+ All Categories
Home > Documents > Initiation and growth of a hydraulic fracture from a ...

Initiation and growth of a hydraulic fracture from a ...

Date post: 14-Mar-2022
Category:
Upload: others
View: 1 times
Download: 0 times
Share this document with a friend
38
1 Initiation and growth of a hydraulic fracture from a circular wellbore X. Zhang 1 , R. G. Jeffrey 1 , A. P. Bunger 1 and M. Thiercelin 2 1 CSIRO Earth Science and Resource Engineering, Private Bag 10, Clayton South, VIC 3169, Australia 2 Schlumberger RTC-Unconventional Gas, 14131 Midway Road, Suite 700, Addison, TX-75001, USA Submitted to Int. J. Rock Mech. Min. Sci. Abstract A two-dimensional (2D) model is presented for initiation and growth of one or more hydraulic fractures from a well that is aligned with either the maximum or intermediate principal in-situ stress. The coupling of fluid flow and rock deformation plays a key role in reorientation and pattern evolution of the fractures formed. After fracture initiation, the fracture can reorient as it extends from the wellbore until it becomes aligned with the preferred direction for fracture growth relative to the far-field stresses. Initiation and growth of multiple fractures are considered to study their interaction and competition with each other. In such cases, some fractures are unable to extend at all or they arrest after some limited growth, but others can grow in length relative to earlier developed fractures. For fractures that are driven by a uniformly distributed internal pressure, which implies injection of an inviscid fluid, fracture closure may occur at the portion of the fracture path adjacent to the wellbore. This local fracture closure does not typically occur when fluid viscous dissipation is introduced, but the local width is greatly reduced and a fluid lag zone develops. The reduced fracture width results in fluid viscous friction and an associated pressure drop near the wellbore, which makes fracture stimulation more expensive and less successful and may reduce well productivity. Initiation of a fracture using a viscous fluid and a higher injection rate causes the fracture to curve more gradually as it seeks to align with the maximum principal stress direction, a result from our model that is consistent with widely used tortuosity remedy methods. A dimensionless parameter is developed that is shown to characterise near wellbore reorientation and curving of hydraulic fractures driven by viscous
Transcript

1

Initiation and growth of a hydraulic fracture from a circular

wellbore

X. Zhang1, R. G. Jeffrey1, A. P. Bunger1 and M. Thiercelin2 1 CSIRO Earth Science and Resource Engineering, Private Bag 10, Clayton South, VIC 3169,

Australia 2 Schlumberger RTC-Unconventional Gas, 14131 Midway Road, Suite 700, Addison, TX-75001,

USA

Submitted to Int. J. Rock Mech. Min. Sci.

Abstract

A two-dimensional (2D) model is presented for initiation and growth of one or more hydraulic

fractures from a well that is aligned with either the maximum or intermediate principal in-situ

stress. The coupling of fluid flow and rock deformation plays a key role in reorientation and

pattern evolution of the fractures formed. After fracture initiation, the fracture can reorient as it

extends from the wellbore until it becomes aligned with the preferred direction for fracture

growth relative to the far-field stresses. Initiation and growth of multiple fractures are considered

to study their interaction and competition with each other. In such cases, some fractures are

unable to extend at all or they arrest after some limited growth, but others can grow in length

relative to earlier developed fractures. For fractures that are driven by a uniformly distributed

internal pressure, which implies injection of an inviscid fluid, fracture closure may occur at the

portion of the fracture path adjacent to the wellbore. This local fracture closure does not typically

occur when fluid viscous dissipation is introduced, but the local width is greatly reduced and a

fluid lag zone develops. The reduced fracture width results in fluid viscous friction and an

associated pressure drop near the wellbore, which makes fracture stimulation more expensive

and less successful and may reduce well productivity. Initiation of a fracture using a viscous

fluid and a higher injection rate causes the fracture to curve more gradually as it seeks to align

with the maximum principal stress direction, a result from our model that is consistent with

widely used tortuosity remedy methods. A dimensionless parameter is developed that is shown to

characterise near wellbore reorientation and curving of hydraulic fractures driven by viscous

2

fluid. For complex multiple fracture cases, the results demonstrate that the misalignment angle

and the number of initiated fractures are important in near-well fracture path selection.

Introduction

One primary role of hydraulic fracturing is to provide a high conductivity pathway along

which reservoir fluids can flow to the well. However, the development of the fractures and,

ultimately, the treatment effectiveness, depends on the evolution of the fracture path in the near-

wellbore region. Near the wellbore multiple subparallel fractures may be generated because of

fracture initiation, either from perforations positioned along and phased around the wellbore or

from natural flaws around the open hole. Perforations and flaws are not usually aligned with the

far-field stress direction. Fractures initiated from these defects will reorient themselves to a plane

perpendicular to the minimum far-field stress direction as they grow away from the wellbore

(Hamison and Fairhurst, 1970; Daneshy, 1973; El Rabaa, 1989; Weng, 1993; Soliman et al.

2008). The reorientation process gives rise to mode II and mode III stress intensity factors along

the fracture leading edge, which has been shown to produce fracture branching and

segmentation, in addition to curving (Daneshy, 1973; Pollard et al., 1982; Hallam and Last,

1990, Soliman et al. 2008). The overall result of such a fracture initiation process is the

generation of near-wellbore fracture tortuosity. As a consequence, the net pressure required to

extend the fractures is increased considerably. Methods to reduce near-wellbore tortuosity and

potential proppant placement problems have been developed that involve using higher viscosity

fluids and higher injection rates to initiate the fracture. Proppant slugs and oriented perforations

are also used in an effort to reduce the fracture complexity. In many naturally fractured

formations (coal seams, naturally fractured reservoirs and geothermal reservoirs), near-wellbore

tortuosity can lead to serious problems that significantly affect the success of hydraulic fracture

stimulation treatments (Cuderman et al., 1986; Chu et al., 1987; Chen, 2009). Methods to avoid

or reduce fracture tortuosity are therefore important. When the source of the tortuosity is

understood, methods to avoid its development can be designed. For example, Manrique and

Venkitaraman (2001) showed results demonstrating that, in wells where otherwise failed

treatments occurred, perforations oriented in the plane of the maximum principal stress direction

are an effective method of eliminating tortuosity and avoiding initiation of multiple fractures. We

present results that quantify the effect of treatment parameters and rock properties on the

development of near-wellbore fracture curvature. These results allow the engineer to design

improved breakdown methods to minimise the development of fracture tortuosity.

3

Hydraulic fracture initiation from a wellbore has been studied theoretically and

numerically, but still remains a subject of interest for petroleum industry research. In particular,

many details of how non-planar fracture geometries contribute to high treating pressures, often

observed in the field, are still lacking. The models should capture pressure loss in the small zone

associated with near-wellbore fracture tortuosity. Generation of near-wellbore fracture

complexity, with associated pinching and competition between multiple fracture branches can

occur simultaneously in the evolution of fracture paths. Daneshy (1973) considered the

conditions for fracture initiation and propagation from vertical and deviated wellbores to find

factors controlling near-wellbore effects. Narendran and Cleary (1983) used numerical methods

to analyse curving and interaction of multiple fractures. Weng (1993) and Yew et al. (1993)

numerically studied the interaction and linkage of multiple fractures in stress fields around a

deviated well, including fracture growth in a plane inclined to the wellbore axis. A finite-element

model for a wellbore and a single bi-wing fracture was presented by Cherny et al. (2009) with a

predetermined fracture trajectory. A quasi-static approach was used by them, as is the case for

most previous studies, and only stationary crack problems were studied, using different fixed

fracture lengths. In their paper, the fracture curving is predetermined to follow an arc with a

given radius. Although narrow fracture width in the near-wellbore region and abnormal fluid

pressure is predicted by their numerical modelling, the predetermined fracture path limits the

application of the results obtained. Moreover, asymmetric multiple fractures are commonly

observed in the field and in laboratory experiments (Rabaa, 1989; Hallam and Last, 1990; Yew

et al., 1993; Weng, 1993].

