivity ratio after fracturing is defined as:
Jf =

qf

3-116

p p wf

Where:
Pwf = bottom hole flowing pressure after fracturing

The productivity ratio is defined as:


J
PR = f
Jo

3-117

The stimulation ratio is defined as:

7
.
13
Jf =

J o ln 0.472( r e )

rw

3-118

The productivity ratio can be calculated from figure 3-35, which was given by
McGuire and Sikora (1960). This nomograph is used for the prediction of
fractured well performance and a comparison of pre- and post-treatment
productivity indices. Their work was based on a physical analog potentiometric
experiments. Figure 3-35 shows the expected multiple increase in the
productivity index from a fracture versus the relative fracture conductivity
(graphed on the abscise) for a variety of penetrations of the fracture as given by
ratio L/re . The McGuire and Sikora curve is for pseudo steady state or in other
words, for closed, depleting reservoirs. The ordinate provides the ratio of the
productivity indices before and after fracturing. (Note: In the McGuire and
Sikora curve, the fracture width is in inches).
3.2.23 OPERATIONAL CONSTRAINTS
Pump rate and treating pressures are usually considered as the operational
constraints for a number of reasons. First, injection at a higher rate has the
potential of fracturing out the productive zone as a result of the higher treating
pressure. Second, for some fluids severe degradation of fracturing fluid may
3-187

occur under high shear rate. Third, high pipe friction pressure results in high
treating surface pressure, which may need to be limited due to available
horsepower and limits on allowable pressure for the well tubular. Therefore, it
is important to determine the pumping parameters that will ensure operating
conditions during the injection not to exceed the pre-determined maximum
allowable pressure and rate.
3.2.24 ECONOMIC CONSIDERATIONS:
One of the basic requirements in designing a fracture treatment is to maximize
economic returns. One criterion to determine the most cost-effective treatment
is to maximize the net present value of the various design options. This
criterion will be considered in the following sections for illustrative purposes.
Britt (1985) used the net present value and discounted return on investment and
compared their differences. The fracture net present value (NPV) will be
calculated by subtracting the total treatment cost from the discounted well
revenue. The equations used to calculate the total treatment cost as well as the
discounted well revenue for a given fracture length is summarized as follows:
3.2.25 Treatment Cost
The equation to calculate the total treatment cost is given by
Treatment cost = (Fluid + Proppant + Horsepower
+ Miscellaneous) cost

3-119

Where
Fluid cost
= $/gallon x gallons of fluid
Proppant Cost = $/pounds x pounds of proppant,
And

HHP Cost = $ / hhp

q i p surf

3-120

40.8

Where qi denotes the pump rate and Psurf is the surface pumping pressure.
Discounted Well Revenue:
The equation to calculate the discounted well revenue for n years is given by:
Discount Well Re venue = nj =1[total annual net revenue during
year J ) /(1 + j )i ]

3-188

3-121

Figure 3-34: Illustrates the shape of a horizontal fracture showing


the fracture extension and then fracture permeability

Figure 3- 33: McGuire and Sikora (1960) curves for folds of increase
PI/PIO in a bonded reservoir of area A (acres).

3-189

Where
i denote the discounted rate.
The net well revenue is calculated by:
Net Revenue = $/bbl (or $/Mscf) x (fractured - Unfractured)
x Production

3-122

Also considerations of taxes and production costs should generally be included.


Net Present Value:
The equation to calculate the fracture net present value is given by:
Fracture NPV = Discounted Well Revenue - Treatment Cost

3-123

3.2.26 Treatment optimization design procedure:


Optimization Criteria:
The criteria to be considered in determining the optimum size of the treatment
are (1) optimize the reservoir deliverability, (2) maximize the proppant
penetration, (3) Optimize the pumping parameters, (4) minimize the treatment
cost, and (5) maximize economic returns of the well. The step-by-step design
procedures to illustrate an optimum fracture treatment are summarized as
follows:
1. Assume a hydraulic fracture length and a maximum proppant concentration.
2. Select the appropriate fracture propagation model (e.g. PKN or KGD)
based on formation characteristic pressure behaviour observed during the
pre-treatment in- situ stress test or mini-fracture.
3. Select several candidates fracturing fluid systems applicable to the
formation being fractured.
4. Select the appropriate proppant type based on sand crushing and formation
embedment characteristics.
5. Determine the maximum allowable treating pressure versus fracture height
relationships based on the insitu stress distribution.
6. Determine the maximum allowable pump rate based on the pipe frictional
pressure losses, shear degradation factors of the fracturing fluid, and
equipment limitation.

