Download as pdf or txt
Download as pdf or txt
You are on page 1of 20

SPE-170130-MS

Quantifying the Uncertainty Associated With Caprock Integrity during


SAGD Using Coupled Geomechanics Thermal Reservoir Simulation
Varun Pathak, David Tran, and Anjani Kumar, Computer Modelling Group Ltd.

Copyright 2014, Society of Petroleum Engineers

This paper was prepared for presentation at the SPE Heavy Oil Conference-Canada held in Calgary, Alberta, Canada, 10 –12 June 2014.

This paper was selected for presentation by an 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
With a growing emphasis on safety of thermal operations in the Canadian oilsands operations, the focus
of the industry on caprock integrity during SAGD has been increasing. A SAGD operation may create a
high degree of deformation in the reservoir rock because of the stresses induced by pressure and
temperature. This might yield to a failure of both the reservoir rock as well as the caprock. Thus, there
is a need to understand the impact of various geomechanical and operational parameters on overall safety
of a SAGD operation.
In this paper, caprock failure during SAGD was modeled using a coupled simulator to integrate
reservoir flow with geomechanical deformation. Representative Canadian oilsands properties were used
for the simulation. Both shear and tensile failures were modeled and the surface heave/deformation was
recorded.
Sensitivity analyses and uncertainty assessments were performed to understand the effect of various
geomechanical properties on caprock integrity and surface heave, including Young’s modulus, Poisson’s
ratio and thermal pore pressure. At the same time, operating conditions such as well spacing and
maximum injection pressure were also altered to see their effect on caprock integrity. This was done for
different geological scenarios and different reservoir depths. Based on this study, we attempt to develop
a generic workflow for quantifying the uncertainty associated with caprock integrity and minimizing the
risk associated with caprock failures in any SAGD project.
Introduction
SAGD has been one of the most successful techniques for the extraction of bitumen in Alberta over the
past several years. One of the important features of a safe and successful SAGD operation is the integrity
of the caprock, typically shale. During steam injection, pressure and temperature of the SAGD reservoir
change very rapidly, which leads to a change in the stress state of the reservoir. This leads to deformations
in the reservoir and the caprock. As a SAGD process continues, the average reservoir pressure increases
and this may lead to a reduction in effective stress and shear strength of the reservoir rock. As a result,
the reservoir rock undergoes deformation and this leads to dilation of the sand and enhancement of
reservoir permeability. This is helpful for the evolution of the steam chamber as this promotes steam
injection, and has been traditionally simulated using the popular dilation-recompaction model (Beattie et.
2 SPE-170130-MS

al, 1991). However, this will be beneficial only if the caprock is able to endure the shear stresses
developed just below it inside the reservoir (Uwiera-Gartner et al., 2011). There have been a number of
examples of various SAGD operations where caprock integrity has been broken such as Total’s Joslyn
Creek SAGD project (ERCB, 2010). Laboratory tests on shale are very complicated and time consuming
because of very low permeability values, which requires slow loading rates (Yuan et al., 2011). As a
result, the role of simulation becomes quite important.
A number of efforts have been made to simulate and analyze caprock integrity using various reservoir
and geomechanical simulators (Khan et al., 2011; Uwiera-Gartner et al., 2011; Chin et al., 2011, Yuan et
al., 2011, Chin et al., 2012). Some of the studies have been aimed at evaluating caprock integrity for gas
storage type projects (Chin et. al, 2011; Tran et al., 2009). Other studies have been targeted towards
determining the probability and potential risk of caprock failure under a given operating scenario or to
determine the maximum operating pressure specifically in case of SAGD (Grizzly Oil Sands, 2010;
Uwiera-Gartner et al., 2011; Chin et al., 2012; Walters et al., 2012).
Caprock failure can happen in SAGD because of the changes in stresses caused by the pressure and
temperature change in oilsands. It can also happen because of the existence of thermal pore pressure inside
the shale, which is caused by the heating up of the water inside shale. Water expands more than the shale
matrix upon heating and it cannot flow out because of the low permeability of shale – creating high pore
pressure inside the shale (Yang et al., 2013).
In addition to the injection pressure, the current study aims to consider other operating parameters too,
such as the vertical separation between the injector and producer well in a SAGD operation and the lateral
spacing between well pairs. Besides, there is an uncertainty in reservoir flow and geomechanical
properties which should also be considered while designing safe and optimal operating conditions for a
project. A reservoir engineer’s objective is to find the optimum operating scenario to maximize the
net-present-value (NPV) of the project. This should always be coupled with another objective, which is
to ensure the safety of caprock under geological uncertainty.
A simple synthetic model has been used in this study to highlight and emphasize these points. The
workflow starts with doing a sensitivity analysis to understand the relative effects of various operating
parameters and focuses on the effect of injection pressure in particular. Subsequently, an optimization
study is done to come up with an economical and safe operating scenario, which might mean some
compromise on the recovery in order to maintain caprock integrity. Lastly, an uncertainty analysis is
performed which is essential to quantify the effect of uncertainty in geomechanical and reservoir
properties on caprock integrity.

