You are on page 1of 13

Design of Caprock Integrity in Thermal

Stimulation of Shallow Oil-Sands Reservoirs


Yanguang Yuan, SPE, and Bin Xu, SPE, BitCan G&E Incorporated; and
Claes Palmgren, SPE, Value Creation Incorporated

Summary and failure (e.g., reactivation of pre-existing weak planes) can


Stakeholders in in-situ oil-sands development take caprock-integ- damage the well casing, breaking its hydraulic-sealing capacity.
rity issues seriously. The industry is faced with the challenge of Whether hydraulic or mechanical, caprock integrity becomes a
determining an optimal operating pressure in the reservoir where, geomechanical issue related to caprock deformation and potential
in general, the pressure should stay significantly low to ensure the failure. Caprock-integrity analysis compares the prevailing stress
caprock integrity while being significantly high for enhanced oil conditions against the material strength. The ongoing stress condi-
production and economics. This paper presents a comprehensive tion is the induced stresses superimposed on the virgin in-situ
work program on the subject for a shallow oil-sands play. stresses. Therefore, there are three major components to any geo-
Caprock integrity considers the induced stress and deformation mechanical work program: determination of the original in-situ
in a caprock during the thermal stimulation of an oil-sands reser- stresses, evaluation of the induced stresses, and measurement of
voir. A minifrac-test program is undertaken to define the original the mechanical properties. Minifrac tests are the most reliable
in-situ stress state. Laboratory tests are carried out to measure the method to measure the in-situ minimum stress; the induced
deformation and strength properties. Simulations are run to calcu- stresses are normally inferred from geomechanical simulations;
late the induced stresses and evaluate them against the mechanical and the mechanical properties should be measured ideally from
strength. This paper describes some important quality-control cores obtained from the project sites. This general geomechanical
issues for these activities. For the minifrac tests, multiple cycles work program is followed in the oil-sands industry now. How-
and use of flowback are promoted for enhanced efficiency and ac- ever, special considerations are warranted for oil-sands develop-
curacy. Laboratory tests are recommended on whole cores in a ment where the formations are relatively shallow. This paper will
drained condition at a slow strain rate. Numerical simulations present some relevant examples of how to ensure quality control
should use site-specific and laboratory-measured material proper- in these special cases.
ties. On the basis of the limited sensitivity analyses, the thermal- Both the industry and government regulatory agencies take the
expansion coefficient of the reservoir and Young’s modulus of the caprock-integrity issue seriously. For the in-situ development of
caprock are found to significantly affect the caprock deformation the Alberta oil-sands reservoirs, systematic efforts in this regard
and/or induced stresses. were first initiated for cyclic-steam-stimulation (CSS) operations
(Smith et al. 2004). More recently, steam-assisted gravity drain-
age (SAGD) has become another mainstream commercial in-situ
Introduction oil-sands recovery process. SAGD is normally thought to be gen-
Collectively, caprock refers to a certain interval in the overburden tler on the caprock because it operates at a lower pressure than
rock formations above a petroleum reservoir containing the reser- CSS. However, as the operational experience with SAGD grows,
voir fluids within the reservoir. It is normally shaly, with high evidence indicates that proactive precautions are still necessary to
clay content and low permeability. Sometimes, it immediately safeguard the caprock integrity. Yuan (2008) presented theoretical
overlies the pay zone. In other cases, there is a buffer zone be- arguments about why attention also should be paid to integrity
tween the caprock and pay zone. During petroleum exploitation, issues in SAGD. The current paper will apply these theoretical
the caprock plays an important role in safeguarding against the principles in the context of shallow oil-sands reservoirs.
hydrocarbon fluid, stimulating materials, and/or their mixture In the following sections, case histories are given to illustrate
invading zones above the caprock. Often, these zones contain important quality-control issues in various caprock-integrity stud-
groundwater aquifers. ies. First, complexities encountered in the oil-sands and the corre-
Ultimately, caprock integrity considers hydraulic integrity— sponding best practices during minifrac tests are described. Next,
no reservoir fluids should escape through the caprock into the geomechanical laboratory tests are explored. The subsequent sec-
groundwater aquifers or to the surface. In general, the hydraulic tion is dedicated to a study of the geomechanical simulations
integrity is already maintained naturally, as in the geological his- which combine both the field-obtained data and laboratory-meas-
tory of the caprock preventing further upward hydrocarbon migra- ured material properties to derive the safe SAGD operating pres-
tion. It is the process-induced mechanical deformation and sure. The effect of reservoir depth on caprock integrity is also
potential failure of the caprock during thermal operations that specifically discussed.
may introduce new hydraulic conduits and thus compromise the The target reservoir discussed in this paper belongs to the
hydraulic integrity. Therefore, hydraulic integrity becomes a me- McMurray formation sandstones of the Lower Mannville in the
chanical integrity issue. Fort McMurray area. Fig. 1 offers a schematic about the general
Caprock integrity also considers caprock mechanical integrity stratigraphic column. The reservoir is generally less than 100 m
(i.e., deformation and failure in the caprock strata). For example, deep, but its thickness is approximately 40 m, making the asset
surface heave, which is rock deformation reflected on the ground, attractive. The Clearwater formation constitutes the regionally
can alter the environment by changing the landscape or the surface continuous caprock, which is mostly shale with interbedded silty-
or shallow subsurface hydrogeological conditions. Such surface to-sandy mudstone. The caprock thickness ranges from 46 to 61
heave could damage surface installations and infrastructures, and m. A Wabiskaw member is sandwiched between the Clearwater
have other unintended impacts. Furthermore, rock deformation caprock and McMurray pay zone. It is a marine shore face system
conformably overlying the McMurray formation. It may also con-
tain sand facies that can be bitumen-saturated. Therefore, the
Copyright V
C 2013 Society of Petroleum Engineers
Wabiskaw member constitutes a transitional buffer zone from the
This paper (SPE 149371) was accepted for presentation at the Canadian Unconventional reservoir pay zone upward to the caprock. A solvent-assisted low-
Resources Conference, Calgary, 15–17 November 2011, and revised for publication.
Original manuscript received for review 22 November 2011. Revised manuscript received for
pressure SAGD is proposed to develop the reservoirs (Palmgren
review 18 December 2012. Paper peer approved 17 May 2013. et al. 2011). Significant efforts have been commissioned by the