3-190

7. Optimize the propped fracture geometry, fluid volume, and proppant


requirements. The optimization procedures involved are to (a) determine the
optimum pumping parameters subject to the treating pressure and pump rate
constraints (b) maximize the fluid efficiency and (c) maximize the proppant
penetration for a designed proppant concentration and the hydraulic fracture
length assumed in Step 1.
8. Construct transient IPR curves for various producing times based on the
propped fracture geometry (optimized) and fracture conductivity calculated
from Step 7.
9. Couple the transient IPR curves with tubing intake curves to obtain the
wellhead deliverability.
10. From the production decline curves obtain the annual cumulative recovery
for various producing times.
11. Calculate the present value of the net revenue of the production based on an
assumed discounted rate.
12. Calculate the treatment cost including the cost associated with fluids,
proppants, hydraulic horsepowers, and miscellaneous items.
13. Calculate the fracture net present value by subtracting the treatment cost
from the well's discounted revenue. This completes the computational
cycle for an assumed hydraulic length.
14. Construct the fracture net present value or appropriate economic criteria
versus fracture length curves by repeating the above computational cycles.
It should be pointed out that any point on the fracture net present value curve is
optimized based on the balance between the fracture characteristics, reservoir
deliverability, and treatment cost. The most cost- effective treatment design can
then be determined by maximizing the well's profitability. Once the optimum
design is selected, the treatment schedule can be calculated based on the
method generally used for design.

3-191

3.2.27 REFERENCES
[1]

Timoshenko, S. and Goodier, N.J. (1951). Theory of Elasticity. 2nd ed.


McGraw Hill, New York.

[2]

Ben-Naceur, K., (1989). Modeling of Hydraulic Fractures in Reservoir


Stimulation. 2nd ed. M.J. Economides and K.G. Notle, eds., Prentice
Hall, N.J.

[3]

Brown, J.E. and Economides, M.J., (1992). An analysis of hydraulically


fractured horizontal wells. SPE paper 24322.

[4]

Mukherjee, H. and Economides, M.J., 1991 (June). A parametric


comparison of horizontal and vertical well performance. SPEFE, pp.
209-216.

[5]

Howard, G.C. and Fast, C.R., (1970). Hydraulic fracturing, Mono. Ser.,
Society of Petroleum Engineering, 2, Texas.

[6]

Geertsma, J. and de Klerk, F., 1969 (December). A rapid method of


predicting width and extent of hydraulically induced fractures. JPT, pp.
1571-1581.

[7]

Hubbert, M.K. and Willis, D.G. Mechanics of hydraulic fracturing.


Trans AIME, 210, 153-168.

[8]

Khristianovic(h), S.A. and Zheltov, Y.O., (1955). Formation of vertical


fractures by means of highly viscous liquid. Proc., 4th World Petroleum
COngresws, Sec. II, 579-586.

[9]

Balen, M.R., Meng, H.Z and Economides, M.J. (1988). Application of


the NPV (New Present Value) in the optimization of hydraulic fractures.
SPE Paper 18541.

[10]

Haimson, B.C. and Fairhurst, C. (1969). Hydraulic fracturing in porous


permeable materials. J. of Pet. Eng. Tech., 811-817.

[11]

Haimson, B.C. and Fairhurst, C. (1967). Initiation and extension of


hydraulic fractures in rocks. S. of Pet. Eng. J., 310-318.

[12]

Nordgren, R.P (1972). Propagation of a vertical hydraulic fracture.


SPEJ, 12, 306-314.

[13]

Detournay, E., Cheng, A. H-D, and McLennan, J.D. Proelastic PKN


hydraulic fracture model based on an explicit moving mesh algorithm.
J. of Energy Resources Tech. Trans. ASME, 112, 224-230.