Calculations
Constitutive model for oilsand
A hyperbolic non-linear constitutive model was used to represent the relation between stress and strain in
oilsand (Duncan and Chang, 1970; Settari et al., 1993). In this model, tangential modulus (similar to
Young’s modulus) and bulk modulus change with changes in minimum principle effective stress and
temperature. Since the elastic moduli are varying, the Poisson’s ratio varies too. Following equations
govern these changes:
(1)

Where,
Ei ⫽ Initial tangential modulus
␴r ⫽ Reference stress
Patm ⫽ Atmospheric pressure
T ⫽ Current temperature
SPE-170130-MS 3

To ⫽ Initial temperature
The reference stress is given by:
(2)

Where,
c ⫽ Cohesion
␸max ⫽ Max. value of friction angle
⫽ Minimum principle effective stress.
Ke, ne, nte and m are the various parameters of this model.
The tangential modulus Et is given by the following relation:
(3)

Where is the maximum principle effective stress; Rf (failure ratio), n2 and n1 are model parameters.
(4)

␸o ⫽ Friction angle at confining pressure of 1 atm


⌬␸ ⫽ Friction angle change per log cycle of minimum principle effective stress
The bulk modulus Bm is given by the equation:
(5)

Where Kb, nb and ntb are all model parameters.


Lastly, oilsand permeability change due to geomechanical effects was modeled based on an empirical
relation (Touhidi-Baghini, 1998)
(6)

Where,
K ⫽ Current permeability
Ko ⫽ Initial permeability
␧v ⫽ Volumetric strain
Cn is an experimental parameter. This formula makes the reservoir rock permeability a function of
volumetric strain.

Criteria for failure of caprock


Consider the Mohr circle and failure envelope as shown in Figure 1. A safety factor can be defined to
estimate the proximity of a rock to shear failure. This is based on the distance of the stress state of the
rock from the shear failure envelope.
From Figure 1, this factor can be defined as: Safety Factor ⫽ CB/CA The Mohr-Coulomb failure
envelope can be represented by the straight line:
(7)

Where c is the cohesion, ␸ is the friction angle, ␶ is the shear stress at failure and ␴ is the normal
effective stress on the failure plane.
Co-ordinates of point C are: and the radius of the Mohr’s circle is: ; where and are
the maximum and minimum principle effective stresses respectively.
Point B lies on the Mohr-Coulomb failure line and has the co-ordinates:
4 SPE-170130-MS

Figure 1—Mohr-Coulomb circle and Mohr-Coulomb failure envelope

Figure 2—Typical geology used in the models. Caprock failure was studied at the point ‘X’

Thus, the length of the segment CB is:


From ⌬ BCB=, CB ⫽ CB= cos ␸; therefore,
(8)

A value of 1 or lower for this safety factor would indicate shear failure of the rock. A value greater
than 1 would indicate that the rock is still in the elastic region of deformation.
Tensile failure was modeled too and its identification was based on the value of minimum principle
effective stress.
Experimental design and response surface based sensitivity study
Several sensitivity runs were done using CMG’s CMOST software in conjunction with CMG Builder and
CMG STARS. CMOST creates several datasets using experimental design concepts to best represent the
search space and then creates a proxy model using response surface methodology. A reduced quadratic
model has been used to represent the response surface which relates the objective function (y) to the
various parameters (xj) using the following equation,
SPE-170130-MS 5

Table 1—Geomechanical parameters assumed for this study


Parameter Value Description

Ke 448.3 Constant in tangential modulus equation (1)


Kb 747.2 Constant in bulk modulus equation (5)
ne 0.882 Exponent in tangential modulus equation (1)
nb 0.882 Exponent in Bulk modulus equation (5)
Rf 0.3 Failure ratio in equation (3)
c 101.325 Cohesion, kPa in equations (2) and (3)
␸o 63° Friction angle at 1 atm confining pressure in equation (4)
⌬␸ 20° Friction angle change per log cycle in equation (4)
␸max 53° Maximum value of friction angle in equation (2)
Patm 101.325 Atmospheric pressure, kPa in equations (1), (4) and (5)
␣ 2.00E-05 Thermal expansion coefficient, 1/°C
Cn 5 Exponent in relation for permeability, equation (6)

Table 2—Shale geomechanical parameters assumed for this study


(9)
Parameter Value Description

E 1.35E⫹06 Young’s modulus, kPa (Constant)


This model takes into account the linear effects v 0.32 Poisson’s ratio (Constant)
␸ 30° Friction angle (Constant)
of various parameters on the objective function
␣ 1.60E-06 Thermal expansion coefficient, 1/°C
(through linear terms xj), the interaction effects c 300 Cohesion, kPa
(through cross terms xixj and the quadratic effects
(through quadratic terms xj2). For example, this
methodology can give a relation between an objec-
tive function like surface heave and parameters such as injection pressure, lateral well-pair spacing etc.
The effect of parameters individually, as well their combined effect can be obtained.

