You are on page 1of 39

Chapter One

General Overview of Hydraulic


Fracturing
General Overview of Hydraulic Fracturing:

1.1 Introduction:

Hydraulic fracturing has been, and will remain, one of the primary engineering
tools for improving well productivity. This is achieved by:

❑ Placing a conductive channel through near wellbore damage, by passing


this crucial zone.

❑ Extending the channel to a significant depth into the reservoir to further


increase productivity.

❑ Placing the channel such that fluid flow in the reservoir is altered. In this
last instance, the fracture becomes a tool for true reservoir management
including sand deconsolidation management and long-term exploitation
strategies.

Complexity arises from two directions: geologic reality and the inherent
multidisciplinary nature of the fracturing process. Historically, the control of
fracturing has rested with drilling and operations groups owing to the nature
of field procedures using pumps, packers, pressure limits, etc. However, the
final results (and thus design) are dominantly a production engineering
exercise, and fracturing cannot be removed from intimate contact with
reservoir engineering. At the same time, designing a treatment to achieve the
desired results is also intimately connected with rock mechanics, fluid
mechanics and the chemistry that governs the performance of the materials
used to conduct the treatment.

However, the design must also be consistent with the physical limits set by
actual field and well environments. Also, treatments must be conducted as
designed to achieve a desired result (i.e., full circle to the critical role of
operations). Proper treatment design is thus tied to several disciplines:

▪ Production engineering,
▪ Rock mechanics,
▪ Fluid mechanics,
▪ Selection of optimum materials, and
▪ Operations.

1.2 What is Fracturing?

If fluid is pumped into a well faster than the fluid can escape into the
formation, inevitably pressure rises, and at some point something breaks.
Because rock is generally weaker than steel, what breaks is usually the
formation, resulting in the wellbore splitting along its axis as a result of tensile
hoop stresses generated by the internal pressure.

Figure 1: Internal pressure breaking a Figure 2: Cross-sectional view of a


vertical wellbore. propagating fracture

1.3. Why Fracture?

Hydraulic fracture operations may be performed on a well for one (or more)
of three reasons:
❑ To by pass near-wellbore damage and return a well to its “natural”
productivity.

❑ To extend a conductive path deep into a formation and thus increase


productivity beyond the natural level.

❑ To alter fluid flow in the formation. In the third case, fracture design
may affect and be affected by considerations for other wells (e.g., where
to place other wells and how many additional wells to drill). Although
these three motivations where the all-important production rate (q) is
related to formation permeability (k), pay thickness (h), reservoir fluid
viscosity (μ), pressure drop (Δp) and formation flow area (A).

The actual philosophy shift for fracturing, from accelerating production from a
single well to reservoir management, occurred with the application of massive
stimulation treatments in tight gas formations .Although outwardly a
traditional application of fracturing to poorer quality reservoirs, these
treatments represented the first engineering attempts to alter reservoir flow
in the horizontal plane and the methodology for well placement (e.g., Smith,
1979). Fracturing for vertical inflow conformance (i.e., reservoir
management) was successfully used in the Gullfaks field (Bale et al., 1994),
where selective perforating and fracturing were used to optimize reserve
recovery and control sand production while maintaining (but not necessarily
increasing) the required production rates.

Without fracturing, the entire zone can be perforated, and a low drawdown
allows a significant production rate on the order of 20,000 STB/D, sand free.
However, sand production is triggered by water breakthrough in the high-
permeability zone (from down dip water injection). The resulting wellbore
enlargement caused by sand production acts to stimulate production from the
high permeability zone. To stop sand production, draw-down must be reduced
even more. The production is then essentially 100% water coming from the
stimulated high-permeability zone, and the well must be abandoned. This
further diminishes production from the large reserves found in the deeper
zones with lower permeability. Open- or cased hole gravel packing could be
used to eliminate the sand production. However, such completions are less
than satisfactory for two reasons.
First, the deeper, lower permeability zones can significantly benefit from
stimulation.