[14]

Perkins, T.K. and Kern, L.R., 1961 (September). Widths of hydraulic


fracture. JPI, 937-949.

[15]

Prats, M., 1961 (June). Effect of vertical fractures on reservoir


behaviour incompressible fluid case. SPEJ, 27-32.

[16]

Valk, P.,and Economides, M.J., 1993 (February). A continuum damage


mechanics model of hydraulic fracturing. JPT, 198-205.
3-192

[17]

Shah, S.N., 1989 (November). Proppant-settling correlations of nonNewtonian fluids. SPE Production Engineering Journal, 446-448

[18]

Shah, S.N., 1991 (October). Rheological characterization of hydraulic


fracturing slurries. SPE 22839, presented at the 66th Annual Technical
Conference and Exhibition of SPE in Dallas, Texas.

[19]

Schlumberger (1989). Log Interpretation Principles. I, Schlumberger


Publications.

[20]

Fertl, W.H. (1976).


Amsterdam.

[21]

Benham, P. and Warnock, F., (1973).


Structures. Pitman, London.

[22]

Matthews, W. and Kelly, J. 1967 (February). How to predict formation


pressure and fracture gradient. Oil and Gas Journal, 92-106.

[23]

Pilkington, P.E., 1979 (May). Fracture gradient estimates in tertiary


basins. Petroleum Engineer. 138-148.

[24]

Eaton, B.A., 1969 (October).


Fracture gradient prediction and
interpretation of hydraulic fracture treatments. Journal of Petroleum
Technology, 21-29.

[25]

Christman, S.A. 1973 (August). Offshore fracture gradients. Journal of


Petroleum Technology, 910-914.

[26]

Behrmann, L and Elbel, J., 1991 (May). Effect of perforation on


fracture initiation. SPE 20661 Journal of Petroleum Technology, 43, 5,
608-615, also presented at SPE Conference and Exhibition, New
Orleans (1992, September).

[27]

Bennett, C.O., Rodolpho, C.V, Reynolds, A.C. and Raghavan, R., 1985
(October). Approximate solutions for fractured wells producing layered
reservations. SPE 11599, SPE Journal, 25, 729-742.

[28]

Bennett, C.O., Rosato, N.D., Reynolds, A.C. and Raghavan, R., 1983
(April). Influence of fracture heterogeneity and wing length on the
response of vertically fractured wells. SPE 9886, SPE Journal, 23, 219230.

[29]

Brown, K.E. et al., (1984). The technology of Artificial lift methods.


PennWell Publishing Co., Oklahoma.

[30]

Cameron, J.R. and Prud'homme, R.K. (1989). Fracturing fluid flow


behavior. Recent Advances in Hydraulic Fracturing. J.L. Gidley, S.A.
Holditch, D.E. Nierode and R.W. Beatch Jr. (eds), Monograph Series,
Society of Petroleum Engineers, Texas, USA.

[31]

Collins, R.E., (1961). Flow of Fluids Through Porous Materials. Van


Nostrand Reinhold, New York, USA

Abnormal Formation Pressures.

3-193

Elsevier,

Mechanics of Solids and

[32]

Cramer, D., 1987 (December). Limited entry extended to massive


hydraulic fracturing. Oil and Gas Journal. 85, No. 50, 40-50.

[33]

Dake, L., (1982).


Fundamentals of Reservoir Engineering,
Development in Petroleum Science. Elsevier Scientific Publishing Co.,
Amsterdam, 8.

[34]

Ely, J.W., (1985). Stimulation Treatment Handbook.


Publishing Co., Oklahoma, USA.

[35]

Geertsma, J. and Haafkens, R., 1979 (March). A comparison of the


theories to predict width and extent of vertical hydraulically inducted
fractures. Trans. AIME, 101, 8.

[36]

Haimson, B.C. and Huang, X., 1993. The Hydraulic Fracturing Method
of Stress Measurement: Theory and Practice.. Comprehensive Rock
Engineering, J. Hudson (ed.), Pergamon Press, U.K., 3, 297-328.