266 July 2013 Journal of Canadian Petroleum Technology


mum stress, it should be detected by the minifrac tests and the frac-
Quaternary Till
ture being measured is horizontal. If the minifrac tests detect an
Smin smaller than the density-derived Sv, it means the fracture
being measured is vertical. The measured stress represents SHmin.
Thus, the fracture-closure pressure measured in a minifrac test is,
at most, equal to Sv and in any case should never be larger than Sv.
Upper Clearwater Shales Three test programs were conducted for the target shallow res-
ervoir in the particular study. They were performed by three dif-
ferent service providers on four different wells. The first test
program was completed in 2009 by a wireline unit. It measured
fracture-closure-pressure gradients at 24 to 36 kPa/m [Appendix 2
Lower Clearwater Shales and Appendix 3 of Clearwater Pilot Application submitted to
Energy and Resources Conservation Board (ERCB) of Alberta,
Wabiskaw Shales
January 2010]. Doubts should be raised about these values
because they are larger than the density-derived Sv at approxi-
mately 21 kPa/m. Indeed, subsequent analysis led to a conclusion
that most measurements provided data of low confidence and
should be deemed inconclusive, although one successful measure-
McMurray Oil Sands ment did provide a closure-pressure gradient of 20.57 kPa/m.
A borehole is a stress concentrator. Stresses around a borehole
are higher than the far-field in-situ stress condition. Therefore, a
proper minifrac test should inject a sufficient volume of liquid for
the fractures to propagate outside of the stress-concentration area
and for the majority of test length analyses to focus on far-field
Fig. 1—A schematic about the general well stratigraphic col- stresses. In the previously described tests completed in 2009, the
umns encountered in the shallow reservoir. total injected volume was only up to 10 L per test. This small injec-
tion volume may be one of the reasons for the unusually high frac-
operator, some of which are still ongoing, to design an operating ture-closure pressure as described previously. The fracture may
pressure that maintains the caprock integrity. have stayed within the influence domain of the borehole. There-
fore, the interpreted fracture-closure pressure reflects the stress
Minifrac Tests concentration near the borehole, not the far-field stress condition.
Minihydraulic-fracturing stress tests, commonly called minifrac Multiple Test Cycles for Consistency Check. The in-situ
tests, have been routinely used by the industry to measure the in- stress magnitudes are small at shallow depths. Thus, accurate
situ stresses. They are regarded widely as the most reliable interpretations of the fracture closure become an important issue.
method to define the in-situ minimum stress, Smin. Using con- For example, a 50-kPa inaccuracy in the interpreted closure pres-
trolled high-pressure water injection, the minifrac tests create a sure, Pc, is equivalent to 0.5 kPa/m at a 100-m depth, but only
fracture and propagate it for a sufficient distance from the injec- 0.05 kPa/m if the test depth is 1000 m. Two measures can help to
tion well and into the formation. This ensures the fracturing enhance the interpretation’s certainty. One is to use multiple
behaviour being dominated by the far-field stress condition. The injection and shut-in cycles, and the other is to use flowback, as
pressure data are analyzed to estimate the fracture-closure pres- described later. When using multiple injection and shut-in cycles,
sure Pc. The latter can then be equated to the in-situ minimum the interpreted closure pressure from each cycle should remain
stress acting perpendicular to the fracture. For more background similar as compared with their average from the multiple cycles,
information on minifrac tests in the industry, one may refer to the thus better representing the in-situ minimum stress. Fig. 2a shows
review paper by Hudson et al. (1993). Bell et al. (1994) summar- one example minifrac test performed in the McMurray sands with
ized the in-situ stress measurements made mostly at deep depths nine cycles. The pressure declined quickly during the shut-in,
in the western Canadian sedimentary basin. Some tests were per- making it difficult to derive a reliable Pc. As a result of on-site
formed in the context of oil sands [e.g., those reported in Settari real-time analysis, the difficulty was spotted immediately and the
and Raisbeck (1979), Chhina and Agar (1985), Chhina et al. corresponding corrective measures were taken. More cycles were
(1987), Kry (1989), Proskin et al. (1990), Kry et al. (1992)]. Our used, which yielded consistent closure behaviour eventually in the
test procedures and analysis methods incorporate the learnings last four cycles (Fig. 2b). Therefore, real-time analysis and multi-
from these earlier works and also implement new developments ple shorter cycles certainly assisted in enhancing the data quality
and/or adjustments dedicated to the shallow depths and the uncon- in this difficult situation.
solidated nature of the oil sands and overlying caprock intervals. High formation-breakdown pressures and fracture-propagation
pressures were observed during the test previously described (Fig.
Sv as a Quality-Check Index. At the shallow depths associ- 2a). This was caused by the significant near-wellbore friction dur-
ated with the target oil-sands reservoir, the original in-situ stress ing the injection. It can be observed during the step-rate test in
condition can be defined by three principal normal stress compo- Cycle 6. Every time the injection rate increased or decreased
nents: the vertical stress Sv and two horizontal stresses, which are abruptly, the injection pressure rose or dropped correspondingly
commonly denoted as the maximum and minimum horizontal by a relatively large amount (Fig. 3), which is evidence of flow-
stresses, SHmax and SHmin, respectively. Sv can be estimated reli- induced friction in the injection system. Near-wellbore complex-
ably from the overburden lithological column by integrating the ities are responsible for the high friction. One example is perfora-
density log, which typically ranges from 20 to 22 kPa/m. As a tion damage, where the explosive action of the perforating
rule of thumb, Sv¼21 kPa/m. operation and the associated high temperatures likely cause signif-
In theory, a properly executed and interpreted minifrac test icant compaction to the rock.
should never measure an Smin larger than the density-derived Sv. Flowback. A further enhancement to the minifrac test is a
Hydraulically driven fracture propagation is always perpendicular flowback procedure. For the flowback, a certain volume of water
to the direction of Smin. This direction represents the least resist- is withdrawn manually from the injection system (wellbore plus
ance (i.e., requiring the smallest pressure to extend the fracture). the fracture) during the shut-in period. Instead of waiting for the
Away from the mechanical influence domain of the borehole, the fracturing pressure to decline in response to the natural leakoff
fracture should be perpendicular to the least of the three stress from the fracture, the fracture now closes because of the manually
components: Sv, SHmax, and SHmin. Therefore, if Sv is the mini- induced pressure drop. As a result, the fracture closure is

July 2013 Journal of Canadian Petroleum Technology 267


(a)
BHP@Wellhead Sv = 2.23 140
Hydrostatic, 1.03 Smin = 2.23
Rate 6 9 120
10
Cycle 3 8
2 4 5 100
8

Injection rate, L/min


80

BHP, MPa 6 60

40
4
20
2
0

0 –20
2:09 PM 2:24 PM 2:38 PM 2:52 PM 3:07 PM 3:21 PM 3:36 PM

(b) Summary
8

5
Pressure (MPa)

2 Preopen
ISIP
1 Fracture closure
Pc,sqrt
Pc, compliance pressure = 2.23 MPa

0
0 2 4 6 8 10
Cycle

Fig. 2—(a) Recorded pressure and injection/flowback rate history for a minifrac test in the McMurray reservoir at 105-m true vertical
depth (TVD). The bottomhole pressures (BHP) were calculated from a surface pressure sensor at the pump plus the hydraulic
head (“Hydrostatic”) from the water column. The overburden weight (Sv) was calculated from the density log. Smin was the in-situ
minimum stress or fracture-closure pressure interpreted from the pressure data. The negative rate means flowback. Similar con-
ventions are used in this paper unless otherwise specified. (b) Various characteristic pressures interpreted from the minifrac test.
Preopen denotes the fracture reopening pressure where the fracture starts to reopen during the subsequent injection. In the first
cycle, Preopen corresponds to the breakdown pressure when the fracture is formed. ISIP is the instantaneous shut-in pressure;
Pc, sqrt refers to the fracture-closure pressure extracted by the sqrt(dt) plot, and Pc, compliance is the fracture-closure pressure
extracted by the compliance plot from the flowback tests.

