You are on page 1of 14

CSUG/SPE 149433

Importance of Shale Anisotropy in Estimating In-Situ Stresses and


Wellbore Stability Analysis in Horn River Basin
Safdar Khan, Sajjad Ansari, Hongxue Han, and Nader Khosravi, Schlumberger

Copyright 2011, Society of Petroleum Engineers

This paper was prepared for presentation at the Canadian Unconventional Resources Conference held in Calgary, Alberta, Canada, 15–17 November 2011.

This paper was selected for presentation by a CSUG/SPE program committee following review of information contained in an abstract submitted by the author(s). Contents of the paper have not
been reviewed by the Society of Petroleum Engineers and are subject to correction by the author(s). The material does not necessarily reflect any position of the Society of Petroleum Engineers,
its officers, or members. Electronic reproduction, distribution, or storage of any part of this paper without the written consent of the Society of Petroleum Engineers is prohibited. Permission to
reproduce in print is restricted to an abstract of not more than 300 words; illustrations may not be copied. The abstract must contain conspicuous acknowledgment of SPE copyright.

Abstract
Shale formations have laminated structures which result in significant differences in mechanical properties along the
orientations parallel to and perpendicular to laminations (bedding planes). These differences lead to anisotropic horizontal
stresses. Failure to consider the effect of anisotropic behavior of shale can have severe consequences for drilling. In rocks
with anisotropic mechanical properties and strength, there is a high risk of wellbore instability while building deviation angle
from vertical sections. Conventional wellbore stability analysis approaches do not consider material anisotropy and laminated
nature of shales, which can result in underestimated stresses leading to incorrect safe trajectory or mud-weights.

Shale formations in the Horn River Basin (HRB) are strongly anisotropic with anisotropic ratios varying from 1.2 to 3.5. In
this paper, the authors demonstrate the importance of considering anisotropy in estimation of in-situ stresses and wellbore
stability analysis. Two field case study examples are presented to underscore the consequences of neglecting anisotropy in
wellbore stability analysis.

INTRODUCTION
The Horn River Basin (HRB), the largest shale gas field in Canada, is located in northeastern British Columbia of the
Western Canadian Sedimentary Basin (WCSB). It is overlain by the Fort Simpson Formation and underlain by the carbonates
of the Lonely Bay Formation, Nahanni Formation or Pine Point Formation. Multi-laterals (horizontal drilling) with multi-
stage hydraulic fracturing are typically used in the HRB. Major drilling challenges in the area include borehole instability due
to sloughing, tight hole, stuck pipe, excessive reaming leading to large wellbore breakout/washout, and moderate to severe
lost circulation or combination of any of these problems. One of the most troublesome formations is Fort Simpson shale
which is about 800-1000m thick [Ross and Bustin, 2008] reactive high clay content shale which swells when contacted by
fresh water and usually consist of high pressure zones. To avoid these types of wellbore instabilities, an intermediate casing
string is typically set right after passing through the Fort Simpson shale [Stewart et al. 2000] or before entering Muskwa. In
this paper, the authors analyzed some of the wellbore instability problems encountered in the HRB and identify possible
causes of these problems.

SHALE ANISOTROPY
The term shale is normally used for the entire class of fine grained sedimentary rocks that contain substantial amount of clay
minerals [Lal, 1999]. Organic-lean Fort Simpson shales are clay rich, with almost equal concentrations of illite, kaolinite, and
chlorite [Ross and Bustin, 2008]. It is Devonian age, weak, brittle shale and calcareous silty laminated with high quartzitic
content [Stewart et al. 2000]. Shale is formed of platy type of grains when it is compacted, the grains are usually distorted
and elongated in one or more directions, and aligned in parallel to sub-parallel orientation, making it a layered material. Thus,
shale formations have laminated structures in the form of bedding, layering, stratification, fissuring which combined with in-
situ stresses result in directionally dependent rock properties. Because of this reason, shales are said to be inherently
anisotropic. When the properties of a material are the same in all directions, the material is said to be isotropic. On the
contrary, when the properties of a material vary with different orientations, the material is said to be anisotropic.

Anisotropy can also result from complex physical and chemical processes associated with transportation, deposition,
compaction, cementation, etc. Rocks which have undergone several depositional and formation processes may have more
2 CSUG/SPE 149433

than one direction of planar anisotropy which are not necessarily parallel to each other, such as foliation and bedding planes
in slates [Amadei, 1996]. Anisotropy in shales is observed at microscopic scale in the form of layering of fibers as seen in
scanning electron microscope (SEM), at core scale as well as at field scale as reservoir inter-beds.