[37]

Haimson, B.C. and Huang, X., 1989. Hydraulic Fracturing Breakdown


Pressure and In-Situ Stress at Great Depth. Rock at Great Depth. V.
Fourmaintraux and D. Maury (eds). Balkema, Netherlands, 939-946.

[38]

Khristianovich, S.A., Zheltov, Y.P, Barenblatt, G.I. and Maximovich,


G.K., 1959. Theoretical Principles of Hydraulic Fracturing of Oil Strata.
Proc. 5th World Petroleum Congress, New York.

[39]

Lamb, H., 1932. Hydrodynamics. 6th ed., Dover Publication, New York,
USA, 581-587.

[40]

Muskat, M. 1937. The Flow of Homogeneous Fluids Through Porous


Media. McGraw-Hill Book Co., New York, USA.

[41]

Muskhelishvili, N.I., 1953. Some Basic Problems of the Theory of


Elasticity. 4th ed., J.R. Radok (trans), P. Noordhoff, Netherlands.

[42]

Penny, G.S., Conway, M.W. and Lee, W., 1985 (June). Control and
modeling of fluid leakoff during hydraulic fracturing. SPE 12486, J. of
Pet. Tech., 37, No. 6, 1071-1081.

[43]

Perkins, T.K., 1973 (January). Discussion of 'On the design vertical


hydraulic fractures'. J. of Pet. Tech., 25, No. 1, 80-81. Also in Trans.
AIME (1973), 255, 93-95.

[44]

Perkins, T.K. and Kern, L. R., 1961 (September). Widths of hydraulic


fracture. SPE 89, J. of Pet. Tech., 13, No. 9, 937-949.

[45]

Prats, M., 1961 (June). Effect of vertical fractures on reservoir behavior


incompressible fluid case. SPE 1575-G. PSE Journal, 1, No. 1, 105118. Also in Trans., AIME, (1961), 222.

[46]

Settari, A., 1985 (August). A new generation model of fluid loss in


hydraulic fracturing. SPE 11625, SPE Journal, 25, No. 4, 491-501.

3-194

PennWell

[47]

Shah, S.N. and Lord, D.L., 1990 (September). Hydraulic fracturing


slurry transport in horizontal pipes. SPE 18994, SPE Drilling
Engineering, 5, No. 3, 225-232.

[48]

Smith, M.B. and Hannah, R.R., 1996 (July). High permeability


fracturing: the evolution of a technology. SPE 27984-P, J. of Pet.
Tech., 628.

[49]

Soliman, M.Y., Hunt, J.L. and El Rabaa, A.M., 1990 (August).


Fracturing aspects of horizontal wells. SPE 18542, J. of Pet. Tech., 42,
No. 8, 966-973.

[50]

Thiercelin, M., 1989 (June).


Fracture toughness and hydraulic
fracturing. International Journal of Rock Mechanics Sciences and
Geomechanics Abstracts, 26, Nos. 3-4, 177-183.

[51]

Van Poollen, H.K., Tinsley, J.M and Saunders, C.D., (1958). Hydraulic
fracturing: fracture flow capacity vs. well productivity. Trans. AIME,
213, 91-95.

[52]

Yew, C.H. and Li, Y., 1993 (August). Fracture tip and critical stress
intensity factor of a hydraulically induced fracture. SPE 22875, SPE
Production & Facilities, 8, No. 3, 171-177.

[53]

Penny, G.S., 1986. An investigation of the effects of fracturing fluids


upon the conductivity of proppants. Final report STIM-LAB, Oklahoma,
USA.

[54]

Penny, G.S., 1988. An investigation of the effects of fracturing fluids


upon the conductivity of proppants. Final report STIM-LAB, Oklahoma,
USA.

[55]

Hubbert, M.K. and Willis, D.G., 1956 (October). Mechanics of


Hydraulic Fracturing. Paper 686-G presented at SPE Annual Meeting,
Los Angeles, USA also in Journal of Petroleum Technology (September
1957), 9, No. 6, 153-168 and Trans AIME (1957) 210.