Second, significant scaling occurs with water breakthrough and quickly plugs
the gravel pack. The fracturing tool selected to manage the Gullfaks field is
termed an indirect vertical fracture completion (IVFC). The IVFC accomplishes
several goals:

▪ Some (although choked) production is achieved from the main zone to


enable the well to reach minimum productivity standards.

▪ Production from the lower, moderate-permeability zone is stimulated,


maximizing reserves from this zone.

▪ Greater drawdown is allowed because the weak high-permeability rock


is separated from the perforations, and greater drawdown increases the
total rate and significantly increases recovery from the lower zones.

▪ If the upper high-permeability zone has sand production tendencies (as


is typically the case), then producing this zone via the fracture totally
avoids the need for sand control.

▪ Any potential for water breakthrough in the high-permeability zone is


retarded, and post water- breakthrough oil production is significantly
increased. To achieve these goals, fracture conductivity must be tailored
by synergy between the reservoir and fracture models. Too much
conductivity accelerates production and the time to water breakthrough
from the high-permeability main zone. Also, too much conductivity,
because of surface or tubular limits for the production rate, restricts
drawdown on the lower zones, and the desired, more uniform vertical
production profile is not achieved.

The fracture design goal is not to simply accelerate the rate but to achieve
maximum reserves recovery with no sacrifice of rate (as compared with a
simple completion in which the entire zone is perforated).
Chapter Two

Literature Review
Literature Review:

2.1 History of Hydraulic Fracturing:

The first experimental hydraulic fracturing treatment in the United States took
place in 1947 in the Hugoton gas field in Grant County, Kansas, Figure (2.1).
It was done on a small scale to bypass pore space near the wellbore in the oil-
bearing rock formation that was clogged by drilling mud during drilling
operations (Montgomery and Smith, 2010). In 1949 a patent was issued to
the Halliburton Oil Well Cementing Company, which then performed the first
two commercial fracturing treatments in Oklahoma and Texas. At that time
the engineering was simple and unsophisticated.

Figure 3: First experimental fracturing job conducted in 1947 by Stanolind


Oil in the Hugoton gas field of southwestern Kansas

Over the years, significant advances have been made in materials and
techniques, fracture modeling, fracturing fluids, and the types and amount of
equipment needed Figure (2.2). Today over 60% of all oil and gas wells drilled
worldwide are fractured, with more than 50,000 fracture stages completed
annually.
Figure 4: A very large, staged hydraulic fracturing job performed recently
on a Marcellus Shale multi-well pad in Pennsylvania (modified from U.S.
DOE and NETL, 2011).

Hydraulic fracturing has been and will remain one of the primary engineering
tools for improving well productivity in old and new wells. Fracturing has been
used successfully in all formations except those that are very soft. Fracturing
operations have proved successful in sand, limestone, dolomite and various
silicates.

2.2 Fracture Mechanism:

Fracture Mechanism can be divided into two steps:

1. Initiation Fracture, and


2. Extension Fracture.
2.2.1 Initiation Fracture:

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.
Compressive stress of the rock when fracture initiation pressure depends on:

▪ Pressure of injection fluid.


▪ The rock type of formation.
▪ Pump at the surface.

2.2.2 Extension Fracture:

As injection of fracturing 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.

2.3 In-situ Stresses:

The in-situ stresses acting on a formation can be decomposed into three


principal compressive stresses, one vertical and two horizontal. The two
horizontal compressive stresses are usually not equal. The vertical stress is
caused by the overburden weight acting on the top of a formation. The
horizontal stresses are the result of the poro-elastic deformation of the rocks
plus externally applied tectonic forces.