Reservoir Properties
The geological configurations used for reservoir models included typical shale and sand properties of
Albertan oilsands reservoirs. For all the models, 30 m thick good quality sand was considered (similar to
McMurray sand). On top of this, a 10 m thick mudstone layer was considered (similar to Wabiskaw). And
the caprock was considered to be 10 m thick shale (similar to Clearwater shale). For all the models, an
element-of-symmetry approach was used, i.e. only half well-pair (lateral) spacing was considered. The
half well-pair spacing was 60 m for the base case. Moreover, the analyses were done using 2D
sub-models. From a 3D model with well length of 1000 m (20 J-layers x 50 m thickness of each J-layer),
one layer was cut off representing 1/20* of the full model. The injection and production rates were
adjusted accordingly. This was done to save simulation time. A 2D approach is justified physically
especially if the heterogeneity in reservoir properties is not prominent. Most of the models used for
analyses comprised of layer-cake type geometry. However, this workflow can be easily extended and
performed on 3D models consisting of complex geological heterogeneity. The schematic of the reservoir
has been shown in Figure 2. Two reservoir top depths were used in this study to represent a shallow and
a deep reservoir.

Geomechanical Properties
Two-way coupled models were used for this study. In this coupling scheme, at each time step, the fluid
flow calculations are performed by the reservoir flow simulator – providing the values of pressure and
temperature. These values become the input for the geomechanics module, which then calculates the
resulting deformation, strain change and stress change. Subsequently, the geomechanics module updates
6 SPE-170130-MS

Table 3—Overburden geomechanical parameters assumed for this study


Georock 1 Georock 2 Georock 3

Parameter Shallow Deep Shallow Deep Shallow Deep

Depth range, m 0–120 0–60 120–240 60–120 240–300 120–150


E, kPa 1.00E⫹05 1.60E⫹05 4.50E⫹05
v 0.28 0.35 0.35
␸ 30° 30° 30°
␣, 1/°C 2.0E-06 1.60E-06 1.60E-06
c, kPa 300 300 300

Figure 3—Initial stress distribution for shallow reservoir

the values of porosity and permeability – which are then communicated back to the reservoir flow
simulator. The reservoir flow simulator updates the values of pressure and temperature and provides these
values for the geomechanics module. This continues till a convergence of the solution has been achieved.
Then the simulator proceeds to the next time step.
For sand, shale and overburden, the following geomechanical properties were considered:
1. Sand – As discussed before, the hyperbolic non-linear elastic model was used to represent the
oilsands. This model has been used by various companies and research groups before and generic
values of the different model parameters for McMurray sand are available in literature (Grizzly
Oil Sands, 2010; Uwiera-Gartner, 2011). The assumed values of various geomechanical param-
eters in equations (1) through (6) are given in Table 1.
2. Shale and overburden – Mohr-Coulomb type linear elasto-plastic constitutive models were used
for shale and overburden. Material properties of shale were taken from the Clearwater shale
properties available in literature (Gu et al., 2009) and are listed in Table 2. Three geomechanical
rock types were used to represent the overburden and their respective properties are given in Table
3 for both, the shallow and the deep reservoirs.
The assumed distributions of initial stresses are shown in Figure 3 and Figure 4.
The boundary conditions used in this study are shown in Figure 5.
SPE-170130-MS 7

Figure 4 —Initial stress versus depth for deep reservoir

Figure 5—Boundary conditions used in this study

Table 4 —Base model properties for the cases studied to see effect of operating parameters
Inj. BHP Lateral well-pair Inj-Prod
Case# Model top depth (m) Shale top depth (m) (kPa) Steam temp. (°C) spacing (m) separation (m)

1 150 150 2000 212.3 120 5


2 300 300 4000 250.3 120 5
3 0 150 2000 212.3 120 5
4 0 300 4000 250.3 120 5

Results and Discussion


The objective functions considered for this study were:
1. NPV (Net Present Value): calculated at the end of 10 years, based on steam injection and oil
production rates. The costs considered were $314.5 per m3 ($50 per bbl) for oil revenue and
$50.32 per m3 ($8 per bbl) for steam cost. The discount rate was assumed as 10%.
2. cSOR (Cumulative SOR): Ratio of volume of steam injected to volume of oil produced
(cumulative values at the end of 10 years).
3. Time taken for caprock to fail: This was considered as the time when the top layer of caprock
(point ‘X’) reached a safety factor of 1.1 or lower, i.e. the point ‘X’ is within 10% of failing.
4. Top heave: This was the heave measured at the top of the caprock (point ‘X’) at the end of 10
years. Surface heave was also used as an objective function when the overburden was modeled.
8 SPE-170130-MS