[56]

Howard, G.C. and Fast, C.R. (1970) Hydraulic Fracturing. Monograph


Series, Texas, USA, Society of Petroleum Engineering, 2.

3-195

3.2.28 HYDRAULIC FRACTURING DESIGN PROBLEM


A fracturing treatment is intended to be conducted in an oil well completed in a
tight limestone formation (Sirt Formation) in order to increase the oil
production rate from the well from 88 bbls/day to 1000 bbl/day. Given the
following information:
Depth
= 7650 ft
Producing Interval
= 7600-7695 ft
Formation thickness
= 95 ft
Fracturing gradient
= 0.7075 psi/ft
Minimum horizontal stress
= 4185 psi
Overburden pressure gradient
= 1.0980 psi/ft
Fracturing angle
= 60 degree
Reservoir oil compressibility
= 0.845x10-5 psi-1
Reservoir water compressibility
= 2.30x10-5 psi-1
Gas saturation
= 0%
Oil saturation
= 0.64%
Gas gravity
= 0.890
Oil formation volume factor
= 1.17 rb/stb
Connate water saturation
= 34%
Formation porosity
= 3.14%
Poissons ratio
= 0.158
Bulk compressibility with porosity
= 3.8x10-7 psi-1
Rock compressibility with zero porosity
= 1.93x10-7 psi
Bulk modulus
= 2.63x106 psi
Biot poroelastic constant
= 0.492
Fracturing fluid viscosity
= 40 cp
Fracturing fluid density (Versa Gel)Sp. Gr. (1.002) = 8.36 ppg (35 API)
Frictional pressure gradient inside tubing
= 0.1501 psi/ft
Original formation permeability
= 5 md
Reservoir temperature (BHST)
= 200 F
Geothermal gradient
= 0.015 F/ft
Average reservoir pressure (BHSP)
= 3530 psi
Reservoir fluid viscosity
= 1.190 Cp
Area of filter medium
= 22.8 cm2
Slope of fluid loss curve at lab.
= 1.8cm/min1/2
Filtration pressure at lab.
= 100 psi
Youngs modulus
= 1.057x107 psi
Casing outer diameter
= 9.6250 in
Casing inner diameter
= 8.6810 in
Wellbore diameter
= 10.75 in
Drainage diameter
= 1640 ft
Proppant size and type (Z-proppant)
= 20/40 mesh
Porosity of packed proppant
= 35%
Specific gravity of proppant
= 2.63
Bottom hole flowing pressure before fracturing
= 1700 psi
Fracturing fluid spurt loss
= 0.0100 gal/ft2
3-196

Tubing outer diameter


Tubing inner diameter
Tubing depth
Gas liquid ratio
Bubble point pressure
Bottom hole temperature
Skin facture before fracturing
Perforation diameter
Perforation discharge coefficient
Number of perforations
Closure stress
Well spacing
Assume

hf = h,

qi = 15 bbl/min,

= 3.5 in
= 2.9910 in
= 7591 ft
= 151 scf/bbl
= 630 psi
= 200F
= 20
=
= 0.87
= 100
= 4550 psi
= 70 acres
Vi = 180 bbls

Calculate
1. The formation fracturing pressure.
2. The effective fracturing fluid coefficient.
3. The fracture volume.
4. The fracture efficiency.
5. The concentration of proppant in the fracturing fluid.
6. Wellhead injection pressure.
7. The wells productivity ratio.
8. The bottomhole flowing pressure after fracturing.
9. The oil flow rate after fracturing.
10. The skin factor after fracturing.

3-197

3.2.29 HYDRAULIC FRACTURING DESIGN PROBLEM SOLUTION


1.