relatively quick. Moreover, the difference in the pressure behav- almost a necessary component. Without flowback, a test may take
iour before and after the fracture closure is magnified by the flow- days to complete properly when multiple cycles are used. With
back, which makes the closure defined more clearly. Fig. 4 shows flowback, a test with multiple cycles can be completed in 2 to 3
one example plot of pressure vs. square root of the shut-in time. hours. Fig. 5a shows one example test performed in the Clear-
The inflecting pattern is clear on the plot, and the intersection of water capping shale. It took approximately 3 hours to finish nine
two different slopes points to the fracture closure. The detailed cycles. A consistent, and thus reliable, fracture-closure pressure
description of fracture flowback tests is outside the scope of this was obtained as shown in Fig. 5b.
paper but can be found in some published papers (Raaen and It is our observation that although minifrac tests are used rou-
Brudy 2001; Raaen et al. 2001). tinely in the industry, significant attention should be paid to the
Natural leakoff from the fracture can be slow if the water mo- unique conditions in the oil-sands development and the corre-
bility is small, which is common when testing caprock shales. It is sponding demands for high-quality data and interpretation accu-
also likely to be small in the oil-sands pay zone if it has only mini- racy. Some measures that can enhance the quality include
mum initial water mobility. In these cases, the flowback becomes multiple injection/shut-in cycles to check for consistency in the

268 July 2013 Journal of Canadian Petroleum Technology


12 BHP@Wellhead 140
Hydrostatic, 1.03
Smin = 2.23 120
10 Rate
100

Injection rate, L/min


8
80

BHP, MPa
6 60

40
4
20
2
0

0 –20
14:55:41 14:57:07 14:58:34 15:00:00 15:01:26 15:02:53 15:04:19

Fig. 3—A history plot of pressure and injection rate in the step-rate test Cycle 6 for a minifrac test in the McMurray reservoir at 105-
m TVD. Other details are the same as in Fig. 2a.

data and on-site preliminary interpretations. Additionally, the low permeability requires slow loading rates; therefore, tests on
flowback technique can help define the fracture closure better and shales take a long time. Osmosis and hydration may be involved
also increase test efficiency. Third, a proper minifrac test should if fluids of different activity contact the shales. Shale properties
not measure an in-situ minimum stress larger than the density- will change if shale is hydrated or dehydrated. Therefore, it is nec-
derived vertical stress. If this does occur, a definitive explanation essary to determine the chemistry and activity of the pore fluid in
should be provided. the shales so that the pore fluid used in the laboratory tests is
chemically compatible.
Fig. 6 shows the X-ray diffraction (XRD) analysis results on
Geomechanical Laboratory Tests the Clearwater shales in the current shallow well. Compared with
Mechanical properties of the caprock, including the compressive the conventional hard-rock shales, the Clearwater shales above
strengths, are essential for determining the maximum thermal the target reservoir have a greater quartz content. Correspond-
operating pressure for the reservoir. Geomechanical laboratory ingly, the clay content is lower. Among the clay minerals, the
tests on the site-specific cores are the most-comprehensive and - illite/smectite (I/S) mixed layers are dominant. A higher smectite
effective means of deriving the necessary mechanical properties. content foretells a high hydration tendency. These characteristics
Mechanical tests on rocks in general are complex, and tests on are observed in the field. Without proper attention to the coring
shales are even more complicated. Attention to the complexity is mud, Clearwater shale cores can readily become hydrated when
important because shale is clay rich with low permeability. The brought to the surface. Therefore, careful attention is required

P–√Δt
1.2

Closure Pressure: 0.859 MPa


1.1

1.0

0.9
BHP (MPa)

0.8

0.7

0.6

0.5

0.4
0 5 10 15 20 25 30 35 40
Square Root of Shut-In Time

Fig. 4—The fracture closure is better defined by a clear inflecting curve at the intersection of two different slopes on a p-sqrt(time)
plot. Cycle 9 with flowback in the test is shown in Fig. 5a.

July 2013 Journal of Canadian Petroleum Technology 269


(a)
BHP@pump sensor Sv = 0.92 Hydrostatic, 0.42
2.0
Cycle 1 Smin = 0.88 Rate
1.8 140
3
5
1.6
120
7 9 10
1.4
100

Injection, L/min
1.2
80
1.0
60
0.8
40
0.6
20
0.4

0.2 0

0.0 –20
12:00 PM 12:28 PM 12:57 PM 1:26 PM 1:55 PM 2:24 PM 2:52 PM 3:21 PM

(b)
Summary
2.0

1.5
Pressure (MPa)

1.0

0.5 Preopen Fracture closure


ISIP pressure = 0.88 MPa
Pc, sqrt
Pc, compliance
0.0
0 2 4 6 8 10
Cycle

Fig. 5—(a) Pressure/rate history from a minifrac test in the Clearwater caprock shale at 43-m TVD. The negative injection rates
represent flowback. (b) Various characteristic pressures interpreted from the minifrac test shown in Fig. 5a. Other details follow
Fig. 2b.

when deciding on the coring mud and preserving the cores. tion and that for C24T1S1 from the lower part. C14T3S1 was
Hydration inhibitors added to the mud become necessary. tested under an effective confining pressure, CP¼0.2 MPa. But its
Both room-temperature and high-temperature triaxial tests peak stress is higher than C24T1S1, which was tested under a
were conducted on more than 10 samples for this project. They higher CP¼1.5 MPa. One possible cause is the difference in the
were used to derive the elastoplastic deformation properties and mineralogy between the two samples. As shown in Fig. 6, the
strength parameters to be used for the geomechanical simulations. upper part of the Clearwater caprock, where C14T3S1 came from,
A detailed description of the test procedures and results (e.g., how is more silty, while the lower part where C24T1S1 was situated
to derive the Mohr-Coulomb failure envelope from the tests) are has a higher clay content and I/S mixed layers. A similar phenom-
beyond the scope of this paper. Interested readers can refer to a enon has been reported in literature on the mechanical behaviour
paper by Xu et al. (2011) for the triaxial-test procedures and refer of soils. For example, some researchers (Bishop 1966; Kenney
to relevant publications by the ERCB for the test results. In the 1967; Mitchell and Soga 2005) documented that soil with a higher
following, the effect of high I/S contents on the mechanical clay content, especially when the clay minerals are mostly mont-
behaviour of the Clearwater shales is elaborated upon. Little morillonite, exhibits a small friction angle. Contribution from a
attention has been directed to this factor so far in the context of higher normal stress or confining pressure to the peak stress is
caprock-integrity studies, and thus it warrants a dedicated discus- muted for a smaller friction angle.
sion herein. Sedimentary rocks (e.g., the Clearwater shales in our current
Fig. 7 plots the stress/strain curves for two representative sam- discussion) rely on the friction for their mechanical strength. The
ples: C14T3S1 from the upper part of the Clearwater caprock sec- normal stress condition or confining pressure in the triaxial tests

270 July 2013 Journal of Canadian Petroleum Technology


80

70

Quartz Clay
60

Bulk content, %
50

40

30

20

10

0
m m m m m m m m m m m m m m ale ea n e e
47 .66 .87 .59 .06 .16 .18 .40 .70 .91 .94 .15 .91 .14 sh
. S gto ierr cen
2 2 2 2 22 3 2 3 3 3 3 3 8 3 8 4 4 4 4 4 9 5 0 5 6 5 7
d o r t h l l i n P o
2, 3, 1, 1, 3, 4, 3, 4, 1, 2, 2, 3, 1, 2, ra No We E
T 1S T1S T2S T3S T3S T3S T2S T2S T3S T3S T3S T3S T1S T1S olo
C
11 11 11 14 14 14 16 16 18 18 21 21 24 24 R,
C C C C C C C C C C C C C C O
I