Numerical hydraulic fracturing models have been applied to the study of problems dealing

with competing fracture growth, including width reduction and associated increased viscous

frictional pressure loss. Based on such numerical analyses, various operational remedial

techniques have been proposed to overcome the near-wellbore effects, such as using higher

injection rate, higher viscosity fluids to reduce tortuosity and oriented perforations to control

fracture initiation orientation (Manrique and Venkitaraman, 2001; Abass, et al., 2009). In

particular, the fracture turning processes, during which the fracture reorients as it grows from a

wellbore to become perpendicular to the minimum principal stress direction, is the focus of this

paper. We recognized that the fracture turning can induce rotation of local principal stresses. The

curving fractures studied here are likely to be less smooth than modelled and may break up into

echelon cracks near the wellbore, as discussed by Pollard et al. (1982). However, formation of

segmented and echelon fractures is beyond the scope of this study, in which we focus on

formation of 2D, smoothly curving fractures.

4

In this paper, we will employ a 2D numerical method that can model hydraulic fracture

growth from a wellbore. The 2D plane strain assumption, used here, is justified if the height,

along the wellbore axis direction, of a slot or natural flaw, which is the site of initiation of the

hydraulic fracture, is relatively large compared to its radial depth. In addition, the wellbore axis

is assumed to be aligned with the maximum or intermediate principal stress direction and axial

longitudinal fractures would be created. In the following, we will first present comparisons to

published results to verify our model and then we will present results related to initiation and

propagation of single and multiple hydraulic fractures from the wellbore. The advantage of our

model is that it can explicitly address the coupling of rock deformation and viscous fluid flow

(Zhang et al., 2005, 2007, 2008, 2009). Energy dissipation related to viscous flow is important

for predicting hydraulic fracture propagation from flaws located along the wellbore wall. By

including viscous flow in the model, narrow portions of the fracture channel which arise from the

fracture curving and interactions with the wellbore, naturally lead to higher pressures and

pressure gradients. This non-uniform pressure distribution is shown to have a strong effect on the

fracture curving and overall path. Most previous studies assume the fluid has zero viscosity,

which leads to a uniform pressure that does not depend on the fracture channel width

distribution.

First we consider a single hydraulic fracture growing from a circular wellbore and develop

a nondimensional parameter whose value can be related to the fracture path. Next we study the

curvature of two or more fractures growing from the wellbore, to validate the proposed

dimensionless parameter. We vary the locations of the starter fractures and show how this affects

the hydraulic fracture growth path. We contrast these results with published results that used,

instead, a pre-determined fracture path. In general, the fracture path depends on the detailed

pressure and width distribution that arises from solution of the coupled problem, including the

non-planar fractures interacting with themselves, with the wellbore, and with other hydraulic

fractures. The solution provides, at each time step, the fracture path and opening and excess

pressure distributions. Although the model considers only 2D fractures, the results presented

provide useful insights into the formation and interaction of fracture propagation patterns in the

presence of a wellbore.

Problem Formulation

The 2D hydraulic fracture problem that we consider in this paper is shown in Fig. 1, which

represents a cross-section through a vertical well that is aligned with the maximum or

intermediate principal stress. Two flaws like perforated or abrasively jetted slots, serve as

5

initiation sites for the longitudinal hydraulic fractures that are driven by injected fluids. If the

perforated interval is long, or if the well is not cemented and cased, these axial longitudinal

fractures and their propagation within the transverse plane, as shown in Fig. 1, is assumed to be

independent of the fracture height and subject to plane-strain conditions. Deviation from the

direction of maximum tensile stress would cause the fractures to reorient themselves to form bi-

wing fractures, as depicted in Fig. 1. Of course, more than two fractures at different angles may

initiate in a homogeneous rock instead of just a single bi-wing fracture.

For the sake of simplicity, the in-situ stress field is given as in Fig. 1 based on the in-plane

maximum H and minimum h principal stresses. The rock is assumed to be impermeable and

the stresses can be considered as total far-field stress values. The well radius is denoted as R and

the initial fracture length L (which is shown exaggerated in Fig. 1) is assumed to be very small so

that the effect of initial fracture length on final fracture path can be neglected. Each fracture is

labeled by its initiation orientation angle with the respect to the x-axis, which is the in-plane

maximum principal stress direction. Fractures that initiate at =0 would propagate as straight

planar fractures along the x-axis direction. The well radius is chosen as 0.1 m for the cases

considered in this paper. The wellbore pressure is denoted as wP , which will be determines as

part of the solution. It should be noted that the wellbore pressure is the same as the entrance

pressure for each fracture.

The properties of the rock, fluid and fractures are as follows. The elastic properties for the

intact rock are Young’s modulus (E), Poisson’s ratio ( ) and the mode I fracture toughness

( ICK ). The injected fluid is incompressible and has a Newtonian dynamic viscosity given by .

The sum of all injection rates of fracture branches growing from the well is assumed to be

constant and is denoted as inQ . Although the total rate injected, inQ , is specified as constant, the

rate into each fracture can be different and can vary with time to allow for different growth rates

of each fracture.

Near the wellbore, the two sides of a fracture may come into contact. The model allows for

contact stresses to develop and uses a cohesionless Coulomb friction criterion to assess the

frictional strength of surfaces in contact, (Jaeger et al., 2007)

( )n fp (1)

where is the shear stress along the closed fractures, n is the normal stress acting across them,

fp is the fluid pressure and is the coefficient of friction.

6

In addition to the tangential displacement discontinuity (DD), v, associated with frictional

slip along the closed fracture segments or shear displacement in general, there exists opening

displacement w along the opened hydraulic fractures, which contributes to fracture conductivity.

All of these elastic and plastic displacements give rise to changes in the rock stress or changes in

the tractions acting along the fracture surfaces. The governing equations for stresses, fluid

pressure and displacements are given in Zhang et al. (2007, 2009), and are given in terms of

Green’s functions. We employ the displacement discontinuity method (DDM) (Crouch and

Starfield, 1990) for the simulation of rock deformation in a 2D homogeneous and isotropic

elastic material, on which a uniform stress (e.g. H does not vary with y) is applied at infinity.

The governing equations for fluid flow in the fracture channel are given by Zhang et al.

(2009). There are two types of fluid movement along fractures. The fluid can diffuse along the

closed fractures without any mechanical opening of the fracture, or it can flow through an open

channel as a result of opening mode hydraulic fracturing processes. In the latter case, the fluid

pressure is balanced by the compressive stress caused by in situ stress and rock deformation, that

is, n fp . We use Reynolds’ equation to describe fluid movement inside the opened hydraulic

fractures:

3( ) ( ) in which fpw q w

qt s s

(2)

where w is the mechanical opening and vanishes for the closed segments, is the pre-existing

hydraulic aperture arising from surface roughness and microstructures, s is a distance measured

along the fracture length, q is the fluid flux and 12 .

However, it must be mentioned that effective stress changes in the fluid-infiltrated and

pressurised fracture portion can produce changes of the hydraulic aperture without fully opening

a fracture (Walsh and Grosenbaugh, 1979; Witherspoon et al., 1980; Brown and Scholz, 1986).