Anisotropic rocks are often modeled as orthotropic or transversely isotropic (TI) media. Orthotropic material has orthorhombic
symmetry, i.e., three orthogonal planes of symmetry exist at each point in the rock and that these planes have the same orientation
throughout the rock. Transverse isotropic material has a rotational axis of symmetry at each point in the rock. Layered materials
such as shales exhibit similar properties within a bedding plane and different properties along different bedding planes. This
type of materials has a rotational axis of symmetry (Z-axis) as shown in Figure 1 which has isotropic properties in a plane
normal to that axis (X-Y plane). This plane is called the plane of transverse isotropy.

Figure 1. Rotational axis of symmetry in transversly isotropic material.

Elastic properties of a material are expressed using Hooke’s law which relates strain with stress. Isotropic materials have only
two independent variables (i.e. elastic constants, Young’s modulus and Poisson’s ratio) in their stiffness and compliance
matrices as opposed to 21 elastic constants in the general anisotropic case. Hooke’s law may be written in its generalized
form as

(1)

Where εij and σkl denote components of the second-order strain tensor ε and stress tensor σ and the Einstein summation
convention, in which a summation is made over repeated indices (i,j,k,l = 1,2,3). The Sijkl denote the components of the
fourth-order elastic compliance tensor, S [Nye, 1985]. Equation 1 can be inverted to express the components of the stress
tensor in terms of the components of the strain tensor as follows:

(2)

where the Cijkl denote the components of the fourth-order elastic stiffness tensor, C.

Although this notation, fourth-order tensor in Eq 1 and Eq 2 with 81 components is often used in theoretical studies, there is
no straightforward way to write out the 81 components in a matrix form [Jaeger et al., 2007]. There are only six independent
stress components and six independent strain components; making it 36 independent coefficients. Therefore, a more concise
approach is to use Voigt notation, in which the stress and strain components are expressed by 1x6 column vectors, and the elastic
compliances or stiffnesses are expressed by 36 coefficients that can be written as a 6x6 matrix:

(3)
CSUG/SPE 149433 3

Number of independent stiffness coefficients in Eq 3 is 21. However, if a material exhibits any physical symmetry, the
number of independent coefficients can be reduced further. For an orthotropic material, i.e., three orthogonal planes of
symmetry, number of independent stiffness coefficients reduces to nine [Amadei, 1996; Jaeger et al., 2007]:

0 0 0
0 0 0
0 0 0
(4)
0 0 0 0 0
0 0 0 0 0
0 0 0 0 0

For an isotropic material, there are only two independent stiffness coefficients; the elastic stiffness matrix takes the form:

0 0 0
0 0 0
0 0 0
and C11 - C12 - 2C44 = 0 (5)
0 0 0 0 0
0 0 0 0 0
0 0 0 0 0

For a transversely isotropic material (which has an axis of rotational symmetry), there are 5 independent stiffness
coefficients, and the elastic stiffness matrix takes the form:

0 0 0
0 0 0
0 0 0
and C11 - C12 - 2C66 = 0 (6)
0 0 0 0 0
0 0 0 0 0
0 0 0 0 0

For a transversely isotropic material, the elastic compliance matrix is

0 0 0
0 0 0
0 0 0
and S11 - S12 – 0.5S66 = 0 (7)
0 0 0 0 0
0 0 0 0 0
0 0 0 0 0

Where coefficients of S are given as follows [Jaeger et al., 2007]:

(8a)

(8b)

(8c)

The compliance coefficient S11 is the ratio of strain (εxx) to stress (σxx), and can be written as l/Ex, where Ex is the Young’s
modulus of the rock in the x direction; similarly for S22 and S33, and S44 is related to shear modulus, as 1/Gyz, where Gyz is the
shear modulus in the y-z plane. The off-diagonal components of S are related to “Poisson effect”. Thus, if Ei denotes the
Young’s modulus corresponding to axis xi , νij denotes the Poisson’s ratio that relates the expansion along axis xj to a
compression applied along axis xi, and Gij denotes the shear modulus in the plane xixj, the elastic compliance matrix for a
transversely isotropic material can be written as
4 CSUG/SPE 149433
ν ν
0 0 0
ν ν
0 0 0
ν ν
0 0 0
and S11 - S12 – 0.5S66 = 0 (9)
0 0 0 0 0
0 0 0 0 0
0 0 0 0 0