Figure 6 —Relative effects of various operating parameters on the heave at the top of shale for shallow reservoir

5. Cumulative oil: The cumulative produced oil volume at the end of 10 years.
The observations have been summarized according to the various cases in this section.
Sensitivity Study: Effect of operating parameters
Initially, four variations of above mentioned reservoir description were used to understand the effect of
various operating parameters on caprock integrity. For these cases, the geomechanical and reservoir
properties were kept constant. The base models for each of these cases had the following properties:
The operating parameters that were varied were:
1. Injector bottom-hole pressure(BHP): Injectors were constrained to inject on bottom-hole pressure
constraint. A CWE steam injection rate constraint was also provided but it was never achieved.
2. Injector-producer separation/Location of injector: The producer was kept fixed at a depth of 50
m from top of shale layer (i.e. completed in the bottom-most layer of the model), and injector
location was moved from a depth of 44 m to a depth of 46 m (measured from shale top in all
cases), thereby providing a an injector-producer separation of 4 m to 6 m.
3. Lateral well-pair spacing: The boundaries of the models were moved to provide a lateral
separation between well pairs from 60 m to 200 m (i.e. half well-pair spacing of 30 m to 100 m).
Case-1 and Case-2:
Shallow reservoir and deep reservoir respectively, overburden not modelled: The results of both these
sensitivity analyses showed that the failure of the caprock is highly sensitive to the injection well BHP.
Higher injection well BHP might lead to a quicker breach of caprock integrity. This can be understood
from the tornado charts for both, the time taken for the top of the caprock to fail as well as the heave at
the top of shale. The tornado chart for the heave at the top of shale caprock is shown in Figure 6, which
demonstrates the relative effect of various operating parameters on the objective function. From Figure 6
(R-squared value of 0.96), the maximum value of heave was about 0.40 m. The figure shows that when
SPE-170130-MS 9

Figure 7—Relative effects of various operating parameters on the NPV from the model for shallow reservoir

the injection pressure was changed from about 1700 kPa to about 3300 kPa, an additional heave of about
0.24 m was created at the top of the shale layer. It also shows that increasing the well spacing caused a
slight reduction in heave at the top of the shale. It must be noted that there are some statistically
insignificant terms in the tornado chart and are shown here just for presentation. These will not have much
effect on the objective function. Additionally, some terms from equation (9) are even more statistically
insignificant and are not shown in these plots. For example, the location of injector well (between 44 m
– 46 m from top of shale) did not create a big impact on the heave at the top of the shale.
Also as an example, the effect of the interaction terms and quadratic terms can be understood by
looking at the effect of ‘Inj. BHP*Inj. Location’ on the heave in Figure 6. This means that:
● If we assume ranges of the parameters ‘Inj. BHP’ and ‘Inj. Location’ as (low1, high1) and (low2,
high2)
● Highest heave will be achieved at (Inj. BHP⫽low1, Inj. Location⫽high2) or (Inj. BHP⫽high1,
InjLocation⫽low2)
● Lowest heave will be achieved at (Inj. BHP⫽ low1, Inj. Location⫽low2) or (Inj. BHP⫽high1,
InjLocation⫽high2)
● The difference between these two values of heave was 0.05 m

The effect of these parameters on the NPV is shown in Figure 7 (R-squared value of 0.94). This shows
that NPV can be increased by over $0.79 million if the injector BHP is increased from about 1700 kPa
to about 3300 kPa. This also shows that the lateral well-pair spacing and the position of the injector well
had a small effect on NPV. The cross plot in Figure 8 shows that in general, there was a less chance for
caprock failure to occur if the heave at the top of shale was lower (inverse relation between heave at the
shale top and time taken for the caprock to fail). However, the NPV increased with increasing heave of
the caprock (direct relation between heave at the shale top and the NPV).
10 SPE-170130-MS

Figure 8 —Cross plot showing relations between heave at shale top and corresponding time taken for caprock to fail (red) and heave at shale top and
corresponding model NPV (blue)

Figure 9 —Relative effects of various operating parameters on the heave at the surface (overburden modelled) for shallow reservoir

Case-3 and Case-4:


Shallow reservoir and deep reservoir, overburden modelled: In these two cases, results obtained were
similar to those obtained in the first two cases. As seen in Figure 9 and Figure 10, the maximum surface
heaves recorded were 0.59 m for the shallow reservoir (R-squared value of 0.95) and 0.68 m for the deep
reservoir (R-squared value of 0.98). Additionally, injector BHP still was the most important parameter on
both, the caprock failure as well as on the NPV of the model.
This creates a dilemma in optimization. A higher NPV might not equate to a safe SAGD operation as
shown by the results above. The solution to this dilemma might lie in carefully analyzing the operating
SPE-170130-MS 11