The dilatation can slightly affect the internal pressure distributions since the resulting fluid

conductivity varies in location and time. As stated above, the hydraulic aperture at the beginning

is assigned with an initial value 0w for each closed fracture segment. The evolution of obeys

a nonlinear spring model in response to any increments of the internal pressure. In particular, the

equation governing the hydraulic aperture change associated with pressure change is given as

follows,

fdPd / (3)

where is a small constant with a value of 810 /MPa in the model. A pressure diffusion

7

equation is applied to calculate the pressure inside a closed natural fracture (Zhang et al., 2009)

0)( 2

s

p

sc

t

p ff (4)

where )12/(1 c . The associated flow rate can be obtained by Darcy’s law. This calculation

applies to all closed portions of fractures, whether they are undergoing or have undergone slip or

not.

To complete the problem formulation, we need boundary and initial conditions for the

above governing equations. At the beginning, all fracture segments are assumed to be evacuated

and the rock mass is stationary. A rock failure criterion is needed, as is a method for handling

fracture intersection and associated fluid flux redistribution. For the fractures connected to the

wellbore, the sum of the fluid flux entering them is equal to the injection rate and the fluid

pressure continuity is enforced, that is,

1

(0, ) and (0, )N

ii in f w

i

q t Q p t P

(5)

where N is the number of starter fractures and (0, )iq t the fluid volumetric flux into a fracture i

and (0, )ifp t is the fluid pressure at the mouth of the fracture.

At the fracture tip, the opening and shearing DDs are zero, that is,

0)()( lvlw (6)

and the fracture opening profile near the crack tip possesses a square root shape with distance

from the tip as predicted by Linear Elastic Fracture Mechanics (LEFM) theory. The method

presented here approximates the viscosity dominated hydraulic fracture tip conditions where

energy dissipated in fluid transport becomes much larger than fracture energy (Detournay, 2004;

Bunger and Detournay, 2008), by an approach that explicitly calculates the fluid lag size and

relies on discretization to capture rapid changes in fluid pressure near the tip.

When the stress intensity factor at the fracture tip reaches the fracture toughness of the

rock, fracture growth occurs. In general, the fracture propagation direction may change as

dictated by the near-tip stress field characterised by the Stress Intensity Factors (SIFs). In

particular, the mixed-mode fracture criterion proposed by Erdogan and Sih (1963) is used for

determining fracture growth direction in the model. This same approach has been used by a

number of researchers (e.g. Mogilevskaya et al. 2000b). It must be noted that the calculated

crack deflection angle varies strongly at the near-well region. To smooth the fracture pathways, a

relaxation method is used in reducing the possibility in creating zig-zaging paths. In particular, if

8

the crack deflection angles for two consecutive growth steps have different sign, the new angle is

reduced by the relaxation factor.

Numerical Scheme and Verification

The coupled fluid flow (to obtain pressure) and rock deformation (to obtain fracture

opening) problem is solved in an implicit manner by incorporating the elasticity equation into the

Reynolds equation. The responses obtained are history dependent because of the irreversible

fluid flow and frictional sliding processes. Additionally, there are three moving fronts inherent in

the problem, namely the slip front, the fluid front and the fracture opening (crack tip) front. The

reader is referred to Zhang et al. (2007, 2009) for a detailed description of the numerical solution

methods applied in modelling these processes. At the wellbore, the fracture entry pressures must

be equal to the wellbore pressure. This condition is enforced through an iteration scheme. The

fracture entry and wellbore pressures are constrained to be equal by enforcing Eqn. (5). The

injection rate into each fracture is found based on the specified current wellbore pressure, using

the method described in our previous work applied to fracture junctions (Zhang et al., 2007). In

the calculation, the convergence of the solution is very sensitive to the time step and the

convergence error tolerance. To rapidly find a converged wellbore pressure, a relaxation factor is

used in updating the wellbore pressure at each iteration step. A tolerance (< 410 ) is set for the

relative error between the wellbore and fracture entry pressures. Therefore, in the program, a

new iteration loop is introduced for obtaining a converged wellbore and fracture entry pressure

solution.

The numerical method has been verified, in our previous papers, by comparison to

published solutions to a range of problems; see Zhang and Jeffrey (2009). However, to treat

problems that include a wellbore, the program was modified to allow correct treatment of the

interior and exterior problems generated when the wellbore is discretized using the DD

(Displacement Discontinuity) method. The resulting DDs along the wellbore boundary are

fictitious, although they are necessary for calculating pressure and displacement at the wellbore

wall and along the fractures. The fictitious DDs do not have a literal physical interpretation as

DDs, but rather are an approach to imposing the correct pressure condition at the wellbore

boundary. Moreover, the interior wellbore disc, if not constrained, is free to move as a rigid body

and therefore the rigid body motion of the disc must be suppressed when applying a DDM to

such problems (Crouch and Starfield, 1983).

A simple method to suppress these rigid-body motions is as follows. Two additional

9

elements are defined inside the wellbore and assigned fictitious DDs. Then we calculate the

displacement on the negative side of an element based on the fundamental elastic solutions

(Crouch and Starfield, 1983). After that, the four displacements at these two additional elements

are set to zero in order to construct a closed, well-posed problem for all DDs. The displacements

on the negative side of the elements are reproduced here as given by Crouch and Starfield (1983,

Eqn. 5.6.1) for both normal and tangential displacement components.

The elasticity equations for the normal and tangential DDs ( ku includes w and v.) at the

middle of each element of the system are

,ijk k f ij ijK u P k n s along the wellbore and the fractures (7)

0 ,ijk kM u k n s for the two additional elements (8)

where ijkK and ijkM are the coefficient matrices derived from the fundamental elastic solution as

is done in conventional boundary element methods (see details in Crouch and Starfield, 1983);

and ij are the resultant stresses along the fracture surface arising from far-field (in situ) stresses.

Here, Eqn (7) enforces equilibrium at DD elements along the fractures and the wellbore, and Eqn

(8) enforces the constraint condition that eliminates rigid body motion of the interior disc by

requiring zero displacements at two additional interior elements. The above simultaneous

equations (Eqns (7) and (8)) are solved for normal and shear DDs of all elements along the

wellbore, the fractures and of the two additional elements, once the fluid pressure is known. The

DDs that are obtained can then be used to calculate stress or displacement at any other point

inside the rock mass. It must be mentioned that the additional discretization along the wellbore

can slow down the computation speed and requires more intense calculation. Moreover, it is

found that the model is not efficient to treated extremely high injection rate and fluid viscosity

problems since the fracture path is very sensitive to the wellbore pressure development.

Figure 2 displays the comparisons in terms of normalised stress intensity factors (SIFs) for

two symmetric cracks emanating from a wellbore under different loading conditions. Both

fractures are of equal length and the loading conditions are illustrated in inset drawings. These

two cases have been thoroughly studied by Newman (1971), Atkinson and Thiercelin (1993) and

Mogilevskaya et al. (2000a). These papers all show very similar results for this problem so we,

therefore, only use the results obtained by Newman (1971) as the reference values. In Fig. 2, our

computed numerical results provide a good approximation to those obtained by Newman (1971).

The small differences that still exist have a maximum relative error of less than 5 precent for the

data in Fig. 2. Considering the sensitivity of SIF to element size, the results are taken as

10

acceptable.

Results for a case which applies a uniform pressure distributed along both the curving

fracture and the wellbore are contained in Fig. 3. A uniform pressure condition is equivalent to

using an inviscid fluid (Detournay, 2004). The pressure magnitude is adjusted at each fracture

growth step to satisfy the fracture extension criterion at the fracture tip. For generality of results,

Mogilevskaya et al. (2000b) demonstrated that the fracture path is controlled by the following

dimensionless parameter,

( )H h

IC

R

K

(9)

In Fig. 3, the coordinates have been normalised by the wellbore radius. The path given in

Fig. 3 by the densely spaced red square symbols are from our DD model and corresponds

approximately to the case of =0.25. The sparse green triangles are the results obtained by

Mogilevskaya et al. (2000b) for the same value. The initial crack, normal to the wellbore, is at

an angle, which is dictated by the wellbore discetisation, of 80 degree with respect to the x-axis

for our numerical result, and 81 degree for the results in Mogilevskaya et al. (2000b). A small

discrepancy in angles does not significantly affect the overall responses. As stated by

Mogilevskaya et al. (2000b), the crack growth, in addition to the dimensionless parameter ,

also depends on the initial fracture length. The initial fracture length is small (equal of 6% of the

wellbore radius) to minimise its influence. Bunger et al. (2010) determined that the initial

fracture length’s influence on breakdown pressure cannot be neglected if it is larger than 4% of

R. Nonetheless, for the purpose of the present calculation of crack paths, its influence is not

significant. For the case depicted in Fig. 3, fracture surface contact at the fracture inlet does not

occur. It is clear that for this small value of our computed path matches well with the path

obtained by Mogilevskaya et al. (2000b).