ν ν ν ν ν ν
Because of the symmetry, , , . Since S11 = S22, S13 = S23, S44 = S55, and S66 = 2(S11 – S12), Eq 9 can
be simplified follows:

E1 = E2 =E
E3 = E′
ν12 = ν21 = ν
ν31 = ν32 = ν′
G23 = G13 = G′
G12 = G =
ν

Where E and E′ are Young’s moduli in the plane of transverse isotropy and in direction normal to it, respectively, ν and ν′ are
Poisson’s ratios characterizing the lateral strain response in the plane of transverse isotropy to a stress acting parallel or
normal to it, respectively and G′ is the shear modulus in planes normal to the plane of transverse isotropy. Thus, elastic
compliance matrix for transverse isotropic material becomes:
ν ν′
0 0 0

ν ν′
0 0 0

ν′ ν′
0 0 0
′ ′
(10)
0 0 0 0 0

0 0 0 0 0

0 0 0 0 0

The two Young’s moduli E and E′, two Poisson’s ratio ν and ν′, and the shear modulus G′ are considered as the five
independent elastic stiffnesses. If bedding planes are considered to be horizontal, then E = Eh and E′= Ev; ν = νh and ν′=νv,
(h: horizontal direction and v: vertical direction) which can be determined from the acoustic stiffness coefficients (Cij):

ν ν ν ν
, (11a)

ν ν
, (11b)

1 ν 1 2 ν ν (11c)

For transversely isotropic material, the shear modulus G′ is often expressed in terms of E, E′, ν and ν′ through the following
empirical equation:
ν′
2 (12)
′ ′ ′

This relation was first introduced by Saint-Venant and later Worotnicki confirmed that most of the published experimental
data support the validity of the Saint-Venant approximation, with some exceptions [Amadei, 1996].
C
CSUG/SPE 1494
433 5

SHALE ANISOTROPY IN HORN RIVER


S R BASIN
a measured on core specimeens cut at diffeerent angles wiith respect to the
T static elasttic properties are
The t apparent diirections of
r
rock symmetry (bedding plan nes). Typically,, core specimenns are taken at least three in orientations,
o paarallel, perpenddicular and
a 45° with respect to bedding planes in a transverselyy isotropic maaterial as illusttrated in Figuure 2. Intact trransversely
at
isotropic rocks are typically stiffer in the plane
p of isotroppy (Eh) than inn the direction of the axis off symmetry (Ev). Amadei
[1996] analyzed 98 sets of daata on the elasttic moduli of anisotropic
a roccks collected byy various reseaarchers and fouund that in
m cases, Eh/Ev ranged betw
most ween 1 and 4.

Figure 2.
2 Core specim
men orientations in a TI materrial.

There is a stronng mechanical anisotropy preesent in the Hoorn River Basinn formations. Typically,
T T the ratio of Eh andd Ev ranges
f
from 0.9 to 2.55; however, th here are some areas where thist range can go up to 3.5.. An example of variation of o Young’s
m
modulus and Poisson’s
P ratio with differentt orientations is
i illustrated inn Figure 3. Thhis figure showws results fromm two core
s
specimens takeen from Musk kwa and Keg River formatiions. Zero deggree orientatioon represents verticalv core (normal
( to
b
bedding planes), and 90 degreees orientation represents horrizontal core (pparallel to beddding planes). As
A it can be seenn in Figure
3 Young’s moodulus measureed parallel to bedding
3, b plane is
i slightly highher than the onne measured noormal to beddinng plane in
b
both Muskwa and a Keg Riverr formations. ThisT was the caase in almost all a the core speecimens taken from HRB forrmations as
w as overlyinng Fort Simpsson formation. Normally, Poiisson’s ratio allso exhibits thee same trend (i.e.
well ( horizontal values are
h
higher b slightly highher than the hoorizontal one as shown in
than verrtical ones). In some exceptioonal cases, verrtical one can be
F
Figure 3 for Keeg River formaation at a particcular depth wheere the core speecimen was takken.

Figurre 3. Variation of Young’s modulus


m and Poisson’s ratio with
w orientationn.