Figure 10 —Relative effects of various operating parameters on the heave at the surface (overburden modelled) for deep reservoir

Figure 11—Comparison of caprock state between Model A and Model B

scenarios and come up with one which can provide a profitable and safe operation. As an example, two
cases from the above study (Case-4), i.e. deep reservoir, have been presented in Figure 11 and Figure 12.
For convenience, these models have been named as ‘Model A’ and ‘Model B’. A summary of these two
models is given in Table 5.
12 SPE-170130-MS

Figure 12—Comparison of yield state in shale caprock between ‘Model A’ and ‘Model B’ after about 2620 days

Table 5—Operating parameters and corresponding results from Model A and Model B
Inj BHP Lateral well-pair Inj-Prod
Model (kPa) Steam temp. (°C) spacing (m) separation (m) Model NPV (M$) cSOR (m3/m3)

A 5500 269.9 100 6 1.03 4.34


B 5090 265.0 100 6 1.01 4.20

Table 6 —Operating parameter ranges for opt1imization study


Case-3 (Shallow) Case-4 (Deep) Looking at the results of ‘Model A’ and ‘Model
B’ above, it is clear that in both the operating
Parameter Min. Max. Min. Max. scenarios, results were very similar. Both the NPV
Inj. BHP (kPa) 1750 4000 3600 6000 and cSOR from these two models were within 5%
Lateral well-pair spacing 60 200 60 200 of each other. However, in one case (‘Model A’) the
Inj-Prod separation (m) 4 6 4 6 caprock underwent shear failure after about 2700
Steam temperature (°C) Steam saturation temperature
days (Shown by Yield State ⫽ 1 and Safety Factor
⫽1, representing shear failure), and in the other case
(Model B) there was no caprock failure for 10 years
(Shown by Yield State ⫽ 0 and Safety Factor ⬎ 1).
Figure 12 shows that after 2620 days, the caprock in ‘Model A’ has failed right above the well pair (failure
initiated at about 2600 days), whereas the caprock in ‘Model B’ did not. Both models showed some degree
of shear failure in the clean sand, which is not unexpected because of the possibility of sand dilation under
injection pressures. The main cause of this difference is the higher injection pressure and temperature in
‘Model A’.
Optimization Study: Finding safest and economical operating scenario
It is evident from the sensitivity analysis that the effect of operating conditions, especially the injector
BHP, on the integrity of caprock can be immense. As a result, any optimization study undertaken to
maximize the profit from the field must pay attention to the safety aspect of the operation. The most
economical operating scenario might not be the safest one, and as a result a compromise might be needed
to ensure safe operation of the project.
SPE-170130-MS 13

Figure 13—Model NPV and time taken for caprock to fail for all optimization jobs conducted for shallow reservoir

Figure 14 —Model NPV and time taken for caprock to fail for all optimization jobs conducted for deep reservoir

For this study, optimization was done on the NPV of the model while keeping in mind the safety factor
and yield state (i.e. the caprock integrity parameters). From here on, only those models have been used
for analysis where overburden has been modelled. So, the base cases from Case-3 and Case-4 (See Table
3) were taken as the starting point for optimization. The operating parameters were varied in following
ranges:
The relationship between the NPV of the model and the time taken for caprock integrity to be breached
is shown in Figure 13 and Figure 14 for the two reservoirs. The ideal situation will be to reach to a
maximum value on both axes, i.e. maximum NPV with no failure of caprock for 10 years (maximum time
taken for caprock to fail ⫽ length of simulation ⫽ 3652 days).
14 SPE-170130-MS

Figure 15—Effect of injection BHP and lateral well-pair spacing on the objective functions for 5 m spacing between injector and producer for shallow
reservoir

Table 7—Results of optimization study


Parameter Case-3 (Shallow) Case-4 (Deep)

Max Inj. BHP or MOP (kPa) 2800 5000


Lateral well-pair spacing (m) 80 80
Inj-Prod separation (m) 6 6
Max. NPV with caprock integrity ($) Steam saturation temperature

Table 8 —Geomechanical property ranges for uncertainty analysis


Parameter Low High Mean St. Dev. Description

Sand
Ke 300 600 450 64.5 Constant in tangential modulus equation (1), kPa
Kb 600 1000 800 86 Constant in bulk modulus equation (5), kPa
ne (⫽nb) 0.5 1 0.75 0.11 Stress exponent in equations (1) and (5)
nte (⫽ntb) 0 1 0.5 0.21 Temperature exponent in equations (1) and (5)
Rf 0.2 1 0.6 0.17 Failure ratio in equation (3)
⌬␸ 15° 25° 20° 2.15° Friction angle change per log cycle in equation (4)
␣ 1.50E-05 3.00E-05 2.25E-05 3.22E-06 Thermal expansion coefficient, 1/°C
c 5 100 52.5 20.4 Cohesion of sand
Cn 1 10 5.5 1.93 Exponent for permeability in equation (6)
Shale
E 8.00E⫹05 2.00E⫹06 1.40E⫹06 2.57E⫹05 Shale Young’s modulus, kPa
v 0.25 0.4 0.325 0.032 Shale Poisson’s ratio
c 200 700 450 107.5 Cohesion of shale
␣ 8.00E-07 3.00E-06 1.90E-06 4.73E-07 Thermal expansion coefficient, 1/°C
␸ 20 40 30 4.3 Friction angle of shale