The parameters that affect the magnitude of the in-situ stresses include:

▪ Overburden weight,
▪ Fluid pore pressure,
▪ Porosity,
▪ Anomalies in the rock fabric (i.e., natural fractures),
▪ Rock mechanical properties (such as Poisson’s ratio), and
▪ Tectonic activity.
Figure 5: Initiation and extension

Knowledge of the in-situ-stress magnitude and direction can impact decisions


and designs throughout the drilling and completion of a well. During drilling,
in-situ stress may affect the mud and cement densities required to prevent
unwanted fracturing of open hole strata in the wellbore. For wells that will be
stimulated at high pressures, casing design must account for the maximum
anticipated stresses.
Figure 6: Functions of in-situ stress

Wellbore-stability calculations, particularly for horizontal wells, require


knowledge of in-situ-stress magnitude and direction. For hydraulic-fracture
treatment applications, the in-situ stresses control:

a) Fracture azimuth and orientation (vertical and horizontal)


b) Fracture-height growth,
c) Fracture width,
d) Treatment pressures, and
e) Fracture conductivity.

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 Figure (2.5).
Figure 7: The stresses that effect the subsurface formations

2.3.1 Vertical Stress:

Absolute vertical stress (σv), in pounds per square inch (psi) corresponds to
the weight of the overburden, and is given by:

σv = ρ×D/144

Where:

ρ = the density of the formations overlaying the target reservoir


(lb/ft3)

D = depth to the target reservoir (ft).

In a porous medium, the weight of the overburden is carried by both the grains
and the fluid within the pores. Accordingly, an effective stress (σv’), is defined
as:

σv’ = σv – (α×P)

Where:

α = Biot’s poro-elastic constant (dimensionless),


P = pore (reservoir) pressure (psi).
2.3.2 Horizontal Stresses:

The vertical stress is translated horizontally through Poisson’s ratio (ν):

σH’ = (ν/(1- ν))×σv’

Where:

σH’ = effective horizontal stress (psi).

The absolute horizontal stress is arrived at by adding the ( α×P) term to the
effective horizontal stress. Due to tectonic components, the horizontal plane
stress varies with direction. The above defined horizontal stress is the
minimum horizontal stress; the maximum horizontal stress is:

σH(max) = σH(min) + σ(tect)

Where:

σ(tect) = tectonic stress contribution (psi),

Horizontal stresses contained within stiff boundaries are generally considered


to be “locked-in place”, whereas the vertical stress follows the geologic history
(e.g., erosion, glaciations) of the overlying layers. Thus, horizontal, or
pancake fractures are likely to occur in a stiff, shallow formation with a
geologic history of surface erosion.

1. Maximum Horizontal Stress (σx): it is along x axis, which is


perpendicular the with overburden stress.

2. Minimum Horizontal Stress (σy): it is along y axis, which is


perpendicular with the overburden stress.

When the vertical stress (overburden) is applied, it tries to prevent the lateral
expansion of the rock. The horizontal strain equals to zero. For rocks in
compression (x) is essentially zero and since the lateral stress (σx) equals the
lateral stress (σy).
𝒗
𝝈𝒙 = 𝝈𝒚 = 𝝈𝒉 = 𝝈
𝟏−𝒗 𝒛
Where:

σh = Horizontal stress, psi

ν = Poisson ratio.

In-situ stress state of the reservoir rock depends on rock mechanical


properties:

▪ Modulus of elasticity (young's model, E),


▪ Poisson's ratio (ν),
▪ Rock strength (σ),
▪ Rock compressibility (Cr, Cb),
▪ Shear modulus (G), and

▪ Strain modulus ().

2.4 Fracture Orientation:

Formation fractures along the plane are perpendicular to the direction of the
least principle stress:

2.4.1 Vertical Fractures:

The fracture orientation is parallel to the overburden pressure (P ob), and the
fracture gradient is 0.7 psi per ft of depth or less, and the vertical fracture is
less than the overburden pressure. Vertical fractures are generally most
effective in the following situations:
▪ Closely spaced horizontal stratification exists in the producing reservoir,
▪ Fluid injection is aided by vertical distribution in the producing interval,
and
▪ Very deeply penetrating fractures are required.