Static Young’s modulus measured on core specimens takken from Fort Simpson,
S S Muskkwa, Evie and Keg River forrmations at
v
various depths are presented in Figure 4 too Figure 6. Datta from more than t 50 core specimens
s m three different areas of
from
H
HRB are presennted in these fiigures. In Areaa 1, Ev ranges from
f 15 GPa too 55 GPa, and Eh ranges from
m 30 GPa to 555 GPa with
6 CSUG/S
SPE 149433

ssome scattered data in higher range of 80 GPPa. Although most


m data in this area exhibitss an anisotropyy ratio, Eh/Ev of 1.0 to 1.8
a shown in thee histogram (Fiigure 4c), somee specimens inndicated a higheer range, up to 2.5.
as

Area 1 Arrea 1

(a) Hoorizontal vs. Veertical Young’ss modulus (b) Vaariation of Eh/E


Ev with depth

Area 1

m of Eh/Ev variaation
(c) Histogram

Figuree 4. Shale anisootropy in Hornn River – Area 1.

There seem to be two differeent ranges in annisotropy withhin the same foormation as maarked in dottedd line shapes inn Figure 4.
T
T could be due
This d to differencces in mineral composition
c orr presence of micro
m fractures created duringg core recoveryy process.

In case of Areaa 2, which is seeveral hundredd meters away from f Area 1, thhe range of vaalues of Ev and Eh is higher by 25%. Ev
r
ranges from 155 GPa to 100 GPa,G and Eh rannges from 30 GPaG to 100 GPaa (one point shhowed a value of 160 GPa, but it seems
too be an outlierr). Area 2 exhib bits a higher annisotropy ratio, Eh/Ev than Arrea 1; it rangess from 0.9 to 2.5 as shown in Figure 5c.
T
There were onlly few sampless available in Area A 3, althouugh the range of
o anisotropy ratio,
r Eh/Ev seeems to be conssistent with
o
other two areass, quite a bit off variation is repported in sampples taken from
m Muskwa form mation.
C
CSUG/SPE 1494
433 7

Area 2 Areaa 2

(a) Hoorizontal Vs. Vertical


V Young’’s modulus (b) Variatioon of Eh/Ev witth depth

Area 2

m of Eh/Ev variaation
(c) Histogram

Figuree 5. Shale anisootropy in Hornn River – Area 2.


2

Area 3 Arrea 3

(a) Hoorizontal Vs. Vertical


V Young’’s modulus (b) Variatioon of Eh/Ev witth depth

Figuree 6. Shale anisootropy in Hornn River – Area 3.

Figure 4 to Figuure 6 illustrate the mechanicaal anisotropy present


F p in shalees due to its fabbric, layering, etc.
e There are many
m other
n
non-structural s
sources that caan contribute too mechanical annisotropy, e.g. presence of frractures, tectonnic activity, streess history,
e Changes inn pore pressuree and stress cooncentration caan also influence mechanicall anisotropy. Inntact rock anisotropy can
etc.
d
decrease with an
a increase in confinement. Micro M cracks in
i rock would close under hiigher confiningg pressures, poore volume
r
reduction/enhan ncement can allso occur with change in poree pressure or coonfining pressuure; this will chhange overall stiffness
s of
thhe rock in all directions
d thus changing the anisotropy.
a Lerau et al. [19811] conducted a series of laborratory tests under triaxial
c
compression att confining preessures rangingg between 0 anda 80 kPa. They reported a linear increaase in values of o Young’s
m
modulus and Poisson’s
P ratio
o in vertical direction,
d and decrease in the t anisotropy ratio from 2.14 to 1.69 with w higher
8 CSUG/SPE 149433

confinement [Amadei, 1996]. Therefore, in order to fully characterize the mechanical anisotropy, not only the effect of
bedding planes but other aspects discussed above should also be considered.

EFFECT OF ANISOTROPY IN ESTIMATION OF IN-SITU STRESSES


Estimation of in-situ stresses is the second fundamental block in geomechanical calculations after rock mechanical properties.
It is often possible to estimate the magnitude and directions of in-situ stresses from sonic logs, but they need to be calibrated
using field measured data. In-situ stresses are vertical and two horizontal stresses. Near ground surface, the principal stress
directions are vertical and horizontal. The vertical normal stress is assumed to be equal to the weight of the overlying rock
and can be computed by integrating the bulk density log data:

z
σ v = g ∫ ρ b ( z )dz (13)
0

where σV is the vertical stress, ρb is the bulk density, g is the acceleration due to gravity, and z is the vertical depth. In the
absence of density log, as a rule of thumb, vertical stress gradient can be assumed as 0.023-0.026 MPa/m or 1.0 to 1.2 psi/ft
on an average.