100
Kaolinite
90 Chlorite
Illite
80
I/S mixed
Clay type/content, %

70

60

50

40

30

20

10

0
m m m m m m m m m m m m m m ale ea n e e
47 .66 .87 .59 .06 .16 .18 .40 .70 .91 .94 .15 .91 .14 sh
. S gto ierr cen
, ,
2 2
22
2 ,2 ,3 ,3 2 3
,
3 8 8 4
3 ,3 ,3 ,4 ,4 ,4 ,5 ,5 ,5 4 9 0 6 7
d o r t h l l i n P Eo
a o e
S2 S3 S1 S1 S3 S4 S3 S4 S1 S2 S2 S3 S1 S2 lor N W
1 T1 1T1 1T2 4T3 4T3 4T3 6T2 6T2 8T3 8T3 1T3 1T3 4T1 4T1 Co
1 1 1 1 1 1 1 1 1 1 2 2 2 2 ,
C C C C C C C C C C C C C C OR Conventional hard rock
I

Fig. 6—Bulk minerals and clay type/contents measured from XRD analysis on a shallow oil-sands well. “C11T1S2, 22.47 m”
denotes Sample identification and depth, respectively. Data for “IOR, Colorado shale” are from the CSS operation in Cold Lake,
Alberta (Smith et al. 2004). Data for “Conventional hard rock” are from the public literature whose sources are not listed here.

controls the strength. However, the previously described observa- late the induced stresses and deformation in the caprock caused by
tions suggest that for shallow depths, the mineralogy, in con- the evolving pore-pressure ( p) and temperature (T) conditions in
junction with the stress condition, also affects the strength the reservoir. The simulation models start with the original in-situ
significantly. For greater depths, the confining pressure is much stress condition measured in the minifrac tests. They use the me-
higher. There, the stress condition may overshadow the mineral- chanical properties extracted from the laboratory tests to describe
ogical influence. Therefore, it is important to characterize the the constitutive deformation behaviour and allow safety factors to
mineralogical variation in the Clearwater shale. Different depth be evaluated for a given thermal operating pressure. Finally, the
intervals may exhibit different clay/quartz contents. Therefore, optimal maximum operating pressure (MOP) is derived by a series
these intervals should be grouped and tested separately. Different of sensitivity-analysis runs for a given geology in a project site.
mechanical properties may be measured for each of such groups. ABAQUS (SIMULIA 2012), a large-scale commercial finite-
element stress-analysis software, was used for the simulations. It
is well-established in the industry and fulfills our requirements
Geomechanical Simulation because of its strong nonlinear solution capability and unrivaled
Geomechanical simulations are used to investigate the impact of strength in simulating the coupled thermal/hydraulic/mechanical
thermal operating conditions on the caprock integrity. They calcu- processes. For the caprock strata, our simulations are coupled

July 2013 Journal of Canadian Petroleum Technology 271


2.2

2.0

1.8

1.6

Deviatoric Stress, q (MPa)


1.4

1.2

1.0

0.8

0.6
C14T3S1: CP = 0.2 MPa
0.4

0.2 C24T1S1: CP = 1.5 MPa

0.0
0.0 1.0 2.0 3.0 4.0 5.0 6.0
Aial Strain, Ea (%)

Fig. 7—Two example stress/strain curves measured from triaxial tests on the shallow Clearwater shales. The sudden stress drops
are likely caused by localized microcracks being activated in the samples. They do not affect the ultimate sample strength.

between the fluid flow, temperature diffusion, and elastoplastic ditions calculated independently by a thermal simulator or from
deformation. The hydraulic/mechanical coupling is two-way— analytical solutions. Their induced reservoir deformation is thus
namely, the pressure diffusion and rock deformation influence simulated by ABAQUS. Reservoir dilation is automatically
each other fully. But the thermal/mechanical coupling is one-way accounted for by the chosen elastoplastic model and input dilation
(i.e., the temperature diffusion affects the rock deformation only, angle. But the thus-calculated reservoir deformation is not allowed
while the mechanical deformation does not affect the thermal to feed back into the thermal simulation, affecting the p and T con-
properties and temperature distribution). Temperature is allowed ditions in the reservoir. It is our opinion that such one-way cou-
to diffuse into the caprock because of the SAGD temperature con- pling from the p and T conditions to the mechanical response in
ditions in the reservoir. But no thermal convection is considered. the reservoir is an adequate approximation and the most cost-
This is valid given the low permeability and thus the slow fluid- effective if the major focus is on the caprock deformation only.
flow velocities in the caprock. The following discussion will compare the induced stresses in
Within the reservoir layer, the coupling is only one-way: Our the caprock for two different reservoir depths: 60 and 200 m. The
specially-designed ABAQUS driver enters the SAGD p and T con- Clearwater shale is assumed to be in the horizontal-fracture stress
regime where the vertical stress Sv is the minimum component.
As a reference point, an MOP at 80% of Sv is at the given reser-
Stress-free ground surface voir depth. Thus, at 60 m, MOP¼80%21 kPa/m60 m 1
MPa. The corresponding steam temperature Ts is assumed at
1818C following the phase behaviour of steam at the 1-MPa pres-
sure. Similarly, at 200-m depth, MOP¼3.36 MPa and Ts ¼2458C.
Fig. 8 shows a conceptual geomechanical model used for the
Clearwater current discussions. An oil-sands pay zone 27 m thick is sand-
caprock, ho wiched between the overburden and underburden. For the sake of
simplicity, the whole overburden is assumed to be the Clearwater
formation and the underburden is taken to be a limestone-like ma-
terial. The model is a 2D plane strain cut perpendicular to the hor-
z izontal SAGD wells. The top of the model is stress free,
Oil-Sands
representing the real ground surface, while the bottom of the
pay zone, hr model is constrained from moving in the vertical direction,
x
reflecting the support from underneath. It is an artificial boundary
condition, but it does not affect the computations in the pay zone
Underburden and overburden because the bottom of the model is sufficiently far
from the area of interest. The left and right sides of the model are
a symmetrical boundary type, which reflects the physical reality
because the left side of the model cuts vertically through the mid-
dle of the SAGD wells and the right side is at the midway between
two neighbouring SAGD well pairs.
Material properties used in the previously described simula-
tions are listed in Table 1. For simplicity of presentation, a linear
elastic model is used for the over- and underburden. But a
Fig. 8—A schematic model used for the geomechanical simula- Drucker-Prager (D-P)-type plasticity model was used to account
tions and analytical derivations concerning induced stresses for the shear-induced dilation in the reservoir. In addition, a po-
and deformation in the caprock caused by SAGD steaming in rous nonlinear elastic model is used in the reservoir to account for
the reservoir. the pressure-dependent elastic moduli (SIMULIA 2012). A