Uniformly Pressurised Fractures

In the above, we have reported results for uniformly pressurised fractures for small values

of . In this section, a special case will be discussed using medium values of to highlight the

effect of fluid viscosity and injection rate on the fracture path. The intermediate stress state

employed here is shown in Fig. 4. For the case of = 4, the strong compressive normal stress

acting near the wellbore can cause the fracture surfaces to be in contact near the fracture mouth

(for example, see the geometrical configuration shown in Fig. 4 with =4). As is

11

proportional to the difference between H and h , the larger this difference, the stronger the

tendency for pinching of the fracture. The same conclusion has been drawn by Cherny et al.

(2009). The opening DD distributions are displayed in Fig. 4, too, and the area with zero opening

DD corresponds to the contact zone. However, such contact conditions were not taken into

account in Mogilevskaya et al. (2000b) and, instead, this portion of the fracture was allowed to

overlap in their model. In Fig. 4(a), the purple line shows our results without any contact

constraint. The crack path without the surface contact restriction follows the trajectory obtained

by Mogilevskaya et al. (2000b), although there is some small discrepancy between two curves.

We checked if this was caused by the different initial fracture lengths. However, changing the

initial flaw length did not affect the results or the path away from the wellbore. In the

calculations, we have already adopted a very fine mesh and we did not see any crack path

zigzagging for this case.

More importantly, from the results shown in Fig.4, it is clear that the surface contact, or

lack of it, plays an important role in fracture development around a wellbore. The fracture

trajectory with contact is less curved and does not approach the y-axis, running through the

center of the wellbore, as rapidly as the results from Mogilevskaya et al. (2000b). Compared with

the results in Fig. 4(a), the fracture offset relative to the horizontal centreline is clearly larger if

contact is taken into account. Also, applying a coefficient of friction, , to the contacting

surfaces can increase this offset a little bit more as shown in Fig. 4(b). The narrow fracture

opening channel near the wellbore will directly increase the fluid injection impedance, increasing

the injection pressure and making it more difficult to carry out the fracture treatment. Hence the

contact stresses near the wellbore, even though the contact area is relatively small compared with

the whole fracture, can significantly affect the fracture growth and pressure. Generally speaking,

the fracture trajectories for uniformly pressurised crack pathways are determined by the values of

both and , but they are more sensitive to than to .

Effects of Fluid Viscosity

In the above section, all fractures are assumed to be uniformly pressurised as implied by a

toughness dominated condition. This is not the case in general for hydraulic fractures. Viscous

dissipation arising from the fluid flowing in the fracture channel is typically very important. The

viscous fluid flow will result in a pressure gradient along the fracture length and the non-constant

pressure distribution leads to different fracture opening. For example, contact between the two

sides of the fracture as discussed above, become much less likely when viscous flow is

12

considered. The resulting pressure distribution is strongly affected by narrow sections of the flow

path, such as those that develop near segments that come into contact for the uniform pressure

cases. For the modelling undertaken here, we do not considered the fluid compressibility and the

fluid volume change caused by changes in wellbore pressure and shape. A constant total

injection rate is specified for fluid flow into the wellbore. In addition, a small initial fracture size

(6% of wellbore radius) is used as a starting point for all fractures emanating from the wellbore.

The effect of different initial fracture lengths is thus not considered in this paper. Therefore, our

model fully accounts for the effect of the wellbore on stress field changes, but it does not fully

account for the effect of the wellbore on the fluid flow (from compressibility induced flow

during the fracture initiation process, for example, when pressure is changing rapidly). For an

injection rate of 0.0004 m2/s, for instances, if the fracture height is 1 m, the given injection rate is

equivalent to 24 litres per minute.

1. Single fracture cases

For this particular case, we employed the following material parameters: Young’s modulus

E= 50 GPa, ICK =1.0 MPa m . Under the same far-field stress conditions and the

geometric setting as given in Fig. 4, the hydraulic fracture path is found using a fluid viscosity

=0.01 Pa s and an injection rate of 0.0004 m2/s. The initial fracture angle is 80 degrees with

respect to the positive x-axis. For these conditions, a significant pressure gradient is generated

along the fracture path as shown in Fig. 5(b). The fluid pressure at the fracture mouth is over 20

MPa. This large wellbore pressure induces tensile hoop stress along the boundary of the

wellbore. The large fluid pressure is not only the driving force for crack growth, but is also a

factor in preventing fracture contact at the fracture entry. This lack of contact is indicated in Fig.

5(a) by the non-zero DDs everywhere along the fracture. The fracture opening in the portion

adjacent to wellbore is, however, very small, although a high pressure exists there. As the

hydraulic fracture grows in length, the near-wellbore fracture width increases and we expect that

proppant could then be placed through the restricted near wellbore width.

The fully opened fracture case, with no fracture contact occurring near the wellbore,

generates a fracture path with less curvature developing as the fracture propagates away from the

wellbore, in contrast to the more curved fractures for uniform pressure case shown in Fig. 4.

When viscous fluid flow is included, the fracture turning radius increases and the opening

distribution become smoother. This result supports the use of high viscosity fluid for initiation of

hydraulic fractures as a means of reducing fracture tortuosity (curvature in this case). Although

13

eventually the fracture becomes aligned in the direction of the maximum principal stress, the

injection pressure cannot decrease below a level sufficient to keep the whole fracture open (e.g.

otherwise a pinching point may develop near the wellbore). Incipient pinching at the wellbore

will result in elevation of the pressure, which would in turn prevent the pinch from developing.

Thus, use of more viscous fluids can produce a less curved fracture but continued propagation of

these fractures will still require elevated injection pressures when compared to the propagation

pressure for a straight planar fracture.

The wellbore pressure will increase, when a constant injection rate is imposed, until the

fracture is opened and propagated. The pressure will increase more slowly and eventually

decrease as the plane strain 2D fracture grows to become long in size compared to the wellbore

diameter. The non-planar nature of the fracture path and the associated restriction in fracture

width in the portion near the wellbore causes the wellbore pressure to remain at a higher level,

compared with the pressure to extend a planar fracture. The interaction between higher wellbore

pressure and in-situ stresses can generate a larger tensile hoop stress along the x-axis direction,

as shown in Fig. 5(c). We note that the hoop stress at the point on the x-axis nearly perpendicular

to the crack initiation site is tensile (positive), as shown in Fig. 5(c). The wellbore may thus

break down again, in a direction parallel to the x axis (the maximum in-situ stress direction), if a

critical stress level is reached for this point. The curved fracture would then likely stop growing

since the pressure to extend the favourably oriented fracture would be considerably lower.

Moreover, a fluid lag zone can be seen near the tip of the curved fracture shown in Figs.

5(b) and 5(d). The lag is a small region near the tip that is open but not pressurised by the fluid.

The existence of a fluid lag, in this case, reflects a larger pressure gradient, which can cause the

fracture physical tip to grow beyond the fluid front and is an important feature of viscous fluid-

driven fracture growth. Also, the lag zone appears to expand with the fracture growth since it is

smaller at the earlier time shown in Fig. 5(d).