Figure 8: Formation of vertical fractures

2.4.2 Horizontal Fractures:

The fracture orientation is perpendicular to the direction of the overburden


pressure (Pob), and the fracture gradient is 1.0 psi per ft of depth or greater.
The horizontal fracture is greater than overburden pressure and mostly occurs
at depths of 3000 ft or less.
Figure 9: Formation of horizontal fractures

Horizontal fractures are generally most effective where the following


conditions exist:

▪ Homogeneous formations are present.


▪ The oil reservoir is under-laid by an active bottom-water drive, or is
over-laid by a gas cap.
▪ Uniform injection of fluid into large areas of the reservoir is required.

2.5 Hydraulic Fracturing Models:

During this early period of hydraulic fracturing, two simple models were
proposed to try to predict the shape and size of a hydraulic fracture based on
the rock and fluid properties, the pumping parameters, and the in situ stresses
(Khristianovic and Zheltov, 1955; Geertsma and de Klerk, 1969; Perkins and
Kern, 1961; Nordgren, 1972). The models are known as the KGD and PKN
models, and their description can be found in the many other summaries or
texts (e.g. Geertsma, 1989; Mendelsohn, 1984a, b). Perkins and Kern (1961)
and Geertsma and de Klerk (1969) also derived a model for radial hydraulic
fracturing.

KGD and PKN models are essentially two dimensional plane strain formulations
with fluid flow only along the length (or radius) of the fracture. The fracture
width and shape are related to the fluid pressure distribution in the fracture;
the KGD model has a constant height and constant width through the height,
while the PKN model has a constant height and an elliptical vertical cross-
section.

The 2D models are not able to simulate both vertical and lateral propagation.
Therefore, pseudo-3D models were formulated by removing the assumption
of constant and uniform height (Settari and Cleary, 1986; Morales, 1989). The
height in the pseudo-3D models is a function of position along the fracture as
well as time. The major assumption is that the fracture length is much greater
than the height, and an important difference between the pseudo-3D and the
2D models is the addition of a vertical fluid flow component. The pseudo-3D
models have been used to model fractures through multiple rock layers with
differing stresses and properties. These models are simple, fast, and relatively
effective. Warpinski et al. (1994) recently provided brief descriptions and a
comparison of predictions for a number of simulators, including 2D and
pseudo-3D models.

Pseudo-3D models cannot handle fractures of arbitrary shape and orientation,


however; fully 3D models are required for this purpose. The literature contains
numerous references to fully 3D simulators; the majority of these are limited
to planar fracture surfaces, however. These are called planar-3D simulators in
this chapter to differentiate them from true fully 3D simulators which can
model out-of-plane fracture growth. Planar-3D simulators have been
developed by Clifton and Abou-Sayed (1979), Barree (1983), Touboul et al.
(1986), Morita et al. (1988), Advani et al. (1990), and Gu and Leung (1993).
Out of plane 3D hydraulic fracture growth has been modeled by Lam et al.
(1986), Vandamme and Jeffrey (1986), and Sousa et al. (1993). To the
authors’ knowledge only Carter et al. (1994), using the predecessor of the
simulator described herein, have modeled 3D fracture in the near-wellbore
region of a cased, perforated, and deviated wellbore.

The increasing use of deviated wellbores implies that fully 3D hydraulic


fracture simulators are vital to the petroleum industry. Hydrofracturing is
often less effective for deviated wellbores as compared to traditional vertical
wells. Some of the problems have been attributed to a poor understanding of
the mechanics of fracture initiation and propagation from a deviated wellbore.

The complex state of stress which is generated around an inclined wellbore


(Yew and Li, 1988; Ong and Roegiers, 1995) means that the fracture
propagates with a complex geometry (Behrmann and Elbel, 1990; Hallam and
Last, 1991; Weijers and de Pater, 1992; Abass et al., 1996).