272 July 2013 Journal of Canadian Petroleum Technology


In order to calculate the evolving p and T distributions in the
TABLE 1—MATERIAL PROPERTIES USED IN reservoir, an analytical solution given by Butler and Stephens
GEOMECHANICAL SIMULATIONS (1981) and Butler (1991) is used. Fig. 9 plots a steam-chamber
shape calculated from the solution and imported into the geome-
Overburden chanical simulations. It captures the general steam-chamber shape
Young’s modulus (MPa) 200 during the SAGD operation. However, differences do exist in the
Poisson’s ratio 0.25 evolving p and T conditions if they are calculated by a numerical
Volumetric thermal-expansion coefficient (1/ C) 8.4010–5 thermal simulator. It is expected that the differences are not im-
Pay zone portant for the current objective of evaluating the impact of differ-
Porous bulk modulus* 0.007 ent reservoir depths on caprock integrity.
Poisson’s ratio 0.4
Volumetric thermal-expansion coefficient (1 /C) 2.0010–4 Induced Deformation in the Caprock. Reservoir expansion
Drucker-Prager friction angle (degrees) 45 caused by the temperature increase is shown in Fig. 9 by magnify-
Drucker Prager cohesion (kPa) 62 ing the deformed meshes. It shows the caprock is being bent by the
Drucker-Prager dilation angle (degrees) 20
expanding reservoir underneath. The surface heave, which may be
a more familiar term in the industry, is given in Fig. 10. It shows
Underburden
the evolving heave directly above the SAGD well pair during the
Young’s modulus (MPa) 1000 SAGD operation in both shallow and deep reservoirs. For these
Poisson’s ratio 0.25 simplified calculations, as much as 90 cm for the shallow reservoir
* Logarithmic bulk modulus for porous elasticity in ABAQUS (SIMULIA 2012). and 130 cm for the deep reservoir are registered at the surface after
approximately 10 years of SAGD operations. These large calcu-
lated surface heaves are a direct consequence of the large thermal-
parameter called logarithmic bulk modulus is introduced for the expansion coefficient aT¼210–4/ C being used in the simula-
oil sands (Table 1). For readers who are more familiar with the tions. As shown in Fig. 10, a smaller thermal-expansion coefficient,
conventional linear elasticity theories, a logarithmic bulk modulus aT¼410–5/ C reduces the surface heave to 30 cm for the 60-m-
of 0.007 corresponds to a Young’s modulus of 142 MPa at a con- deep reservoir. A higher steam temperature is responsible for the
fining pressure of 1 MPa if other relevant material properties fol- larger deformation observed for the deeper reservoir. As analyti-
low those given in Table 1. It should be noted that the coefficient cally derived in Appendix A, the thermally induced displacement
of thermal expansion for reservoir sands came from the upper is proportional to the thermal expansion. A large thermal-expan-
bound of the high-temperature triaxial tests conducted by Agar sion coefficient and/or a large temperature increase means more
et al. (1986). Their laboratory results indicated that, at 100 C, the thermal expansion and, correspondingly, greater surface heave.
coefficient of volumetric thermal expansion for the oil sands is Induced Stresses in the Caprock. Reservoir expansion indu-
2.410–4/ C in the undrained tests and 410–5/ C in the drained ces additional stresses inside the caprock. In fact, it is these
tests. Also, this simplified numerical simulation did not consider induced stresses that are responsible for a caprock failure if it is to
the glaciation-induced slickenside structure inside the Clearwater occur. Stress paths are normally used to depict the evolving state
shale formation. Thus, cohesion is used to simulate the rock-ma- of stresses at a given location. Two elements in the caprock above
trix deformation and failure. More-sophisticated numerical mod- the SAGD chamber are used for the illustration as shown in
els are needed to simulate these weak slickenside structures, Fig. 11: Elem 1 is closer to the top of the reservoir than Elem 2.
which is beyond the scope of this paper. Two failure envelopes of D-P type are also plotted for reference,
In order to explain some of the numerical-simulation results, representing two different friction angles: 30 and 45 , respec-
analytical derivations are given in Appendix A for the induced tively. All of them use a D-P cohesion¼0.3 MPa, notwithstanding
stresses and vertical displacement in an interval caused by a uni- that 45 is stronger than 30 . Failure occurs when the stress path
form temperature or pore-pressure increase. Despite the simplicity touches the failure envelope. Because elastic deformation is
of the model, the analytical expressions help greatly to explain assumed in the caprock, it is possible that the stress path goes
how various material and/or geometric parameters affect the rele- beyond the failure envelope. If the plastic deformation model is
vant caprock-integrity issues. used, the stress path can only follow the failure line.

NT11 S, S11
+1.810e+02 (Avg: 75%)
+1.666e+02 +2.329e+01
+1.522e+02 +0.000e+00
+1.378e+02 –5.942e+02
+1.233e+02 –1.188e+03
+1.089e+02 Caprock –1.783e+03
+9.450e+01 –2.377e+03
+8.008e+01 –2.971e+03
+6.567e+01 –3.565e+03
+5.125e+01 –4.159e+03
+3.683e+01 –4.754e+03
+2.242e+01 –5.348e+03
+8.000e+00 –5.942e+03
–6.536e+03
–7.131e+03

Reservoir

Y Y

X Step: Step–24, SAGD process t = 1225.35days X Step: Step–24, SAGD process t = 1225.35days
Increment 17: Step Time = 53.28
Increment 17: Step Time = 53.28
Primary Var: NT11 Primary Var: S, S11
Deformed Var: U Deformation Scale Factor: +5.000e+01 Deformed Var: U Deformation Scale Factor: +5.000e+01

Fig. 9—Left: temperature distribution (NT11 in the unit of 8C) in the reservoir and the caprock after 1,225 days of SAGD operation.
Right: the corresponding effective horizontal stress (S, S11 in the unit of kPa). Negative stresses are compressive in a 60-m reser-
voir depth. The mesh is magnified by 50 times to show the deformed shape.

July 2013 Journal of Canadian Petroleum Technology 273


60 m, E = 200 MPa, Expansion = 2e–4/C
140
200 m, E = 200 MPa, Expansion = 2e–4/C
120 60 m, E = 20 MPa, Expansion = 2e–4/C

60 m, E = 200 MPa, Expansion = 4e–5/C


100

Surface Heave (cm) 80

60

40

20

0
0 500 1,000 1,500 2,000 2,500 3,000 3,500
SAGD Operating Time (Days)

Fig. 10—Temporal variation of the vertical displacement on the ground surface (also called heave) directly above the SAGD well
pair for the various sensitivity cases simulated. “60m, E5200 MPa, expansion52e–4/C” denotes, for example, a 60-m reservoir
depth, the Young’s modulus of the caprock shale of 200 MPa, and volumetric thermal-expansion coefficient of the reservoir of
2310–4/º C. The same holds for the legends in Figs. 11 and 12.

Fig. 11 demonstrates that the shear stress or deviatoric stress Lower shear stresses and a more compressive nature in the
as plotted increases in the caprock during the SAGD operation in induced stresses all favour the integrity of the caprock. Unfortu-
both shallow and deep reservoir depths. The increase is more in nately, because of its shallow depth, the initial stress state is under
the deeper reservoir (> 2 MPa) than in the shallow one (approxi- a smaller confinement in the shallow reservoir than in the deep
mately 1.5 MPa). In the case of a deeper reservoir, the stress state one. This is reflected by the fact that the original state of stresses
becomes less compressive (i.e., the stress path moves to the left for the shallow reservoir depth is located to the left of (i.e., less
from the original condition). This is caused by the higher steam compressive than) that for the deep reservoir depth. The rock
pressure and thus a higher pore pressure diffusing into the cap- strength is mostly derived from its friction. As a result, the cap-
rock. On the contrary, for the shallow reservoir, the stress state rock touches the failure envelope more readily in the shallow
moves toward more compressive or toward the right on the plot. depth than in the deep one. As shown in Fig. 11, the stress path
This reflects the low operating pressure and the dominating nature for the shallow reservoir has passed the 30 failure envelope at
of the compressive thermally induced stresses. both Elem 1 and Elem 2 and has touched the 45 envelope

6.0
30°
45°
60 m, E = 200 MPa, Expansion = 2e–4/C, Elem 1
5.0
60 m, E = 200 MPa, Expansion = 2e–4/C, Elem 2
200 m, E = 200 MPa, Expansion = 2e–4/C, Elem 1
Deviatoric Stress (MPa)

4.0 200 m, E = 200 MPa, Expansion = 2e–4/C, Elem 2

3.0

2.0

1.0

0.0
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5 5.0
Mean Effective Stress (MPa)

Fig. 11—Stress paths at two elements in the caprock above the SAGD steam chamber for the two different reservoir depths. 30º
denotes the D-P friction angle of 30º . D-P cohesion is all at 0.3 MPa. Other details follow Fig. 10. Elem 1 or 2 denotes the location
shown in the insert where the stress path is taken.