To characterize the effect of fluid viscous dissipation on fracture paths, we introduce

another parameter which provides an apparent fracture toughness arising from viscous

dissipation, as given in Jeffrey(1989), based on an earlier somewhat different formulation given

by Settari and Price (1984), and Detournay (2004) in the case of zero lag. This apparent

toughness, when substituted for the rock fracture toughness in Eqn. 9 provides a dimensionless

parameter, given by Eqn. 10, that we hypothesize will control the curving of a hydraulic fracture

driven by viscous fluid as it grows from a wellbore.

14

3 1/4

( )

( )H h

Fin

R

Q E

(10)

In Eqn. 10, E is the plane-strain modulus and other parameters have been previously defined. In

this formula, the product of injection rate inQ and fluid viscosity can be controlled by a

fracturing engineer to directly change the tendency of the fracture to curve.

Figure 6 shows the fracture trajectories for various values of the product Qinµ with other

parameters fixed to values used for Fig. 5. It is found that as this product increases, i.e.,

decreasing F , the fracture path is less curved. This means that the fracture radius of curvature

becomes large and fracture paths are less tortuous. The first two cases listed in Fig. 6 use

different Qin and µ values but keep the value of Qinµ the same. The corresponding value of F is

equal to 0.44 for both cases. The fracture paths produced are coincident for these two cases, as

shown in Fig. 6. This indicates that this product is the controlling factor to characterize fluid

viscous dissipation for fixed values of E . For higher values of the product Qmµ, a higher

wellbore pressure develops and this can leads to unstable fracture growth and crack path

zigzagging in the model with associated difficulties in numerical accuracy. However, the trend

for the fracture path to become less curved as the product is increased is evident in the results

given in Fig. 6. Moreover, if we simultaneously vary Qin, µ, the modulus and in situ stresses to

keep the value of F unchanged, the fracture pathway will follow the curves for this value

of F , as shown in Fig.6. This highlights the importance of the dimensionless parameter in

characterizing the effect of viscous flow on fracture interactions with a wellbore. Also, the

results imply that the fracture radius of curvature becomes large and fracture paths are flatter if

the pumping rate is increased or the ratio of the maximum to the minimum horizontal stresses is

smaller. This conclusion is consistent with previous experimental studies by El Rabaa (1989),

Veeken et al. (1989) and Soliman et al. (2008).

2. Two fracture cases

When two fractures are arranged in a configuration as shown in Fig. 7, the normal and

shear stresses acting on both fractures are identical in both magnitude and direction. Therefore a

single bi-wing fracture path will be formed as two wings of the fracture propagate anti-

symmetrically with respect to both the x- and y-axes. For this and following subsections, the

parameters that remain fixed are Young’s modulus E= 65 GPa, Poisson’s ratio and

fracture toughness ICK =1.35 MPa m . For the following example, the fluid viscosity is 0.01

15

Pa s and the injection rate is 0.0004 m2/s. The alignment angle is 150° and 330° for both

fractures in Fig. 7(a) and 120° and 300° for the case shown in Fig. 7(b). The former case is called

as low-angle case since the interior angle between the fracture and the horizontal stress is 30°,

while the latter is called a high-angle case with an interior angle of 60°. The far-field stresses are

increased to 50 and 80 MPa for the in-plane vertical and horizontal principal stresses,

respectively. The fracture path is different for the two cases shown with the low-angle fractures

in Fig. 7(a) growing back toward the centre line of the wellbore. The two fracture wings tend to

eventually both grow parallel and along the centreline at y=0, while for the high-angle fractures

in Fig. 7(b), there is an offset left between the two wings on either side of the wellbore, which is

about equal to the wellbore diameter. The fracture growth path generated by the high-angle

fractures resembles the fracture geometry used by Weng (1993) and Cherny et al. (2009).

The net wellbore pressure, which is the difference between the wellbore pressure and the

far-field minimum principal stress, is initially over 40 MPa for the high-angle case shown in Fig.

7(c) and later in time this net pressure is approximately reduced to 25 MPa. This very large net

pressure arises because of the growth of the fracture with an orientation not aligned

perpendicular to the minimum stress and because of the restriction in fracture width near the

wellbore resulting from the interaction of the fracture with the wellbore. An increase in the

initiation angle results in an increase in the injection pressure. There is a strong pressure drop

post fracture initiation, after which the wellbore pressure tends to be more constant with a slight

increasing trend. The pressure is maintained at a higher level throughout the short time

simulated. However, if we compare the fluid pressure distribution along the fractures at the same

time for two different initiation angles as shown in Fig. 7(a) and (b), the pressure level along

most of the fracture extent is the same. The strong pressure drop occurs only in the vicinity of the

wellbore, and is associated with the local fracture width reduction there and with the higher

normal stress acting across the fracture at this location. This phenomenon of friction pressure

loss with near-wellbore localisation has been reported in Weng (1993) although different near-

wellbore fracture paths were produced by his numerical modelling, in part because he used a

uniform pressure to drive the fracture. It is reasonable to conclude that the fracture length for the

two different initial angles will be similar, too, and the fracture growth outside the zone of

influence of the wellbore appears to depend on parameters which control growth of a fracture

without a wellbore, although there is a high friction loss near the wellbore. Moreover, the

pressure in the fracture away from the wellbore is only slightly above the far-field minimum

stress magnitude as shown in Fig. 7(a) and 7(b).

16

The comparison of fracture trajectories in Fig. 7(d) further indicates that the fracture

pathways are determined by the dimensionless parameter F , for the case where viscous

dissipation is important. One can see the two pathways in Fig. 7(d) are exactly the same.

3. Multiple fracture cases

In the presence of multiple fractures, some starter fractures cannot propagate because of

local stress states at their initiation sites. The fractures subject to strong compression are

suppressed and starter fractures subject to low confinement extend. In Fig. 8, four starter

fractures are arranged every 90 degrees around the wellbore with angles of 60°, 150°, 240° and

330°. Despite the additional flaws along the wellbore, a bi-wing fracture geometry is generated

for this fracture configuration. Two of the four starter fractures grow significantly and become

dominated, while the other two fail to extend from the initial flaw when subject to the pressure in

the wellbore. Clearly, fracture initiation is sensitive to the local stress state at the site of the flaw.

The time-dependent fracture path development, as calculated by the model, is shown at two

times in Figs. 9(a) and (b) for cases that include five starter fractures with a special angular

arrangement. A single fracture is dominant at early time (Fig. 9(a)) but the pattern eventually

develops into a system with two dominant asymmetric fractures on both sides of the wellbore

shown in Fig. 9(b). The right-hand side fracture is dominant at the beginning because the stress

at this starter fracture location allows growth at a lower pressure. The left-hand side fracture,

which is in a slightly less favourable initiation direction, then starts to grow and extends to a

similar length as the right fracture, eventually. This emphasizes the interaction among two

fractures and the wellbore. When one fracture at higher internal pressure becomes longer, the

compressive stresses across other fractures, for the geometry studied, is affected in a way that

favours growth of a second fracture. In general, growth of other starter fractures may occur as a

result of the extension of one or more dominant fractures. The growth rate of the second fracture

may then increase, as demonstrated by comparing Figs. 9(a) and (b). When a second starter

fracture starts to grow, more fluid will enter this new fracture as its growth is favoured under the

current near wellbore stress field. Less fluid will then be available to extend the first or early-

growing fracture with the result that the second fracture to initiate can grow to become

approximately equal in size to the first, as shown in Fig. 9(b). After the fracture reorientation of

the newly activated fractures is completed and the second fracture has grown to a size

comparable to the wellbore diameter and to the first fracture, there will be another balance

established in fracture growth and fluid flow among these fractures all approximately aligned in

17

the preferred direction. Eventually, two dominant fractures are extended, in a fashion similar to

the single bi-wing fracture lying in the preferred direction. Therefore, a slight deviation from the

preferred direction does not greatly affect long term fracture growth. Similar fracture growth

paths, with one following the preferred direction and a secondary one turning quickly into this

direction have been found in experimental results (Abass et al., 2009).