The complex stress state and fracture geometry can limit the fracture width
at the wellbore and hinder the injection of proppant into the fracture leading
to premature screen out (Hallam and Last, 1991; Soliman et al. 1996).
Nevertheless, the advantages of drilling inclined wellbores are significant. For
example, the ability to drill several wells from a single location minimizes
production infrastructure and impact on the environment. Therefore, the
ability to model hydraulic fracturing from deviated wells is of ever increasing
importance.

In addition to inadequate modeling of the fracture geometry, many of the


current hydraulic Fracturing simulators do not predict the correct wellbore fluid
pressure or fracture geometry even for planar fractures The proposed reasons
for this are numerous (Medlin and Fitch, 1983; Warpinski, 1985;
Shlyapobersky et al., 1988; Jeffrey, 1989; Palmer and Veatch, 1990; Johnson
and Cleary, 1991; Gardner, 1992; Papanastasiou and Thiercelin, 1993; de
Pater et al., 1993 and van den Hoek et al., 1993).

2.5.1 Perkins and Kern Model of a Vertical Fracture (PKN):

Perkins and Kern (1961) assumed that a fixed height vertical fracture is
propagated in a well confined pay zone; 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 the conditions as shown in Figure (2.8), 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 (i.e., 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).
Figure 10: A PKN fracture model (Perkins and Kern (1961)).

2.5.2 Khristianovich-Geertsma-De Klerk Model (KGD):

Khristianovich and Zheltov (1955) derived a solution for the propagation of a


hydraulic fracture by assuming the width of the crack at any distance from the
well is independent of vertical position (i.e., a rectangular cross section with
slip at the upper and lower boundaries), which is a reasonable assumption for
a fracture with a height much greater than its length. Their solution includes
the fracture mechanics aspects of the fracture tip.

They recognized that to solve this problem analytically it was necessary to


simplify the solution. They did this by assuming that the flow rate in the
fracture is constant and that the pressure in the fracture could be
approximated by a constant pressure in the majority of the fracture body,
except for a small region near the tip with no fluid penetration, and hence no
fluid pressure. This assumption can be made because the pressure gradient
caused by fluid flow is highly sensitive to fracture width and therefore occurs
primarily in the tip region.
Figure 11: A KGD fracture model (Khristianovich and Zheltov (1955))

The concept of fluid lag remains an important element of the mechanics of the
fracture tip and has been validated at the field scale (Warpinski, 1985). They
showed that provided this dry region is quite small (a few percent of the total
length), the pressure in the main body of the fracture is nearly equal to the
pressure at the well over most of the length, with a sharp decrease near the
tip.

2.5.3 Three Dimensional & Pseudo Three Dimensional Models (3D):

The simple models discussed in the previous sections are limited because they
require the engineer to specify the fracture height or to assume that a radial
fracture will develop. This is a significant limitation, because it is not always
obvious from logs and other data where or whether the fracture will be
contained. Also, the fracture height usually varies from the well (where the
pressure is highest) to the tip of the fracture. This limitation can be remedied
by the use of planar 3D and pseudo-3D (P3D) models. The three major types
of hydraulic fracture models that include height growth are categorized
according to their major assumptions.
Planar 3D models are based on the assumption that the fracture is planar and
oriented perpendicular to the far-field minimum in-situ stress. No attempt is
made to account for complexities that result in deviations from this planar
behavior. Simulators based on such models are also computationally
demanding, so they are generally not used for routine designs.

They should be used where a significant portion of the fracture volume is


outside the zone where the fracture initiates or where there is more vertical
than horizontal fluid flow. Such cases typically arise when the stress in the
layers around the pay zone is similar to or lower than that within the pay.