274 July 2013 Journal of Canadian Petroleum Technology


2.0
30°

1.8 45°

60 m, E = 200 MPa, Expansion = 2e–4/C, Elem 1


1.6
60 m, E = 200 MPa, Expansion = 4e–5/C, Elem 1

60 m, E = 20 MPa, Expansion = 2e–4/C, Elem 1


1.4

Deviatoric Stress (MPa)


1.2

1.0

0.8

0.6

0.4

0.2

0.0
0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0
Mean Effective Stress (MPa)

Fig. 12—Different stress paths corresponding to the sensitivity-analysis parameters denoted in the legend. The other details follow
Fig. 10 or Fig. 11.

already. For the deeper reservoir, the stress path goes beyond the stresses become significantly smaller because of the smaller
30 failure envelope at Elem 1, but stays just below, although Young’s modulus, so they are comparable to or even overshad-
close to, the 45 failure. Location of Elem 2 is safe for the two owed by the decreasing effective stresses caused by the pore-pres-
strength envelopes. sure increase. The latter makes the stress state approach the less-
Limited Sensitivity Analyses. Caprock integrity is a complex compressive state (i.e., to the left in Fig. 12). In summary, the
issue and numerous factors contribute (Yuan 2008). Ideally, a se- softer or more-compliant nature of the shallow caprock shale sug-
ries of sensitivity analyses should be run in the simulations. How- gested by the laboratory tests helps maintain the caprock integrity
ever, the major purpose of the simulation works presented in this where only the induced stresses are concerned.
paper is to compare the induced stresses in the caprock during the Thermal expansion is affected dramatically by the drained or
SAGD operation between a shallow (60 m) and deep (200 m) res- undrained heating. The pore fluid has a thermal-expansion
ervoir, respectively. Therefore, to limit the scope of this presenta- capacity generally at an order of magnitude larger than that of the
tion, the sensitivity analysis is limited to the following two rock matrix. For low-mobility porous media (e.g., the bitumen-
material parameters: (1) Young’s modulus of the caprock shale saturated oil sands), the excessively expanded fluid phase cannot
and (2) thermal-expansion coefficient of the reservoir. The analy- dissipate in time and, consequently, more volumetric expansion
sis is further limited to the shallow reservoir depth and illustrated results in the undrained heating (Agar et al. 1986) or thermally
by the changes in the surface heave (Fig. 10) and induced stresses induced pore pressure (Butler 1986). With the intention of con-
in the caprock (Fig. 12). Details are summarized next. servative design, the earlier simulations used a larger thermal-
The earlier presentation demonstrated that the mineralogical expansion coefficient for the reservoir. Fig. 10 shows that a
content of the caprock shale varies across the depth and the smaller thermal-expansion coefficient for the reservoir,
impact on the mechanical strength can be significant for the shal- aT¼410–5/ C, causes a much smaller surface heave. But Fig. 12
low reservoirs. Fig. 7 suggests that the Young’s modulus E for demonstrates that the ultimate level of the induced stresses in the
Sample C24T1S1 is approximately 20 MPa if measured by a se- caprock at the end of the SAGD operation is similar between the
cant line across the major portion of the stress/strain curve data two thermal-expansion coefficients. The stress path is simpler in
before the peak stress. The surface heave is not significantly the case of small thermal-expansion coefficients aT¼410–5/ C.
affected by the Young’s modulus of the caprock (Fig. 10). This It is a nearly straight line throughout the SAGD operation, extend-
can be explained by the analytical derivations in Appendix A. The ing right and upward. On the contrary, the larger thermal-expan-
vertical displacement caused by the thermal expansion (Eq. A-5) sion coefficient, aT¼210–4/ C, causes a nonmonotonic evolution
has no variable E involved. E does affect the poroelastically history in the induced stresses. The difference reflects the two
induced vertical displacement (Eq. A-6), which is responsible for major mechanisms causing the caprock to deform. Reservoir de-
the difference observed in the numerical calculation, as shown in formation and thermal stress take place when the temperature
Fig. 10. front diffuses into the caprock. These two mechanisms differ in
But much smaller induced stresses will result if the caprock is time and magnitude when they contribute to the caprock deforma-
softer or more compliant—having a smaller Young’s modulus tion. The similar end state of the induced stresses shown in Fig.
(Fig. 12). When E¼20 MPa, the state of stresses all stay below 12 suggests that, ultimately, the thermal stress induced by the
the strength lines (i.e., the caprock at the shallow depth is safe temperature changes in the caprock dominates the stress state
against the two strength envelopes: 30 or 45 for the D-P friction therein.
angle in Fig. 12). Again, the analytical derivations in Appendix A
can help explain the previously described sensitivity-analysis
results. The thermally induced stresses are proportional to the Conclusions
Young’s modulus E and the thermal-expansion coefficient. A The Alberta oil-sands/heavy-oil industry is exemplary and proac-
smaller E generates a smaller induced thermal stress. Also note tive in safeguarding caprock integrity. The intent is to design a
that when E¼20 MPa, the stress path moves toward the left (i.e., safe operating pressure and steam temperature condition so that
less-compressive state, as shown in Fig. 12). The induced thermal the caprock maintains the sealing capacity during thermal