With an increase in fluid viscosity, the fracture turning radius becomes larger as shown in

Fig. 9(c). The left-hand side fracture is reoriented in a smoother way or in a path with a lower

curvature. The path followed by the left-hand side fracture should then result in less viscous

friction loss near the wellbore. It is of interest to note that the right-hand side fracture path is

insensitive to the fluid viscosity change since it already lies in the most preferred direction for

fracture growth. Additionally, since the higher-viscosity fluid results in a higher fluid pressure

and wider fracture opening, the fracture growth becomes slower for both wings when compared

to the growth using lower viscosity fluids. This leads to a shorter fracture for the more viscous

fluid case, a well known result for planar fractures, as shown in Fig. 9(c).

Next, we consider six starter fractures arranged around the wellbore with an uneven

angular distribution with respect to the x-axis of 30°, 120°, 150°, 210°, 240° and 330°, as shown

in Fig. 10. The fractures at angles of 120° and 240° are in a more compressive initial stress state,

which causes their growth to be completely suppressed compared with other four fractures. The

remaining four starter fractures are all subjected to the same stress magnitudes, but the shear

stresses have different directions. The starter fracture at angle 150° has a slightly longer initial

fracture length, which causes it to extend first. As all the fractures grow and turn to align with the

horizontal x-direction, they grow nearly symmetrically at the early time because the strongest

interaction is between each fracture and the wellbore. But when they reach a certain length, the

fractures at the angle of 150° and 330° become more dominant, and form a bi-wing fracture

configuration. However, the other two fractures, at angle 30° and 210°, continue to extend but

more slowly and to a shorter distance. These two fractures propagate towards the longer fractures

and develop a kink as they enter the region where they interact with the longer bi-wing fractures,

as shown in Fig. 10. The potential for these short fractures to link into the longer fractures is also

clear. The slowly growing and partially suppressed fractures propagate towards the dominant

growing fractures and may link-up with them as previously found by Weng (1993). These results

suggest that initiation of several fractures from a wellbore subject to a larger stress ratio favours

interactions leading to link-up as seen in Fig. 10. After link-up, the connection pathway has a

step-like geometry but the shorter fractures may be suppressed before or after link-up. Hallam

and Last (1990) argued that the steps observed by Daneshy (1973) and El Rabaa (1989) can be

18

attributed to the interaction of starter fractures, instead of to growth of fractures in shear. This

issue needs further work as steps in the fracture path can exert a strong effect on pressure and

growth (Jeffrey et al., 2009; Zhang et al. 2009).

Conclusions

A two-dimensional plane-strain analysis of some aspects of near-wellbore fracture

tortuosity has been carried out based on a coupled hydraulic fracture model for rock deformation

and fluid flow. In contrast to quasi-static approaches used in previous studies, the model used

here provides a fully coupled solution for the evolution of the multiple fractures emanating from

the wellbore. The fluid pressure and the fracture conductivity are found for each time step and

the fracture propagation paths are found as part of the solution rather than being pre-determined.

The advantages in including fluid viscous friction loss effects are clear, compared with other

models.

When multiple fractures are initiated and then reorient to the maximum in-situ stress

direction, they interact with each other and with the wellbore. As a result, some fractures are not

able to extend at all or they arrest or close after some limited growth, but others grow more

quickly to eventually exceed the size of the earlier developed initially longer fractures. The

model captures this competition mechanism, considering growth of up to six starter fractures.

The direct consequence of such interactions is fracture width restriction and reorientation

resulting in increased treating pressure when compared to the pressure to extend a simple bi-

wing planer fracture.

For fractures that extend in a uniformly pressurised or toughness dominated regime,

fracture closure may occur along the portion of the fracture path adjacent to the wellbore. By

using a fully coupled hydraulic fracture model that includes viscous fluid flow, we have clarified

some contradictions in the literature that occur because of fracture surface contact was neglected.

Local fracture closure does not occur when fluid viscous dissipation is introduced, but the local

width is greatly reduced. Our fully coupled modelling predicts fluid friction losses and pressure

gradients near the wellbore that can be large, reflecting a high injection impedance condition.

Initiation of a fracture using a viscous fluid injected at a higher rate results in a fracture with a

more gradual curving path as it re-orients to align with maximum principal stress direction. Thus,

these numerical results support the use of more viscous fluid and higher injection rates in

initiating hydraulic fractures as a method to reduce near wellbore fracture tortuosity. In addition,

we have extended the dimensionless parameter (Mogilevskaya et al., 2000b) that characterizes

reorientation of a uniformly pressurized fracture near a wellbore to the more realistic case where

19

the hydraulic fracture is growing in a regime dominated by viscous dissipation. This new

dimensionless parameter together with the one presented by Mogilevskaya et al. (2000b)

characterizes the single fracture turning mechanisms and the formation of single bi-wing fracture

paths. Therefore, these dimensionless parameters quantify the degree of reorientation expected in

terms of physical parameters that can be measured and/or controlled.

The results demonstrate that the misalignment angle and the number of fractures are also

important in determining the fracture path near the wellbore. These two factors can significantly

affect the wellbore pressure magnitudes and thus the fracture trajectories. When a fracture is at a

favourable position in response to the dominant fracture growth, it can grow more rapidly

because of the favourable stress conditions that are generated by the dominant fracture, although

its early growth may have been suppressed by the far-field stresses. Growth of multiple hydraulic

fractures from the wellbore is complicated, but it appears that a single bi-wing fracture geometry

would be a final pattern for most cases studied.

In this paper, development of curving fractures with associated width reductions near the

wellbore is dealt with. However, the out-of-plane fracture twisting, which requires three-

dimensional models, cannot be addressed here. On the other hand, the solutions presented here

for axial longitudinal fracture growth will benefit from comparison to additional experimental

results.

References:

Abass, H. H., Soliman, M. Y., Tahini, A. M., Surjaatmadja, J., Meadows, D. L., Sierra, L. (2009) Oriented fracturing: A new technology to hydraulically fracture openhole horizontal well, SPE 124483, Proceedings of the 2009 SPE Annual Technical Conference and Exhibition, New Orleans, Louisiana, October 4-7 2009

Atkinson. C. and Thicerelin, M. (1993) The interaction between the wellbore and pressure-induced fractures, International Journal of Fracture 59, 457-467.

Aud, W.W., Wright, T.B., Cipolla, C.L., Harkrider, J.D., Hansen, J.T. (1994) The effect of viscosity on near-wellbore tortuosity and premature screenouts, SPE Annual Technical Conference and Exhibition, 25-28 September 1994, New Orleans, Louisiana

Brown, S. R. and C. H. Scholz (1986), Closure of rock joints, J. Geophys. Res. 91, 4939-4948. Bunger A.P. and E. Detournay. (2008), Experimental validation of the tip asymptotics for a fluid-driven crack, J.

Mech. Phys Solids 11, 3101-3115. Bunger A.P., A. Lakirouhani, and E. Detournay. (2010). Modelling the effect of injection system compressibility

and viscous fluid flow on hydraulic fracture breakdown pressure. Proceedings 5th International Symposium on In-situ Rock Stress. August 25-27, 2010, Beijing, China. To be submitted.

Chen, D. (2009) Work scope: modelling the effect of perforation-jetting on fracture initiation during stimulation, Geodynamics Ltd Internal Confidential Report.