Calculations of the formation fracturing pressure:



h min = Pob

+ Pp 1
1
1

Pob = 0.433 x Pb x D

= 0.433 x 2.536 x 7650 = 8400 psi

1 t s

2 t c
t s

t c

= 1

Cb =

1 130

1
2 80
2

130

1
80

= 0.195

Cr
Cb

1
Kb

3t s2 42c
10
K b Pb
x 1.34 x 10
2
2
3t s c
3 x130 2 4x80 2
10
6
= 2.536
x 1.34 x 10 = 2.629 x 10 psi
2
2
3 x130 x80

Cb =
Cr =

1
= 3.8 x 10 7 psi 1
6
2.629 x 10
2
3t 2ma t sma

2
Pb 3t sma
4 t 2ma x 1.34 x 1010

3 x 55.5 2 x 88 2

2.536 3 x 88 4 x 55.5 x 1.34 x 10

= 1

1.9299 x 10 7
3.8 x 10 7

10

= 0.492

198

= 1.9299 x 10 7 psi 1

Calculation of minimum fracturing pressure:


Pob = Gob x D
= 1.098 x 7650 = 8400 psi

2
( v Pp ) + T0
1
Pf =
+ Pp
1 2
2

1
0.195
2x
(8400 3530 ) + 705
1 0.195
Pf =
+ 3530
1 0.195
2 0.492 x

1 0.195

Pfrac = 5412 psi


Calculation of fracturing fluid co-efficient:
P (closure stress) = Pf - Pres

= 5412 3530 = 1882 psi


C = 0.0469
= 0.0469

K P
frac

0.005 x 0.0314 x 1882


= 0.00403 ft
min
40

CT = coso + cwsw + cgsg


= 8.45x10-6x0.64+23x10-6x0.34=13.228x10-6 psi-1
C c = 0.0374 P

KC T
res

= 0.0374 x 1882

0.005 x 0.0314 x 13.228 x 106


= 0.00294
1.19

m act = m lab x

Pact
1882
= 1.8 x
= 7.809
Plab
100

199

ft
min

Cw =

0.0164
0.0164
x 7.809 = 0.00562 ft
x m act =
min
22.8
Af

1
1
1
1
=
+
+
CT C CC Cw
1
1
1
1
=
+
+
C T 0.00403 0.00294 0.00562

C T = 0.00131 ft

min

Calculation of Fracture Dimensions:


Assume:

t=

qi

= 15 bbl/min

Vi

= 180 bbls

Vi 180
=
= 12 min
i
15

8 x 0.00131 x 12
8 C T t
=
Ww + 8S p Ww + 8 x 0.01
12
7.48
=

0.0643
0.262 Ww + 0.011

8 x 0.01 2
2

1
12 W w + 7.48 e erfc ( ) +

L=

q i x 5.615
64 h f C T

L=

8 x 0.01 2
2

1
Ww +
e erfc () +

7.48
64 x 95 x 0.00131 12

15 x 5.615

2
2
= 2571[0.262Ww + 0.011]e erfc () +
1

Ww
0.05
0.1
0.2
0.3
0.4

2.6681
1.7285
1.0142
0.7176
0.5553

L
136
117
91
76
69

A
51680
44460
34580
28880
26220

200

AQ
307
264
205
172
156

0.5
0.6
0.7
0.8
0.9
1.0

0.4528
0.3823
0.3303
0.2915
0.2605
0.2355

55
48
43
40
38
35

20900
18240
16144
15200
14440
13300

124
108
96
90
86
79

A = 4 L hf = 4 x 95 x L = 380 L
AQ =
E =

A
A
A
=
=
2q i 2 x 15 x 5.615 168.45

1.06 x 10 7
E
=
= 1.087 x 10 7 psi
2
2
1
1 0.158

q L2
W w = 0.350 i
E h f

using KGD model

40 x 15 x L2
= 0.350

7
1.087 x 10 x 95
L

re

0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1.0

= 0.0097 L

Ww

AQ

82
164
246
328
410
492
574
656
738
820

0.0875
0.1238
0.1516
0.1751
0.1958
0.2144
0.2316
0.2476
0.2626
0.2769

31160
62320
93480
124640
155800
186960
218120
249280
280440
311600

185
370
550
740
925
1110
1295
1480
1665
1850

A = 4 L hf = 4 x 95 x L = 380 L
AQ =

A
A
A
=
=
29 2 x 15 x 5.615 168.45

201

In order to solve for L and Wf. a plot of log AQ versus log Ww (assumed fracture
width) can be constructed. On the same graph paper, assumed values of fracture
length L are plotted versus log AQ. The intersection of these curves represents a
solution that satisfies both length and width equations. Therefore, the fracture
width and the corresponding AQ values, are obtained from the intersection of
these curves.
Fracture width, ww = 0.1 in.
AQ = 260
A
29 i x 5.615