July 2013 Journal of Canadian Petroleum Technology 275


stimulation that it has inherently displayed throughout geological lower strength. Mineralogical contents can vary across the
history. The design and analysis of the caprock integrity serves to depth. The impact on the mechanical properties is more pro-
weigh the balance between the induced stresses and the material nounced for the shallow depths. All of these speak for the
strength. Major activities include (1) minifrac tests to define the need of a detailed geomechanical laboratory program to mea-
initial stress state before the manual disturbance introduced during sure the mechanical properties of the caprock shale and their
the SAGD operation; (2) geomechanical laboratory tests to mea- potential variations across the depth.
sure the mechanical properties including compressive strengths of (5) Young’s modulus of the caprock shale does not greatly affect
the caprock; and (3) geomechanical simulations to calculate the the surface heave. A smaller thermal-expansion coefficient of
induced stresses. These components of a caprock-integrity pro- the reservoir reduces the surface heave and influences the evo-
gram are common to both deep and shallow reservoir depths. lution of the induced stresses in the caprock. But for the simu-
However, when applied to an oil-sands development at shallow lated cases, it appears that the ultimate stress state in the
depths, these components merit some special considerations. This caprock is not altered significantly by the thermal expansion in
paper has discussed the various quality-control issues, as follows: the underlying reservoir. It reflects the dominating role played
• Minifrac tests for oil-sands development require increased ac- by the thermal stresses in the caprock induced by the diffusing
curacy in executing the tests, acquiring the field data, and inter- temperature front from the underlying steam chamber.
preting the data. Interpretation of the density log can reliably This paper is not intended to support or argue against develop-
give the vertical stress Sv. A minifrac test, if properly executed ing a particular type of reservoir depth. In fact, it is our expecta-
and interpreted, should not yield a minimum stress larger than tion that no reservoirs should be categorically branded as safe or
Sv. unsafe for undergoing processing. Such generalizations can result
• Multiple cycles supplemented with flowback can enhance the in certain valuable bitumen resources remaining unextracted, or
quality of the minifrac tests. A consistent fracture-closure pres- conversely some resources being extracted without proper regard
sure should be observed between multiple test cycles. The flow- to protecting the groundwater aquifers and the environment. The
back can shorten the duration of each test without key is a rigorous engineering program, including the geomechani-
compromising the test quality, and help define the fracture-clo- cal engineering, to design the operation and then an effective
sure better. monitoring program to detect signals of negative impacts on the
• Clearwater shale above the McMurray oil-sands reservoirs typi- caprock integrity. Frequent updating and inversion of the moni-
cally has a higher quartz content compared with the conven- tored data will offer a continuous health check on the caprock
tional hard rock, but an I/S mixed layer is the dominant clay integrity.
type. In the case of shallow caprock depths, which are the focus It cannot be overstated that the caprock integrity is a complex
of the current paper, different clay contents affect the mechani- issue. Numerous factors contribute which may compete against
cal properties and strength significantly. Tests on samples with each other. Dominance of some factors over the others may
higher clay and I/S content showed a lower compressive change with time and in space. They all depend on geology, mate-
strength. Thus, it is important in the shallow reservoirs to group rial properties, and operating conditions. Numerous mechanisms
the rock facies properly when planning the laboratory tests and or factors have not been considered in the current paper. For
geomechanical simulations. example, the current paper does not consider the thermally
Geomechanical simulations carried out in this paper serve to induced pore pressure. It can be important in both thye reservoir
compare the induced stresses in the caprock by the underlying and the caprock shale. Second, the present paper does not con-
SAGD operation in the shallow (60 m) or deep (200 m) reservoirs. sider impact of weak planes. Examples of weak planes include
A limited sensitivity analysis is also run to investigate the impact slickensides, joints, fractures, or even thin bentonite layers. In all,
of Young’s modulus of the caprock and thermal-expansion coeffi- readers are warned not to apply the results in this paper in design-
cient of the reservoir. Some observations can be drawn as follows: ing their operation without a proper site-specific geomechanical
(1) It is appropriate to limit the SAGD operating pressure as a study program. But the methodologies and quality-control meas-
function of the reservoir depth. In this paper, the 80% propor- ures illustrated and promoted in this paper should benefit the
tionality according to the fracture-closure pressure in the cap- industry in carrying out geomechanical works.
rock is used as an example. A shallow SAGD operation
should be conducted at a lower pressure and therefore a lower Nomenclature
steam temperature. As a result of lower pressure and tempera- Cp ¼ effective confining pressure in triaxial tests, kPa
ture conditions, the induced deformation and stresses in the E ¼ Young’s modulus
caprock become smaller for the shallow reservoirs than for G ¼ shear modulus
the deep ones. This is beneficial for the caprock integrity. MOP ¼ maximum operating pressure in, kPa
(2) However, also inherent with the shallow depths is the original Pc ¼ fracture-closure pressure, kPa
lower confining stress conditions—a disadvantage for main- SHmin ¼ minimum horizontal stress, kPa
taining the caprock integrity. Consequently, despite the SHmax ¼ maximum horizontal stress, kPa
smaller induced stresses by the lower steaming temperature Sv ¼ vertical stress, kPa
and pressure in the reservoir, it is still possible in the case of Ts ¼ steam temperature in SAGD,  C
shallow reservoir depths that the caprock may enter into plas- a ¼ volumetric thermal-expansion coefficient
tic failure state. This highlights the challenges in developing eij ¼ strain tensor
the shallow reservoirs and the necessity for a rigorous cap- g ¼ Biot’s stress coefficient, g¼a(1þ)/[2(1þ)] with a
rock-integrity study before production and for monitoring pro- here being the Biot’s coupling coefficient
grams during production.  ¼ Poisson’s ratio
(3) In addition to the lower operating pressure and temperature, ui ¼ displacement vector
the integrity of the shallow caprock can be enhanced by a rij ¼ stress tensor with subscripts i ( j) running from x, y, to z
softer and more-compliant mechanical nature. The sensitivity
analysis shows that if the caprock shale has a smaller Young’s
modulus, the induced stresses are further diminished and the Acknowledgements
caprock becomes more resilient against the plastic yielding We thank the management of Alberta Oilsands Incorporated and
under the same SAGD operating pressure and temperature. BitCan G&E Incorporated for allowing the publication of this pa-
(4) The mechanical properties in a given location and depth must per. Thanks are also due to Jeffrey Tailleur, Natalie Murray, and
be adequately characterized and considered in effective mod- Donald Nicholson for editing the manuscript. Xu is partly sup-
elling. Moreover, geomechanical laboratory tests show that ported by Alberta Innovates through an industrial research and de-
some intervals can indeed have a lower Young’s modulus and velopment associate fund program.