Thus it is clear that although the maximum NPV obtained for the shallow reservoir was $1.35 million
– that would mean caprock failure. The maximum NPV that could be achieved with a safe SAGD
operation was $1.10 million. Similarly, maximum NPV that could be achieved in the deep reservoir was
$1.16 million, but a safe SAGD operation could only give a maximum NPV of $1.10 million. All the
parameters affected the NPV of the model. So to identify their effects, plots like Figure 15 were made for
both shallow and deep reservoir. Figure 15 shows the variation of NPV with changing injector well
pressure for different well spacings. It is clear that 80 m spacing is a better solution, and a higher BHP
increases the revenue. But the maximum BHP that should be used for injection is limited to 2800 kPa as
the caprock started to fail beyond that. All the spacing and well separation combinations were checked and
the maximum operating pressure (MOP) was found for both the cases and are shown in Table 7.
SPE-170130-MS 15

Figure 16 —Cumulative probability and probability density functions of surface heave under assumed geomechanical uncertainty for shallow
reservoir

Uncertainty Analysis
Effect of uncertainty in geomechanical properties
After optimization, the best operating conditions were identified for a fixed geology. The next step was
to quantify the uncertainty associated with caprock integrity because of geological uncertainty. Thus,
optimized models were run with fixed operating conditions but varying geomechanical properties of the
reservoir sand and shale. The various parameters of constitutive models for reservoir sand and shale were
assumed to follow normal distributions, their properties were altered within reasonable ranges and their
effect on the surface heave was observed. These ranges are given in Table 8.
The results of these simulations suggest that with the given operating constraints, the shallow reservoir
was not likely to undergo caprock failure based on the property ranges above. Figure 16 shows the
cumulative probability and probability density functions of surface heave for the shallow reservoir under
the above mentioned uncertainty in geomechanical parameters. These results have been obtained using the
response surface method described earlier along with Monte-Carlo simulation. It can be seen that the
surface heave was likely to be in the range of 0.27– 0.65 m with a P50 value of 0.40 m (R-squared value
for response surface is 0.99). The caprock did not undergo failure in any of the models simulated for
uncertainty in shallow reservoir for the simulation time of 3652 days. Similar results were obtained for
the deep reservoir: surface heave was in the range of 0.45– 0.98 m with a P50 value of 0.68 m (R-squared
value of 0.99).
However, one interesting observation was made for the deep reservoir. Under the optimum operating
conditions with a safe MOP, when geomechanical uncertainty was incorporated, the deep reservoir
showed caprock failure in some cases. The P50 value for the time taken for the deep reservoir to fail under
geomechanical uncertainty was 3175 days. This correlates well with the fact that the surface heave was
also found to be generally higher for the deep reservoir (P50 value of 0.68 m in deep versus P50 value
of 0.40 m in shallow). The reason for this is the higher injection pressures used in the deep reservoir, along
with correspondingly higher injection temperatures (same as steam saturation temperatures). The addi-
16 SPE-170130-MS

Figure 17—Cumulative probability and probability density functions of time taken by caprock to fail under assumed geomechanical uncertainty for
deep reservoir

Figure 18 —Poorer quality geobody distributions to mimic shaley-sand – 0% poor geobody in sand (left), 10% poor geobody in sand (center) and 20%
poor geobody in sand (right)

tional stress change caused by these parameters is not able to dissipate completely into the thicker
overburden in case of deep reservoir, thus giving a higher deformation at shale top as well as at surface,
and in some cases causing a caprock failure. It must be noted that the P50 value for failure time was fairly
high. So it is quite likely that caprock failure will happen in the deep reservoir but it will take a fairly long
time to initiate. The distribution for failure time for the deep reservoir has been shown in Figure 17.
Uncertainty because of reservoir heterogeneity
A quick analysis was done to see the effect of the presence of poorer quality rock in the main reservoir
sand. An extensive geostatistical study was not done for this purpose; rather, a poorer quality rock was
populated using basic geostatistics inside the mudstone and sand. This poorer quality rock was assumed
to have 10 times lesser permeability than the reservoir and half the porosity of the reservoir. This was done
as a quick check to see if the caprock deformation is delayed because of reservoir heterogeneity and if it
could mean a higher MOP while ensuring safe operation. Three geobody distributions were considered
and are shown in Figure 18.
The models were run at injection pressures of 3200 kPa (shallow reservoir) and 5400 kPa (deep
reservoir) with saturated steam temperature. These injection pressures were well in excess of the safe
SPE-170130-MS 17