Although, there are several methods used to estimate minimum horizontal stress, σh, the following poroelastic model is the
most commonly used one:

σ h = K 0 (σ v − αPp ) + αPp + Tectonic (14)

ν
In this equation, α = Biot’s poro-elastic coefficient, and K 0 = , where ν is Poisson’s ratio. The coefficient, K0 is often
1 −ν
called in the geotechnical literature the coefficient of earth pressure at rest. This expression was derived based on the
following assumptions: (1) the rock is a continuous, homogeneous, isotropic, linearly, elastic (CHILE); (2) the rock is under
gravity alone with vanishing horizontal displacements; and (3) the loading history has no influence on how in-situ stresses
build-up. It also implies that horizontal and vertical stresses vanish at the Earth’s surface. This classical approach for
estimating in- situ stresses has several limitations in predicting the magnitude of horizontal in situ stresses. Quite often, the
difference between predicted and measured stresses is attributed to the effects of tectonic, residual and thermal stresses,
anisotropy, etc.

The effect of anisotropy on gin-situ stresses in homogeneous rock masses has been addressed by Amadei et al. [1987] and
Amadei and Pan [1992] who proposed expressions for the coefficient, K0 in transversely isotropic, orthotropic and generally
anisotropic rock masses. For transversely isotropic material with horizontal plane as the plane of transverse isotropy,

Eh ν v
K 0 _ aniso = (15)
E v 1 −ν h

Eq. 15 indicates that the horizontal stresses increase as the anisotropy ratio, Eh/Ev increases. Since vertical stress depends on
gravitational loading, it is not affected with anisotropy. The new expression for the minimum horizontal stress with
anisotropy considerations becomes:

σ h _ aniso = K 0 _ aniso (σ v − αPp ) + αPp + Tectonic


Eh ν v
σ h _ aniso = (σ v − αPp ) + αPp + Tectonic (16)
E v 1 −ν h

For an isotropic case, Eh/Ev =1, substituting this value in the above equation will result the standard poro-elastic stress
equation (Eq. 14). Ignoring the effect of anisotropy will underestimate the horizontal stress. As it was discussed above, in
HRB, the anisotropy factor varies from 1.0 to 2.5, if this aspect is not considered, horizontal stresses will be underestimated
by a factor of 2.5. Figure 7 compares minimum horizontal stress estimated from isotropic and anisotropic mechanical
properties in a well located the Horn River Basin. The anisotropic ratio, Eh/Ev in this well ranges from 1.0 to 1.5. The green
profile is the isotropic stress profile computed using Eq. 14; it looks quite flat, nearly constant throughout the formation, the
stress was underestimated by 33%.
CSUG/SPE 149433 9

Isotropic σh

Anisotropic σh

Figure 7. Comparison of minimum horizontal stress profile estimated from isotropic and anisotropic mechanical properties.

Figure 8 shows srip logs of four main formations taken from a mechanical earth model constructed for another well in HRB.
In this figure, track 9 shows variation of minimum horizontal stress gradient. Solid red profile represents the isotropic stress
gradient and dotted red profile represents the anisotropic stress gradient. Considerable amount of difference (4 KPa/m) in
anisotropic and isotropic stress gradients is observed in this well within the Fort Simpson formation as highlighted in the
figure. For other formations, this difference reduces to 1.2 to 1.8 KPa/m. This indicates that Fort Simpson formation exhibits
stronger anisotropy than other formations.

EFFECT OF ANISOTROPY IN WELLBORE STABILITY


When a wellbore is drilled, the rock stresses in the vicinity of the wellbore are redistributed as the support originally offered
by the drilled out rock is partially replaced by the hydraulic pressure of the drilling mud. The redistributed stresses (i.e.
wellbore stresses) are normally referred to as the hoop stress, σθ, which acts circumferentially around the wellbore wall, the
radial stress, σr, and the axial stress, σa, which acts parallel to the wellbore axis. Since drilling mud pressure is uniform in all
directions, it cannot balance the earth stresses which may not be equal in all directions. Consequently, rock surrounding the
wellbore is strained and may fail if the redistributed stresses exceed the rock strength, either in tension or compression,
resulting in instability. If hoop stress at the wellbore wall reduces to zero or to a negative value, tensile fractures (drilling
induced fractures) are created along a plane perpendicular to the the minimum stress, which can cause lost circulation.
Alternatively, the formation can fail in compression or shear when shear stress exceeds the formation strength. In this case,
the formation caves in or spalls off, creating breakouts or sometimes, hole pack off. Therefore, wellbore stresses and rock
strength are key parameters in wellbore stability analysis.