Cherny, S., Chirkov, D., Lapin, V., Muranov, A., Bannkov, D., Miller, M., Willberg, D., Medvedev, O. and Alekseenko, O. (2009) Two-dimensional modelling of the near-wellbore fracture tortuosity effect, Int. J. Rock Mech. & Mining Sci. 46, 992-1000.

Chu, T. Y., Jacobson, R. D. and Warpinski, N. and Mohaupt, H. (1987) Geothermal well stimulated using high energy gas fracturing, Proceedings of 12 Workshop on Geothermal Reservoir Engineering, Stanford, January 20-22, 1987.

20

Crouch, S. L. and Starfield, A. M. (1990), Boundary Element Method in Solid Mechanics, Unwin Hyman, Boston Cuderman, J. F., Chu, T. Y., Jung, J. and Jacobson, R. D. (1986) High energy gas fracture experiments in fluid-filled

boreholes – potential geothermal application, Sandia Report, SAND85-2809. Daneshy, A. A. (1973) A study of inclined hydraulic fractures, SPE J 13: 61-68. Detournay, E. (2004), Propagation regime of fluid-driven fractures in impermeable rocks, Int. J. Geomechanics 4,

35-45. Detournay, E. and Jeffrey, R. (1986) Stress conditions for initiation of secondary fractures from a fractured

borehole, proceedings of the international symposium on rock stress and rock stress measurements, Stockholm, Sweden, 1-3, September, 1986.

Erdogan, F. and Sih, G. C. (1963), On the crack extension in plates under plate loading and transverse shear, J. Basic Engineering 85, 519-527.

Hallam, S. D. and Last, N. C. (1990) Geometry of hydraulic fractures from modestly deviated wellbores, JPT 43, 742-748.

Hamison, B. and Fairhurst C. (1970) In-situ stress determination at great depth by means of hydraulic fracturing, in Rock Mechanics—Theory and Practice, (ed.) W. H. Somerton, Am. Inst. Mining Engrg., 559–584, 1970

Jaeger, J.C., N. G. W. Cook, N.G.W. and R. W. Zimmerman, (2007) Fundamental of Rock Mechanics, 4th Ed., John Wiley & Sons.

Jeffrey, R. G. (1989) The combined effect of fluid lag and fracture toughness on hydraulic fracture propagation, SPE 18957, Proceedings of SPE Joint Rocky Mountain Regional/ Low Permeability Reservoirs Symposium and Exhibition, Denver, Colorado, March 6-8, 1989.

Jeffrey, R.G., Zhang, X., and Thiercelin, M. (2009) Hydraulic Fracture Offsetting in Naturally Fractured Reservoirs: Quantifying a Long-Recognized Process, Paper SPE 119351 presented at the SPE Hydraulic Fracturing Technology Conference, The Woodlands, Texas, 19-21 January.

Linkov, A. M. and Mogilevskaya, S. G. (1994) Complex hypersingular integrals and integral equations in plane elasticity, Acta Mech. 105, 189-205.

Manrique, J.F. and A. Venkitaraman, (2001) Oriented fracturing – A practical technique for production optimization, Proceedings of the 2001 SPE Annual Technical Conference and Exhibition held in New Orleans, Louisiana, 30 September–3 October 2001

Mogilevskaya, S. G., Rothenburg, L. and Dusseault, M. B. (2000a) Interaction between a circular opening and fractures originating from its boundary in a piecewise homogeneous plane, International Journal for Numerical and Analytical Methods in Geomechanics 24, 947-970.

Mogilevskaya, S. G., Rothenburg, L. and Dusseault, M. B. (2000b) Growth of pressure-induced fractures in the vicinity of a wellbore, International Journal of Fracture 104, L25-L30.

Newman, J. C. (1971) An improved method of collocation for the stress analysis of crack paths with various shaped boundaries. Report D-6376, NASA TN, 1-45.

Naredran, V. M. and Cleary, M. P. (1983) Analysis of growth and interaction of multiple hydraulic fractures, SPE 12272, in Proceedings of Reservoir Stimulation Symposium, San Francisco, CA, Nov 15-18, 1983.

Pollard, D. D., P. Segall, and P. T. Delaney. 1982. Formation and interpretation if dilatant echelon cracks, Geol. Soc. Am. Bull. 93, 1291-1303.

Settari, A. and H. S. Price. 1984. Simulation of hydraulic fracturing in low-permeability reservoirs, SPEJ April 1984, 141-152.

Soliman, M. Y., East, L. and Adams, D. (2008) Geomechanics aspects of multiple fracturing of horizontal and vertical wells, SPE 86889, SPE Drilling and Completion, Sept. 2008, 217-228.

Timoshenko, S. P. and Goodier, J. N. (1951) Theory of Elasticity, 3rd edition, Mc-Graw-Hill, New York. Walsh, J. B. and M. A. Grosenbaugh (1979), A new model for analysing the effect of fractures on compressibility, J.

Geophys. Res. 84, 3532-3536. Witherspoon, P. A., J. S. Y Wang,., K. Iwai and J. F. Gale (1980), Validity of cubic law for fluid flow in a

deformable rock fracture, Water Resour. Res. 16, 1016-1024. Weng, X. (1993) Fracture Initiation and Propagation from Deviated Wellbores, SPE 26597, presented at the 1993

ATCE, Houston, TX, 3-6 October, 1993. Yew, C. H., Mear, M. E., Chang, C. C. and Zhang, X. C. (1993) On perforating and fracturing of deviated wellbores,

SPE 26514, Proceedings of 68th Annual Tech. Conf., Houston, Oct, 1993. Zhang, X., Jeffrey, R. G. and Detournay, E. (2005), Propagation of a fluid-driven fracture parallel to the free surface

of an elastic half plane, Int. J. Numer. Anal. Meth. Geomech. 29, 1317-1340. Zhang, X. Jeffrey, R. G. and Thiercelin, M. (2007), Deflection of fluid-driven fractures at bedding interfaces: A

numerical investigation, J. Structural Geology 29, 396-410. Zhang, X. and Jeffrey, R. G. (2008), Re-initiation or termination of fluid-driven fractures at frictional bedding

interfaces, J. Geophys. Res. -Solid Earth 113, B08416. Zhang, X. Jeffrey, R. G. and Thiercelin, M. (2009), Mechanics of fluid-driven fracture growth in naturally fractured

reservoirs with simple network geometries, J. Geophysical Research -Solid Earth, 114, B12406.

21

List of Figures

Figure 1 Fractures around a wellbore, in-situ stress and coordinate system. The origin of the coordinate

system is at the wellbore centre. The size of flaws is exaggerated and should be much smaller than the well

radius in the computations.

h

H

Flaw

L

Wellbore

Pw x

y

22

0

0.2

0.4

0.6

0.8

1

1.2

0 0.5 1 1.5 2 2.5 3 3.5

(R+L)/R

No

rm S

IFs

SIF_internal p

SIF_internal p (Newman)

SIF_Farfield

SIF_Farfield (Newman)

p

R

L

L

R

Figure 2. Normalised SIFs (KI/ (p) ( )R L ) for a pair of cracks emanating from a wellbore subjected

to vertical far-field tension (left drawing) and subjected to uniform internal pressure p (right drawing).

For comparison, the results from the numerical work of Newman (1971) are shown.

23

x (m)

y(m

)

-0.2 -0.1 0 0.1 0.2 0.3 0.4 0.5

-0.1

0

0.1

0.2

0.3

0.4

0.5

0 MPa

0.8 MPa

Figure 3. Fracture trajectory for a crack emanating from the wellbore subjected to vertical loading with

0.25 and comparisons with the results from Mogilevskaya et al. (2000b). Fracture toughness of the rock

is 1 MPa m0.5. The triangle symbols have been digitised from a figure in Mogilevskaya et al. (2000b) for this

comparison.