Figure 19 —Comparison of caprock state under geological uncertainty in shallow reservoir – the reservoir with 0% of poor geobodies undergoes
caprock failure at about 1800 days

Figure 20 —Comparison of caprock state under geological uncertainty in deep reservoir- the reservoir with 0% of poor geobodies undergoes caprock
failure at about 2300 days

MOPs determined during the optimization study (Table 6). The results are as expected. The base case
(with best quality sand, 0% poor geobody) had caprock failure at about 1800 days in shallow reservoir and
at about 2200 days in deep reservoir as shown in Figure 19 and Figure 20. The reason is that the injection
pressure and temperature are high enough to cause the failure of the rock. However, when there was a
presence of poor quality geobodies inside the sand, it impeded the growth of the steam chamber – and that
18 SPE-170130-MS

lead to a poorer recovery and poorer cSOR. But it had also lead to a lesser deformation of the caprock
and the caprock remained intact for the full simulation duration even under high injection pressures. Thus,
it will be safe to say that for reservoirs with poor quality rock/mudstone distribution, the value of MOP
will be higher.

Conclusions
1. It is very important to assess caprock integrity while performing reservoir studies and simulations
for SAGD. A workflow has been presented in this paper which underlines the steps for achieving
this by doing 2-way coupled geomechanics thermal simulations. The first step was to do a
sensitivity study to identify the relative impact of various operating parameters like injection
pressure, lateral well-pair spacing and injector-producer separation on a variety of objective
functions such as field NPV, cSOR, heave on surface and time taken for caprock to undergo
failure. This was followed by an optimization study which was aimed at maximizing the NPV,
while keeping the caprock intact. Lastly, the impact of uncertainty in geomechanical and reservoir
parameters on caprock integrity was studied using varying geomechanical properties and varying
poor quality sand distributions in the reservoir.
2. The injection BHP was found to be the most influential parameter on all the objective functions
considered in the sensitivity study. A high injection BHP generally meant a high NPV from the
model, but it also meant that there is a greater chance of caprock failure. Keeping this in mind,
an MOP was found for both reservoirs-shallow and deep.
3. Based on the uncertainty analysis, it was found that under an optimized operating scenario, the
deep reservoir can still undergo caprock failure for some configurations of geomechanical
properties. However, under the considered uncertainty in geomechanical parameters, caprock
integrity was always maintained in the shallow reservoir. This correlated to the larger surface
heaves obtained in the deep model. The reason for this difference was the much higher injection
pressure used in the deep reservoir along with the high injection steam temperature.
4. Poorer quality shale/mudstone geobodies distributed in a good quality reservoir impedes the
growth of the steam chamber and the associated stress change. It was found that the poor quality
reservoir with lenses of geobodies underwent lesser caprock deformation and lesser chance of
caprock failure. Although this also meant that the economics from the poorer quality rock was
worse.

Acknowledgements
The authors wish to thank CMG for supporting the work done in this paper and for providing the
permission to publish it.

Nomenclature
Ei ⫽ Initial tangential modulus
␴r, ⫽ Reference stress
Patm ⫽ Atmospheric pressure
T ⫽ Current temperature
To ⫽ Initial temperature
c ⫽ Cohesion

max ⫽ Maximum value of friction angle
⫽ Maximum principle effective stress
⫽ Minimum principle effective stress
Et ⫽ Tangential modulus
SPE-170130-MS 19

Ke ⫽ Parameter of the tangential modulus equation


ne ⫽ Parameter of the tangential modulus equation
nte ⫽ Parameter of the tangential modulus equation
m ⫽ Parameter of the tangential modulus equation
␸ ⫽ Friction angle
␸o ⫽ Friction angle at confining pressure of 1 atm
⌬␸ ⫽ Friction angle change per log cycle of minimum principle effective stress
Bm ⫽ Bulk modulus
Ko ⫽ Parameter of the bulk modulus equation
nb ⫽ Parameter of the tangential modulus equation
ntb ⫽ Parameter of the tangential modulus equation
K ⫽ Current permeability
Ko ⫽ Initial permeability
␧v ⫽ Volumetric strain
␶ ⫽ Shear stress at failure
␴’ ⫽ Normal effective stress
␣ ⫽ Thermal expansion coefficient
y ⫽ Objective function
ao ⫽ Coefficient in objective function equation
aj ⫽ Linear term coefficient in objective function equation
ajj ⫽ Quadratic term coefficient in objective function equation
aij ⫽ Interaction term coefficient in objective function equation
MOP ⫽ Maximum Operating Pressure
BHP ⫽ Bottom Hole Pressure