Wellbore stresses are normally computed using classical Kirsch equation. For an isotropic material, in a vertical borehole,
hoop stress is expressed as:

σ 'θ = (σ 'H +σ 'h ) − 2(σ 'H −σ 'h )cos(2θ ) − (Pw − Pp ) (17)


10 CSUG/SPE 149433

Isotropic σh

Anisotropic σh

Figure 8. Stress variation in different formations.


C
CSUG/SPE 1494
433 11

Where σ′θ = effective hoop stress, σ′H andd σ′h = effectivve maximum annd minimum horizontals
W h streesses, Pw = mudd pressure,
Pp = pore presssure. Eq. 17 do oes not containn any materiall property, is suitable
s for isootropic and hom mogeneous maaterial. For
a
anisotropic matterial, Jaeger an
nd Cook [20033] provided thee following equuation to eastimmate hoop stresss:

σ 'θ = aσ 'θ ,a +bσ 'θ ,b +cσ 'θ , pw (


(18)

Where variablees, a, b, and c are related too material anissotropy, refer to


W t Jaeger and Cook [2003] and Suárez-Riivera et al.
[2006] for morre details on th hese parameters. Using this equation, Deeenadayalu and Suárez-Riveraa [2010] compputed hoop
s
stress gradient for various aniisotropic factorrs as shown in Figure 9.

Case of Eh/E
Ev =2.5

Figure 9. Elastic hoop stresses around a hoorizontal wellboore function off the degree of elastic anisotroopy [after, Deeenadayalu
and Suárrez-Rivera, 2010].

As can be seen from Figure 9,


A 9 the top and bottom
b orientatiions of the horrizontal wellbore are the mostt prone to fractturing. The
isotropic case (dark
( blue currve) exhibits thhe highest commpression and this decreases with increasinng anisotropy. The more
a
anisotropic rocck exhibits thee highest tensioon (red curve)), thus, even under
u balancedd wellbore preessure conditioons, tensile
f
fractures can bee created. One of the key parrameters in sheear failure criterion is unconfiined compressiive strength (UUCS) of the
r
rock. The UCS values for anisotropic rock considerably
c vaary with orienttation.

In the Horn Rivver Basin, anissotropic factor, Eh/Ev ranges between


b 1.0 annd 2.5. From Figure 9, the hooop stress gradiient for the
c
case of Eh/Ev = 2.5 is approxiimately, 0.2 psii/ft, whereas foor an isotropic case (Eh/Ev=1)), hoop stress gradient
g is 0.5 psi/ft.
p This
m
means, isotropiic case will ressult in a relativvely wider saffe mud weight window than anisotropic case, which can mislead to
b
believe there exist
e a higher mud weight limit l for tensille fractures. Inn such cases, it is relativelyy easier to creeate tensile
f
fractures or reoopen existing frractures, leadinng to lost circullation, in somee cases, total loosses.

Wellbore stabillity (WBS) anaalyses were coonducted for tw


W wo different wells
w from HRB B using isotroppic as well as anisotropic
a
r
rock propertiess and horizontaal stresses. Figgure 10 compaares WBS resuults of isotropicc with anisotroopic horizontal stress for
W A. In thiss example, thee mud weight window
Well w becam
me narrower when
w anisotropic horizontal stresses
s were considered.
c
M weight wiindow became further narrow
Mud wer for second well. Both WB BS analyses ressults were founnd to be consisstent and in
g
good agreemennt with breakouut and induced fractures seen in image logs as shown in Fiigure 11.
12 CSUG/SPE 149433

Isotropic horizontal stress:


Wide mud weight window

Anisotropic horizontal stress :


Narrow mud weight window

Figure 10. WBS analyses results with isotropic (top) and anisotropic (bottom) horizontal stress inputs.
CSUG/SPE 149433 13

Drilling induced
fractures

Figure 11. Comparison of drilling induced fractures predicted vs. image logs.

These examples clearly demonstrate that ignoring anisotropy can lead to underestimation of stresses and overestimation of
breakdown pressure limits, leading to incorrect safe mud weight windows. Overestimation of breakdown pressures can result
in lost circulation or serious losses, especially when natural fractures are present.