P3D models attempt to capture the significant behavior of planar models


without the computational complexity. The two main types are referred to
here as “lumped” and cell-based. In the lumped (or elliptical) models, the
vertical profile of the fracture is assumed to consist of two half-ellipses joined
at the center as shown in Figure (2.10).

Figure 12: A Lumped fracture model

The horizontal length and wellbore vertical tip extensions are calculated at
each time step, and the assumed shape is matched to these positions.
2.6 Fracturing Fluids:

To select the proper fluid for a specific well it is necessary to understand the
properties of fluids. The fluid design must be considered these characteristics:

1. Low leak off rate.

2. The ability of the fluid to carry the propping agent.

3. Friction loss.

4. Fluid viscosity.

2.6.1 Types of Hydraulic Fracturing Fluids:

1. Water-Base fluids .
2. Oil-Base fluids:

Napalm gels ,viscous refined oil and lease crude oils.

3. Acid based fluids .


4. Foams .
5. Emulsions .

2.7 Propping Agents:

Propping agents are required to (prop-open) the fracture once the pumps are
shut down and the fracture begins to close. The ideal propping agent will be
strong, resistant to crushing, resistant to corrosion, have a low density, and
readily available at low cost. The products that best meet these desired traits
are silica sand, resin-coated sand, and ceramic proppant.
Chapter Three

Hydraulic Fracturing Design


Procedure
Hydraulic Fracturing Design Procedure:

3.1 Introduction:

Hydraulic fracturing is the process of using hydraulic pressure to create an


artificial fracture in a reservoir. The fracture grows in length, height and width
by pumping a mixture of fluid and proppant into the formation above fracture
pressure.

In this section will present step by step the main data required for hydraulic
fracture design and design procedure steps for hydraulic fracture in vertical
wells (Suggested by Dr: Mohamed Nasr).

3.2 Data Required for Hydraulic Fracturing Design:

Parameter value unit

Depth ft

Producing interval ft

Formation thickness ft

Average reservoir pressure (BHSP) psi

Average bulk density (from density log or mud log) gm/cc

Horizontal tensile strength of rock psi

Reservoir oil compressibility psi-1

Reservoir water compressibility psi-1

Reservoir Gas compressibility psi-1

Oil formation volume factor res.bbl/STB

Oil saturation %

Water saturation %

Gas saturation %
Formation porosity %

Original formation permeability md

Fracturing fluid viscosity (YF 140 HTD X-linked gel) cp

Fracturing fluid density (Versa Gel) Sp.Gr: (1.002) ppg

Reservoir oil viscosity cp

Area of filter medium cm2

Slope of fluid loss curve at lab. cm/min½

Filtration pressure at lab. psi

Casing outer diameter in

Wellbore diameter in

Drainage diameter ft

Proppant size and type (Carboliteproppant) mesh

Porosity of packed proppant %

Specific gravity of proppant unitless

Fracturing fluid spurt loss gal/ft2

Tubing inner diameter in

Tubing depth ft

Gas oil ratio scf/bbl

Bubble point pressure psi

Reservoir temperature (BHST) °F

Frictional pressure gradient inside tubing psi/ft

Perforation diameter in

Perforation discharge coefficient unitless

Number of perforations #
Additional Data:

Formation Type #

Matrix compressional sonic wave transient time (∆tma) µ sec

Matrix shear sonic wave transient time (∆tsma) µ sec

Sonic compressional wave transient time log (∆tc) µ sec

Sonic shear wave transient time log (∆ts) µ sec

Assume: hf=h, qi= bbl/min Vi= bbls

Type of Completion: Cased or Open Hole

3.3 Hydraulic Fracturing Design Procedure:

1. Calculation the Formation Fracturing Pressure:

▪ The average formation density (ρb):

As given from open hole logs data.