276 July 2013 Journal of Canadian Petroleum Technology


References Technol 43 (2): 39–46. PETSOC-04-02-03. http://dx.doi.org/10.2118/
Agar, J.G., Morgenstern, N.R., and Scott, J.D. 1986. Thermal expansion 04-02-03.
and pore pressure generation in oil sands. Canadian Geotechnical Xu, B., Yuan, Y.G., and Wang, Z.C. 2011. Thermal Impact On Shale De-
Journal 23 (3): 327–333. http://dx.doi.org/10.1139/t86-046. formation/Failure Behaviors—Laboratory Studies. Presented at the
Bell, J.S., Price, P.R., and Mclellan, P.J. 1994. In-situ Stress in the West- 45th US Rock Mechanics/Geomechanics Symposium 2011, San Fran-
ern Canada Sedimentary Basin. In Geological Atlas of the Western cisco, California, USA, 26–29 June. ARMA-11-303.
Canada Sedimentary Basin (WCSB), G.D. Mossop and I. Shetsen, Yuan, Y. 2008. Overburden/Casing Integrity in SAGD without High Oper-
Chap. 29. Available online from Alberta Geological Survey, http:// ating Pressures. Presented at the 2008 Canadian International Petro-
www.ags.gov.ab.ca/publications/wcsb_atlas/atlas.html. leum Conference/Petroleum Society s 59th Annual Technical Meeting,
Bishop, A.W. 1966. The Strength of Soils as Engineering Materials. Geo- Calgary, 17–19 June. CIPC-2008-206.
technique 16 (2): 89–130.
Butler, R.M. 1986. The Expansion Of Tar Sands During Thermal Recov-
Appendix A: Thermoelastic and Poroelastic
ery. J Can Pet Technol 25 (5). PETSOC-86-05-05. http://dx.doi.org/
10.2118/86-05-05. Solutions to the Stress and Deformation in a
Butler, R.M. 1991. Thermal Recovery of Oil and Bitumen. Englewood Uniformly Heated or Pressured Formation
Cliffs, New Jersey: Prentice Hall. In order to derive analytical solutions, the reservoir is assumed to
Butler, R.M. and Stephens, D.J. 1981. The Gravity Drainage of Steam- be flat and laterally infinite with a constant thickness, as shown in
Heated Heavy Oil to Parallel Horizontal Wells. J Can Pet Technol 20 Fig. 8. The temperature increase, T, is uniform throughout the
(2): 90–96. JCPT Paper No. 81-02-07. http://dx.doi.org/10.2118/81- reservoir.
02-07. Relevant equilibrium equations according to the theory of ther-
Chhina, H.S., Luhning, R.W., Bilak, R.A. et al. 1987. A Horizontal Frac- moelasticity can be written as
ture Test In The Athabasca Oil Sands. Presented at the 38th Annual   
Technical Meeting of the Petroleum Society of CIM/SPE Annual
@rij aT E @T
 ¼ 0; . . . . . . . . . . . . . . . . ðA-1Þ
Technical Meeting, Calgary, 7–10 June. CIM 87-38-56. http:// @xj 3ð1  2Þ @xi
dx.doi.org/10.2118/87-38-56.
Chhina, H.S. and Agar, J.G. 1985. Potential Use of Fracture Technology
with the constitutive relation as
for Recovery of Bitumen from Oil Sands. Presented at the 3rd Interna-  
2G aT E
tional Conference on Heavy Crude and Tar Sands, Long Beach, Cali- rij ¼ 2Geij þ ekk dij  Tdij ; . . . . . . ðA-2Þ
1  2 3ð1  2Þ
fornia, USA, 22-31 June.
Hudson, J.A., Brown, E.T., Fairhurst, C. et al. 1993. Rock Testing and and boundary conditions of
Site Characterization. In Comprehensive Rock Engineering: Princi-
ples, Practice & Projects, Vol. 3, Chap. 17–21. Oxford, UK: Perga- ux ¼ uy ¼ 0 x ¼ 61
mon Press. rzz ¼ rxz ¼ ryz ¼ 0 z ¼ ho þ hr ; x 2 ð1; 1Þ :
Kenney, T.C. 1967. The Influence of Mineral Composition on the Residual
Shear Strength of Natural Soils. Proc., Geotechnical Conference on uz ¼ 0 z ¼ 0; x 2 ð1; 1Þ
Shear Strength Properties of Natural Soils and Rocks, Oslo, Norway,                    ðA-3Þ
Vol. 1, 123–129, TRB 00237620.
Kry, P.R. 1989. Field Observations of Steam Distribution during Injection The coordinate system is shown in Fig. 8. rij is the stress
to the Cold Lake Reservoir. In Rock at Great Depth: Rock Mechanics tensor with subscripts i ( j) running from x, y, to z. eij is the strain
and Rock Physics at Great Depth—Proceedings of an International tensor, ui is the displacement vector, aT is the volumetric thermal-
Symposium, Pau, 28–31 August 1989, V. Maury and D. Fourmain- expansion coefficient,  is the Poisson’s ratio, G is the shear mod-
traux, Vol. 3. London: Taylor & Francis. ulus, and E is the Young’s modulus.
Kry, P.R., Boone, T.J., Gronseth, J.M. et al. 1992. Fracture Orientation Using the boundary conditions at the ground surface (Eq. A-3),
Observations from an Athabasca Oil Sands Cyclic Steam Stimulation it is easy to show that
Project. Presented at the 43rd Annual Technical Meeting of the Petro-
rxz ¼ 0
leum Society of CIM, Calgary, 7–10 June. CIM 92-37.
Mitchell, J.K. and Soga, K. 2005. Fundamentals of Soil Behavior. New
ryz ¼ 0 : . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ðA-4Þ
York: John Wiley & Sons. rzz ¼ 0
Palmgren, C., Walker, I., Carlson, M. et al. 2011. Reservoir design of a
The remaining solution variables are
shallow LP-SAGD project for in-situ extraction of Athabasca bitumen.
 
Presented at the 2011 World Heavy Oil Conference (WHOC11), Ed- aT E
monton, Alberta, Canada, 14-17 March. WHOC11-520. rxx ¼ ryy ¼  T
3ð1  Þ
Proskin, S.A., Scott, J.D., and Chhina, H.S. 1990. Current Practice in the   ; . . . . . . . . . . . . . . . . . ðA-5Þ
Interpretation of Microfrac Tests in Oil Sands. Presented at the SPE aT ð1 þ Þ
Duz ¼ Thr
California Regional Meeting, Ventura, California, USA, 4–6 April. 1
SPE-20040-MS. http://dx.doi.org/10.2118/20040-MS.
Raaen, A.M. and Brudy, M. 2001. Pump-in/Flowback Tests Reduce the Esti- in which Duz is the vertical displacement in the system which is
mate of Horizontal in-Situ Stress Significantly. Presented at the SPE An- reflected on the ground as the surface heave. Upward movement
nual Technical Conference and Exhibition, New Orleans, 30 September–3 or tensile stresses are positive.
October. SPE-71367-MS. http://dx.doi.org/10.2118/71367-MS. A similar result can be derived for the poroelastic solution
Raaen, A.M., Skomedal, E., Kjørholt, H. et al. 2001. Stress determination (i.e., stress changes and displacements in the reservoir caused by
from hydraulic fracturing tests: the system stiffness approach. Int. J. a uniform pore-pressure change, p):
Rock Mech. Min. Sci. 38 (4): 529–541. http://dx.doi.org/10.1016/
rxx ¼ ryy ¼ 2gp
S1365-1609(01)00020-X.  
Settari, A. and Raisbeck, J.M. 1979. Fracture Mechanics Analysis in In- g : ..................... ðA-6Þ
Duz;i ¼ hr p
Situ Oil Sands Recovery. J Can Pet Technol 18 (2): 85–94. PETSOC- G
79-02-07. http://dx.doi.org/10.2118/79-02-07.
SIMULIA 2012. ABAQUS Version 6.12 documentation, http:// Yanguang Yuan is both a professional engineer and a geo-
www.3ds.com/products/simulia/support/documentation/. scientist in Alberta. In his more than 25 years of experience, he
Smith, R.J., Bacon, R.M., Boone, T.J. et al. 2004. Cyclic Steam Stimula- has consulted with petroleum companies in North America,
tion Below a Known Hydraulically Induced Shale Fracture. J Can Pet the Middle East, and Far East, and has authored or

July 2013 Journal of Canadian Petroleum Technology 277


coauthored more than 150 technical reports and papers, one Claes Palmgren is Vice President – Resource Development,
monograph, and a 600-page course note on petroleum geo- Technical at Alberta Oilsands Incorporated. He has more than
mechanics. In 2000, Yuan founded BitCan Geosciences & En- 20 years of experience with design, development, and imple-
gineering Incorporated, which has now grown into being the mentation of bitumen and extraheavy-oil in-situ production
only independent integrated petroleum geomechanics firm in technologies at the laboratory scale, the field pilot scale, and
the world. He is a member of SPE. for larger commercial projects. Palmgren also has manage-
ment and executive experience with pioneering oil-sands
Bin Xu is a senior geomechanics specialist in research and de- companies and development of companies after working as
velopment with BitCan Geosciences & Engineering, where he reservoir-engineering lead during Suncor’s Mackay River
is involved in geomechanical numerical simulations, labora- startup and as overall lead for Petro-Canada vapour-assisted
tory testing, and DFIT/minifrac testing. Xu holds a PhD degree petroleum-extraction pilot. He holds MSc and PhD degrees
from the University of Calgary and MSc and BSc degrees from from Delft University of Technology’s faculty of Mining and Pe-
Tongji University in China. troleum Engineering in the Netherlands.

278 July 2013 Journal of Canadian Petroleum Technology

You might also like