Numerous wellbore instability problems have been reported in the Horn River Basin (HRB); most of these problems occurred
while building angle within Fort Simpson formation. In many of these cases, operators had to halt or alter their planned
trajectories or sidetrack to avoid problematic zones. Although various drilling challenges such as sloughing, tight hole, stuck
pipe, excessive reaming leading to large wellbore breakout/washout were reported, cases of lost circulation and hole pack off
were predominant among all others. This could have caused primarily due to neglecting anisotropy aspects of shales which
resulted in underestimation of stresses and overestimation of breakdown pressures.

CONCLUSIONS
Shale formations are strongly anisotropic due to their laminated structures which result in significant differences in
mechanical properties along the orientations parallel to and perpendicular to bedding planes. These differences lead to
anisotropic horizontal stresses. Mechanical anisotropy ratio in HRB ranges from 1.2 to 2.5, and in some cases it goes up to
3.5. Lower part of the Fort Simpson formation seems to have higher anisotropy ratios. Failure to consider the effect of
anisotropic behavior of shale can result in underestimation of stresses and overestimation of breakdown pressures. Sever mud
losses and lost circulation reported in Fort Simpson and upper Muskwa formations clearly indicate that breakdown pressures
were overestimated. Therefore, to avoid such problems, characterizing shale anisotropy should be an essential part of well
design program and all geomechanical analyses must account for mechanical anisotropy.

ACKNOWLEDGEMENTS
The authors thank Schlumberger for its support and the clients for releasing the data for this publication.

REFERENCES
Amadei, B. [1996]. Importance of Anisotropy When Estimating and Measuring In Situ Stresses in Rock. Int. J. Rock Mech.
Min. Sci. & Geomech. Abstr. Vol. 33, No. 3, pp. 293-325.
Amadei, B. and Pan E. [1992]. Gravitational stresses in anisotropic rock masses with inclined strata. Int. J. Rock Mech. Min.
Sci. & Geomech. Abstr. 29, 225-236.
14 CSUG/SPE 149433

Amadei, B., Savage W. Z. and Swolfs H. S. [1987]. Gravitational stresses in anisotropic rock masses. Int. J. Rock Mech.
Min. Sci. & Geomech. Abstr. 24, 5-14.
Deenadayalu, C. and Suárez-Rivera R. [2010]. The Effect of Horizontal Anisotropic Gas Shales. 44th US Rock Mechanics
Symposium and 5th U.S.-Canada Rock Mechanics Symposium, Salt Lake City, UT June 27–30, 2010.
Jaeger, J. C. and Cook, N. G. W. [2003]. Fundamentals of Rock Mechanics. Third edition, Chapman and Hall, p297.
Jaeger, J. C., N. G. W. Cook, and R. W. Zimmerman [2007]. Fundamentals of Rock Mechanics (Fourth Edition), Blackwell
Publishing.
Lal, Manohar [1999]. Shale Stability: Drilling Fluid Interaction and Shale Strength. SPE Latin American and Caribbean
Petroleum Engineering Conference held in Caracas, Venezuela, 21–23, April 1999. (SPE 54356)
Lerau J., St. Leu C. and Sirieys P. [1981]. Anisotropie de la dilatance des roches schisteuses. Rock Mech. 13, 185-196
(1981). (In French).
Nye, J.F. [1985]. Physical Properties of Crystals: Their Representation by Tensors and Matrices. Oxford University Press.
Ross, D. J. K. and Bustin, R. M. [2008]. Characterizing the shale gas resource potential of Devonian–Mississippian strata in
the Western Canada sedimentary basin: Application of an integrated formation evaluation. AAPG Bulletin, v. 92, no. 1, pp.
87–125.
Stewart, M. R, Kosich, B., Warren, B. K., Urquhart, J. C., and McDonald, M.J. [2000]. Use of Silicate Mud to Control
Borehole Stability and Overpressured Gas Formations in Northeastern British Columbia. SPE/CERI Gas Technology
Symposium, Calgary, Alberta Canada, 3–5 April 2000. (SPE 59751)
Suárez-Rivera R.,Green, S. J., McLennan, J. and Bai, M. [2006]. Effect of Layered Heterogeneity on Fracture Initiation in
Tight Gas Shale. SPE Annual Technical Conference and Exhibition, San Antonio, TX, USA, 24-27 Sep. 2006 (SPE 103327).

You might also like