AQ =

A = (AQ) x 2 i x 5.615
= 260 x 2 x 15 x 5.615 = 43797 ft2
L=

A
43797
=
= 115 ft 2
4hf
4 x 95

Calculation of Fracturing Efficiency:


vf
x 100
vi

Eff =

vf =
=

Eff =

L ww x h f
x 115 x

0.1
x 95 = 143 ft 3
12

143
x 100 = 14%
180 x 5.615

Calculation of Proppant Weight Needed:


(weight of proppant needed) = vf x (1 - ) x Pprop
= 143 x (1-0.35) x 2.63 x 62.4
= 15254 lbm

202

Proppant concentration, cprop =


c prop =

wt
vi

15254
= 2.02 lbm
gal
180 x 42

Fracturing fluid density including proppant

p mix =

8.34 x 8.36 / 8.34 + 2.02


8.34 + x
=
= 9.505 lbm
gal
1 + 0.0456 x
1 + 0.0456 x 2.02

Fracturing fluid injection rate including proppant:


wt of proppant

Vi +
2.63 x 62.4 x 5.615
qi =
tp

180 +
=

15254
2.63 x 62.4 x 5.615
= 16.38 bbl
min
12

Calculation of wellhead injection pressure:


Phydrostatic = 0.052 Pmix x D
= 0.052 x 9.51 x 7650 = 3783 psi
Pfriction = G friction x D

= 0.1501 x 7650 = 1148 psi

Pperf

qi
= 0.237 x Pmix
2
c p x N d perf

Pperf = 0

open hole completion

203

Pw = Pf Phydro + Pfric + Pperf

= 5412 3783 + 1148 + 0 = 2777 psi


Calculation of Fracture conductivity using propped fracture permeability curves
Kf = 200 x 103 md
Fracture conductivity, FC
FC = k f x w f

0.1
= 200 x 10 3 x
x = 1308 md _ ft.
12 4

Calculation of production increase:


The productivity ratio can be calculated from the production increase curves after
McGuire and Sikora (1960).
Relative Conductivity =

3
0.1 x x 200 x 10
4

=
5
115
L =
= 0.14
re 820

wf k f
ki

40
A

40
= 2520
62

Entering the production increase curve with these values:


Jf

7013
= 2.7

J o ln 0.472 re / rw
Jf
Jf
Jf

Jo

= (2.7 x ln 0.472 re rw) / 7.13

Jo

= (2.7 x ln 0.472

Je

820 x 12
) / 7.13 = 2.13
10.75

= 2.3

Productivity ration = 2.3


204

Production of IPR curve before fracturing


qo =
PI =
=

7.08kh( Pe Pwf )

Bo ln re rw
7.08 kh
Bo ln re

rw
7.08 x 0.005 x 95

1.19 x 1.17 x ln (820 x 12 x 2

PI =

= 0.32 B / D / psi
10.75

q
Pe Pwf

q = PI (Pe Pwf)
q = 0.32 (3530 Pwf)
Assumed
Pwf (psi)
1000
2000

qe (bbl/D)
814
490

Construction of tubing intake curve using Brown correlations


Assumed
q
800
1000
2000

Pth (psi)

Pwf (psi)

120
160
280

1390
1520
2000

From the intersection of IPR curve with the tubing intake curve TPC:
qo

optimum before fracturing = 750 B/D

Pwf

optimum before fracturing = 1300 psi

Pth

= 200 psi

IPR Curve after fracturing:

205

qf

=
J o Pe Pwf

qf
2.3 =
3530 P
wf

qe

P P
wf
e

Jf

750

3530 1300

Assumed
Pwf psi

qf (B/D)

1000
2000
3000

1958
1184
410

From IPR and tubing intake curves:


qf
Pwf

after fracturing
after fracturing

= 1450 bbl/d
= 1650 psi

206

Anda mungkin juga menyukai