▪ Overburden pressure gradient (Gob):

Gob = 0.433  b

▪ The Poisson’s ratio ( ):


 2 
 1  t s  
 2   t  − 1 
=  c 
 2 
  t s  − 1 
  t  
  c 

▪ The Shear modulus (G):

  
G = 1.34  1010  b 
 2 
 t s 

▪ The Young’s modulus (E):

E = 2 G (1 + )

▪ Bulk modulus (kb):

 1 4 
kb = b   −   1.34  1010
 2 
 t c 3t2s 
The Bulk modulus also can be calculated from Young’s modulus by using
the following equation:
E
kb =
3  (1 − 2)

▪ Rock bulk compressibility with porosity (Cb):

1
Cb =
kb

▪ Rock grain compressibility with zero porosity (Cr):

1
Cr =
  4  
b   1 −  1.34  1010 
  2 2  
  tma 3tsma  
▪ Poroelastic constant (α):

 C 
 = 1 − r 
 Cb 

▪ Poroelastic stress coefficient (  ):

 (1 − 2 ) 
 =    
 2 (1 − ) 

❑ Calculation of Initiation Fracturing Pressure (Pf):

▪ The overburden pressure (Pob = σv):

Pob = Gob  D

▪ The minimum fracturing pressure (Pf):

  
2  (Pob − PP ) + To
 1− 
Pf = + PP
1 − 2  
2 −  
 1− 

❑ Calculation of Minimum Fracturing Pressure (Pf):

▪ The extension pressure (Pfrac = σh(min.)):

  
Pfrac. =    (v − PP ) + PP
1 −  

2. The Effective Fracturing Fluid Coefficient:

P (closure stress) = P f − P P

▪ The viscosity control coefficient (CV):


k  P
C V = 0.0469
 frac.

▪ The compressibility control coefficient (C C):

Total compressibility ct=co.So+cw.Sw+cg.Sg

k  ct
CC = 0.0374  P
 res.

▪ The wall building control coefficient (C W):

Pact.
mact. = mlab. 
Plab.

 0.0164 
CW =    mact.
 Af 

▪ The total control coefficient (CT):

1 1 1 1
= + +
CT CV CC CW

3. The Fracture Volume (Fracture Dimensions):

▪ Assume (qi = bbl/min), and (Vi = bbls):

▪ Pumping time (tp):

V 
tp =  i 
 qi 

▪ Calculate (α):
 16 C T  t 
= 
  W + 16 SP 
 w 

▪ Calculate (L):

 q  5.615    SP   2 2 
L= i  W + 16   
 7.48  e • erfc () + − 1
 64  h C2  12 w
 
f T      

▪ Calculate (A):

A = 4 L hf

▪ Assume (WW) and calculate (L) and (A):

WW α L A

0.05

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1.0
▪ Calculate the plane strain modulus (E’):

 E 
E' =  
 1 − 2 

▪ Calculate the fracture width (WW) by using KGD Model:

0.25
  qi L2 
WW = 0.345   
 E' hf 
 

▪ Assume (L/re) and calculate (WW) and (A):

L/re L WW A

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1.0
Figure 13: Fracture width vs. the corresponding fracture length values by
using KGD model

In order to solve for (L) and (Wf), the two nonlinear differential (L) and (Wf)
equations are solved simultaneously. The calculated (L) values are plotted
versus assumed (Wf) values on log-log graph paper. On the same graph paper,
assumed values of fracture length (L) are plotted versus calculated values of
log (Wf). The intersection of these curves represents a solution that satisfies
both length and width equations. Therefore, the fracture width and the
corresponding fracture length values are obtained from the intersection of
these curves.

Consequently;

▪ The fracture width (WW) and length (L) by using KGD Model are:

Fracture width (WW) = in.


Fracture length (L) = ft.
4. Calculation of Fracturing Efficiency (Eff):

▪ Calculate fracture volume (Vf):


Vf =   L WW hf
2

▪ The fracture efficiency (Eff):

V 
Eff =  f   100
 Vi 

5. The Concentration of Proppant in the Fracturing Fluid:

A. Calculation of Proppant Weight Needed (Wt Propp.):

▪ The weight of proppant needed (Wt):

wt = Vf  (1 −  )  ρprop.