24

x (m)

y(m

)

-0.2 -0.1 0 0.1 0.2 0.3 0.4

-0.2

-0.1

0

0.1

0.2

0.3

0.4 4.22 MPa

16.9 MPa

0.1 mm

(a)

25

x (m)

y(m

)

-0.2 -0.1 0 0.1 0.2 0.3 0.4

-0.2

-0.1

0

0.1

0.2

0.3

0.4 4.22 MPa

16.9 MPa

0.1 mm

(b)

Figure 4. Fracture trajectory and opening DD distributions for a hydraulic fracture emanating from the

wellbore subjected to higher deviatoric stress, resulting in a 4.0 value. Surface contact along part of the

fracture near the wellbore affects the results for these cases. (a) is for a small friction coefficient ( =0.01)

and (b) is for a large friction coefficient ( =0.8). The triangle symbols are the path from the paper by

Mogilevskaya et al. (2000b) and the purple lines are obtained from our model for cases that allow fracture

surface overlapping. Also, our initial crack is at an angle of 80 degree with respect to the x-axis, while it is 81

degree in Mogilevskaya et al. (2000b). The fictitious DD distributions around the borehole are not shown in

these figures.

26

x (m)

y(m

)

-0.2 -0.1 0 0.1 0.2 0.3 0.4-0.2

-0.1

0

0.1

0.2

0.3

0.4

0.1 mm

16.9 MPa

4.22MPa

Time=0.072 second

(a)

27

x (m)

y(m

)

-0.2 -0.1 0 0.1 0.2 0.3 0.4-0.2

-0.1

0

0.1

0.2

0.3

0.4

10 MPa

16.9 MPa

4.22MPa

Time=0.072 second

(b)

28

(Degree)

0 60 120 180 240 300 360

(MP

a)

-40

-30

-20

-10

0

10

20

30

40

t= 0.002 s t= 0.027 s t= 0.073 s

(c)

29

x (m)

y(m

)

-0.2 -0.1 0 0.1 0.2 0.3 0.4-0.2

-0.1

0

0.1

0.2

0.3

0.4

10 MPa

16.9 MPa

4.22MPa

Time=0.019 second

(d)

Figure 5. Fracture trajectory, opening DD (a) and internal fluid pressure (b) for the case of the borehole

radius =0.1 m and crack angle of 80 degree. The fluid viscosity is 0.01 Pa s and the injection rate is specified

as 0.0004 m2/s. The fictitious DD distributions around the borehole are not shown in Figure 5 (a). (c) The

hoop stress distributions around the borehole with a distance of 0.105m to the borehole centre at three

different times. In this figure, tensile stress is positive. The fracture is located at the angle of 80 degree with

respect to the positive x-direction. (d) internal pressure distribution along the fracture at the earlier time.

30

x (m)

y(m

)

-0.1 0 0.1 0.2 0.3 0.4

-0.1

0

0.1

0.2

0.3

16.9 MPa

4.22 MPa

Qin=0.0001 m2/s, =0.04 Pa.sQin=0.0004 m2/s, =0.01 Pa.s

Qin=0.0001 m2/s, =0.01 Pa.s

Qin=0.001 m2/s, =0.01 Pa.s

Variable parameters , but F=0.44

Figure 6 Fracture trajectories for different pairs of injection rate and fluid viscosity under the same

conditions as Fig. 5. The corresponding values of F decrease from 0.62, to 0.44 and 0.35 when the fracture

pathways shift from top to bottom.

31

x (m)

y(m

)

-0.4 -0.3 -0.2 -0.1 0 0.1 0.2 0.3 0.4-0.2

-0.1

0

0.1

0.2

0.3

0.4

0.5

0.6

100 MPa50 MPa

80 MPa

Time=0.165 second

(a)

32

x (m)

y(m

)

-0.4 -0.3 -0.2 -0.1 0 0.1 0.2 0.3 0.4-0.2

-0.1

0

0.1

0.2

0.3

0.4

0.5

0.6

100 MPa50 MPa

80 MPa

Time=0.165 second

(b)

33

Time (s)

0.00 0.05 0.10 0.15 0.20

Bor

ehol

e P

ress

ure

(MP

a)

40

50

60

70

80

90

100

30 degree60 degree0 degree

(c)

34

x (m)

y(m

)

-0.4 -0.2 0 0.2 0.4-0.2

-0.1

0

0.1

0.250 MPa

80 MPa

Qin=0.0004m2/s, =0.01 Pa.s

Qin=0.0001m2/s, =0.04 Pa.s

(d)

Figure 7 A pair of fractures growing from a wellbore: fracture trajectory and internal fluid pressure (a) for

low-angle fracture initiation; (b) for high-angle fracture initiation at the same instant and (c) the time-

dependent wellbore pressure responses for three fracture initiation angles. The fluid viscosity is 0.01 Pa s and

the injection rate is specified as 0.0004 m2/s. (d) comparison of fracture pathways for two different pairs of

fluid viscosity and injection rate for low-angle crack initiation. The red scatters have been given in (a) and the

blue scatters are based on the condition that the fluid viscosity is 0.04 Pa s and the injection rate is specified

as 0.0001 m2/s. The corresponding value of F is equal to 0.84 for both cases.

35

x (m)

y(m

)

-0.4 -0.3 -0.2 -0.1 0 0.1 0.2 0.3 0.4-0.2

-0.1

0

0.1

0.2

0.3

0.4

0.5

0.6

100 MPa40 MPa

60 MPa

Time=0.132 second

Figure 8. Fracture trajectory and internal fluid pressure in the presence of four fractures around the

wellbore. The fluid viscosity is 0.005 Pa s and the injection rate is specified as 0.0004 m2/s.

36

x (m)

y(m

)

-0.4 -0.3 -0.2 -0.1 0 0.1 0.2 0.3 0.4-0.2

-0.1

0

0.1

0.2

0.3

0.4

0.5

0.6

100 MPa50 MPa

80 MPa

Time=0.014 second

(a)

x (m)

y(m

)

-0.4 -0.3 -0.2 -0.1 0 0.1 0.2 0.3 0.4-0.2

-0.1

0

0.1

0.2

0.3

0.4

0.5

0.6

100 MPa50 MPa

80 MPa

Time=0.06 second

(b)

37

x (m)

y(m

)

-0.4 -0.3 -0.2 -0.1 0 0.1 0.2 0.3 0.4-0.2

-0.1

0

0.1

0.2

0.3

0.4hmin

Hmax

(c)

Figure 9 Fracture trajectory and internal fluid pressure in the presence of five starter fractures around the

wellbore, which are distributed evenly around the wellbore (a) at early time and (b) at large time in the case

where the fluid viscosity is 0.001 Pa s and the injection rate is specified as 0.0008 m2/s. The in-plane horizontal

and vertical far-field stresses are 50 and 80 MPa, respectively. The fracture alignment angle is 12°, 84°, 156°,

228° and 300°. (c) comparison of fracture pathways for different fluid viscosities with maxH =80 MPa and

minh =50 MPa at the same time t=0.06 s; the red line is for =0.001 Pa.s and the blue for =0.005 Pa.s.

38

x (m)

y(m

)

-0.4 -0.3 -0.2 -0.1 0 0.1 0.2 0.3 0.4-0.2

-0.1

0

0.1

0.2

0.3

0.4

100 MPa

50 MPa

80 MPa

Time=0.0846 second

Figure 10. Fracture trajectory and internal fluid pressure in the presence of six fractures around the wellbore

under the in-situ stresses H =80 MPa and h =50 MPa, where the fluid viscosity is 0.005 Pa s and the

injection rate is specified as 0.0012 m2/s. The fracture alignment angle is 30°, 120°, 150°, 210°, 240° and 330°.


Recommended