References
Beattie, C.I., Boberg, T.C., McNab, G.S. 1991. Reservoir Simulation of Cyclic Steam Stimulation in
the Cold Lake Oil Sands. SPE Res Eval & Eng 6 (2): 200 –206
Chin, L.Y., Silpngarmlert, S., and Schoderbeck, D.A. 2011. Heave Prediction by Coupled Modeling
of Geomechanics and Reservoir Simulation for Methane Hydrate Reservoirs. Paper ARMA 11–307
presented at the 45th US Rock Mechanics / Geomechanics Symposium 2012, San Francisco, CA, USA,
26 –29 June.
Chin, L.Y., Tomberlin, T.A., Ramos, G.G., and Chalaturnyk, R.J. 2012. Evaluation of Caprock
Stability by Coupled Modeling of Geomechanics and Reservoir Simulation under Steam Injection for
Producing Oil Sands Reservoirs. Paper ARMA 12–139 presented at the 46th US Rock Mechanics /
Geomechanics Symposium 2012, Chicago, IL, USA, 24 –27 June.
CMOST User’s Guide, Version 2013.12. 2014. Calgary, Alberta: Computer Modelling Group Ltd.
Duncan, J.M., and Chang, C. 1970. Nonlinear Analysis of Stress and Strain in Soils. Journal of Soil
Mechanics and Foundation Engineering, ASCE 96 (5), pp. 1629 –1653.
Energy Resources Conservation Board (ERCB), now AER, 2010. Total E&P Canada Ltd. Surface
Steam Release of May 18, 2006 Joslyn Creek SAGD Thermal Operation. ERCB Staff Review and
Analysis. Document submitted February 11, 2010.
Gu, F., Chan, M., and Fryk, R. 2009. Geomechanical Data Acquisition, Monitoring and Applications
in SAGD. Paper 2009 –177 presented at the Canadian International Petroleum Conference 2009, Calgary,
16 –18 June.
Grizzly Oil Sands. 2010. Project: Safe Operating Pressures for SAGD at Algar Lake. Report prepared
by Taurus Reservoir Solutions. http://www.grizzlyoilsands.com/upload/mediaelement/95/01/13app 5.pdf
(Downloaded 5 March 2014)
20 SPE-170130-MS

Khan, S., Han, H., Ansari, S., Vishteh, M. and Khosravi, N. 2011. Caprock Integrity Analysis in
Thermal Operations: An Integrated Geomechanics Approach. Paper WHOC11– 609 presented at the
World Heavy Oil Congress, Edmonton, Alberta, Canada, 14 –17 March.
Settari, A., Ito, Y., Fukushima, N. and Vaziri, H. 1993. Geotechnical Aspects of Recovery Processes
in Oil Sands, Can. Geotech. J., Vol. 30, pp. 22–33.
STARS User’s Guide, Version 2013.10. 2013. Calgary, Alberta: Computer Modelling Group Ltd.
Touhidi-Baghini A. 1998. Absolute Permeability of McMurray Formation Oil Sands at Low Confining
Stresses, PhD Thesis, University of Alberta.
Tran, D., Shrivastava, V., Nghiem, L., and Kohse, B. 2009. Geomechanical Risk Mitigation for CO2
Sequestration in Saline Aquifers. Paper SPE 125167-PP presented at the 2009 SPE Annual Technical
Conference and Exhibition, New Orleans, LA, 4 –7 October.
Uwiera-Gartner, M.M.E., Carlson, M.R., Walters, D., and Palmgren, C.T.S. 2011. Geomechanical
Simulation of Caprock Performance for A Proposed, Low Pressure, Steam-Assisted Gravity Drainage
Pilot Project. Paper CSUG/SPE 148886 presented at the 2011 Canadian Unconventional Resources
Conference, Calgary, Alberta, Canada, 15–17 November.
Walters, D., Wang, J., and Settari, A. 2012. A Geomechanical Methodology for Determining
Maximum Operating Pressure in SAGD Reservoirs. Paper SPE 157855 presented at the SPE Heavy Oil
Conference Canada, Calgary, Alberta, Canada, 12–14 June.
Yang, B., Xu, B., and Yuan, Y. 2013. Impact of Thermal Pore Pressure on the Caprock Integrity
during the SAGD Operation. Paper SPE 165448 presented at the SPE Heavy Oil Conference Canada,
Calgary, Alberta, Canada, 11–13 June.
Yuan, Y., Xu, B. and Palmgren, C. 2011. Design of Caprock Integrity in Thermal Stimulation of
Shallow Oil-Sands. Paper CSUG/SPE 149371 presented at the 2011 Canadian Unconventional Resources
Conference, Calgary, Alberta, Canada, 15–17 November.

You might also like