▪ The proppant concentration (CPropp.):

W 
Cprop. =  t 
 Vi 

B. Calculation Fracturing Fluid Density including Proppant (ρmix.):

▪ Fracturing fluid density including proppant:

 8.34  + x 
ρ mix. =  
 1 + 0.0456 x 

C. Calculation Fracturing Fluid Injection Rate including Proppant (qi):

▪ Fracturing fluid density including proppant:


 wt of propant 
Vi +  
 prop.  62.4  5.615 
qi =  
tp

6. Calculation of Wellhead Injection Pressure (Pw):

▪ The hydrostatic pressure (∆PH.S):

Phydro = 0.052  mix.  D

▪ The pressure drop due to friction (∆Pfric.):

Pfric. = Gfric  D

▪ The pressure drop due to perforation (∆Pperf.):

Type of completion: Cased Hole.

2
 qi 
Pperf. = 0.237  mix.  2

 cp N dperf. 

▪ The wellhead pressure (Pw):

Pw = Pfrac. − Phydro + Pfric. + Pperf.

7. Calculation of Fracture Conductivity using Propped Fracture Permeability


Curves:

▪ The fracture permeability (kf):

kf = md,
▪ The fracture conductivity (FC):

FC = kf  w f

8. Calculation of Production Increase (McGuire & Sikora):

▪ Relative Conductivity (R.C):

kf  w f 40
Re lative Coductivity (R C) =
ke A

▪ Calculate reservoir area (A), assume square shape:

A = L2 = d2e

1,6402
A= = 61.75 = 62 acres
43,560

▪ Relative Conductivity (R.C):

kf  w f 40
Re lative Coductivity (R C) =
ke A

▪ Calculate ratio of (L/re):

L 180
= = 0.220
re 820

▪ Entering the production increase curve (McGuire & Sikora) with these
values:

Jf  7.13 
 
Jo  Ln(0.472 re / rw ) 
Figure 14: Fracture Conductivity using Propped Fracture Permeability
Curves (McGuire & Sikora).

9. The Well's Production Analysis (IPR Curves):

A. Production of IPR Curve before Fracturing:

▪ Productivity Index (PI=Jo) by using Darcy’s law:

7.08  10−3 k h (Pe − Pwf )


qo =
o Bo  Ln(re / rw ) 

Then;

7.08  10−3 k h
PI =
o Bo  Ln(re / rw ) 
While,

q
PI =
(Pe − Pwf )

Then,

q = PI  (Pe − Pwf)

▪ Assume bottom hole flowing pressure (Pwf):

Assumed Calculated
Pwf (psi) qo (BPD)
Pe 0
* *
* *
0 = AOFP

▪ Construction of tubing intake curve (using Brown correlations):

Assumed
Pth Pwf
qo (BPD)
(psi) (psi)
* * *
* * *
* * *

▪ From the intersection of (IPR) curve with the tubing intake curve (TPC):

qo = optimum before fracturing= BPD

Pwf = optimum before fracturing= psi


B. Production of IPR Curve after Fracturing:
▪ Productivity ratio (Jf / Jo):

Jf
= 3.33
Jo

Jf  qf   qe 
=    
Jo  Pe − P' wf   Pe − Pwf 

▪ Assume bottom hole flowing pressure (P’wf):

Assumed Calculated
P’wf (psi) qo (BPD)
Pe 0
* *
* *
* *
0 = AOFP
▪ From the intersection of (IPR) curve with the tubing intake curve (TPC):

qo = optimum after fracturing= BPD

P’wf = optimum after fracturing= psi

▪ Productivity ratio (production increase ratio) after fracturing:

qf q after
=
qo q before
Figure 15: Construct complete IPR curve

You might also like