Anda di halaman 1dari 19

CANADIAN HEAVY

OIL ASSOCIATION

SPE/PS-CIM/CHOA 97742
PS2005-332
Steam-Injection Strategy and Energetics of Steam-Assisted Gravity Drainage
I.D. Gates, U. of Calgary; J. Kenny and I.L. Hernandez-Hdez, Atech Application Technology Ltd.; and
G.L. Bunio, Paramount Resources Ltd.

Copyright 2005, SPE/PS-CIM/CHOA International Thermal Operations and Heavy Oil Symposium
This paper was prepared for presentation at the 2005 SPE International Thermal Operations
and Heavy Oil Symposium held in Calgary, Alberta, Canada, 13 November 2005.
This paper was selected for presentation by an SPE/PS-CIM/CHOA Program Committee
following review of information contained in a proposal submitted by the author(s). Contents of
the paper, as presented, have not been reviewed by the Society of Petroleum Engineers,
Petroleum SocietyCanadian Institute of Mining, Metallurgy & Petroleum, or the Canadian
Heavy Oil Association and are subject to correction by the author(s). The material, as
presented, does not necessarily reflect any position of the SPE/PS-CIM/CHOA, its officers, or
members. Papers presented at SPE and PS-CIM/CHOA meetings are subject to publication
review by Editorial Committees of the SPE and PS-CIM/CHOA. Electronic reproduction,
distribution, or storage of any part of this paper for commercial purposes without the written
consent of the SPE or PS-CIM/CHOA is prohibited. Permission to reproduce in print is
restricted to a proposal of not more than 300 words; illustrations may not be copied. The
proposal must contain conspicuous acknowledgment of where and by whom the paper was
presented. Write Librarian, SPE, P.O. Box 833836, Richardson, TX 75083-3836, U.S.A., fax
01-972-952-9435.

Abstract
Steam-Assisted Gravity Drainage (SAGD) is being operated
by several operators in Athabasca and Cold Lake reservoirs in
Central and Northern Alberta. In this process, steam, injected
into a horizontal well, flows outwards, contacts and loses its
latent heat to bitumen at the edge of a depletion chamber. As
a consequence, the viscosity of the bitumen falls, its mobility
rises, and it flows under the action of gravity towards a
horizontal production well located several meters below and
parallel to the injection well. In practice, the temperature
difference between the injected steam and produced fluids,
called the subcool, is maintained between 15 and 30C.
Despite many pilots and commercial operations, it remains
unclear what the impact of subcool on the performance and
thermal efficiency of SAGD especially in reservoirs with a top
gas zone. The objective of this study was to define a steam
chamber operating strategy that leads to optimum oil recovery
for minimum cumulative steam to oil ratio in a reservoir with
a top gas zone. These findings were established from
extensive simulation runs that were built from a detailed
geostatistically generated static reservoir model. The strategy
devised uses a high initial chamber injection rate and pressure
prior to chamber contact with the top gas. Subsequent to
breakthrough of the chamber into the gas cap zone, the
chamber injection rates are lowered to balance pressures with
the top gas and avoid or at least minimize convective heat
losses of steam to the top gas zone. The results are also
analyzed by examining the energetics of SAGD.
Introduction

A cross-section of the Steam-Assisted Gravity Drainage


(SAGD) process is displayed in Figure 1. Steam is injected
into the formation through a horizontal well. In Figure 1, the
wells are portrayed as points that extend into the page.
Around and above the injection well, a steam depletion
chamber grows. At the edge of the chamber, heated bitumen
and (steam) condensate flow under the action of gravity to a
production well typically placed between 5 and 10 m below
and parallel to the injection well. Usually, the production well
is located several meters above the base of pay. In industrial
practice1,2, injection and production wells lengths are typically
between 500 and 1000 m. The injection pressure, because the
steam chamber operates at saturation conditions, controls the
operating temperature of SAGD.

Steam Chamber

Native
bitumen

Reservoir
Thickness
Injection
Well
Bitumen
flow zone

Production
Well

Figure 1: Cross-section of the Steam-Assisted Gravity


Drainage (SAGD) process.

SAGD has been extensively piloted in Athabasca and Cold


Lake reservoirs in Alberta2-13 and is being used as a
commercial technology to recover bitumen in several
Athabasca reservoirs11.
These pilots and commercial
operations have demonstrated that SAGD is technically
effective but it has not been fully established whether its
operating conditions are at optimum values. This is especially
the case in reservoirs in contact with gas or water zones where
the optimum operating strategy remains unclear.
The

variability of the cumulative injected steam (expressed in Cold


Water Equivalents, CWE) to produced oil ratio (cSOR) shows
that some SAGD wellpairs operate fairly efficiently (with
cSOR between 2 and 3) whereas others operate at much
greater cSOR up to 10 and higher13. Higher cSOR means that
more steam is being used per unit volume bitumen produced.
The higher the steam usage, the greater the amount of natural
gas combusted, and the less economic is the process.
One key control variable in SAGD is the temperature
difference between the injected steam and the produced fluids.
This value, known as the subcool, is typically maintained in a
form of steam-trap control between 15 and 30C. The subcool
is being used as a surrogate variable instead of the height of
liquid above the production well. The liquid pool above the
production well prevents flow of injected steam directly from
the injection well to the production well thus promoting
injected steam to the outer regions of the depletion chamber
and delivery of its latent heat to the bitumen. It remains
unclear what is the value of the optimum steam-trap subcool
temperature difference and how the operating pressure impacts
the optimum subcool value. It also remains unclear how the
subcool should be controlled in heterogeneous reservoirs that
have top gas.
In this study, a sequence of SAGD simulations were
conducted on a reservoir with a detailed heterogeneous
geological description obtained from geostatistical analysis of
log and core data. After over 100 simulations, an optimized
steam injection strategy was devised that produced a
reasonable cSOR for design of a SAGD operation in
McMurray reservoir with a top gas zone. In addition, the
energetics of SAGD are examined by evaluating the flowing
steam quality in the steam chamber.
Geological Model
To allow the assessment of geological variation and
uncertainty on the SAGD process a detailed statistically based
static geological description was prepared. The overall
domain of the parent geological model covers 32 sections in
the vicinity of Paramount Resources oilsands lease at
Surmount described by Robinson et al. (2005). Petrophysical
interpretations and picks for seven horizons were available at
33 wells within the parent geological domain and were
mapped from the well inputs by using a Local B-Spline
algorithm in Roxar's IRAP RMS geostatistical package12. Of
the 33 wells, facies logs from fourteen of them were prepared
to populate the geological model14.
A stacked rectangular grid was used to build the geostatistical
model. In the East to West direction, 136 columns, each 50 m
wide, were used giving a total width equal to 6 800 m. In the
North to South direction, 259 rows, each 50 m wide, were
used to give a total length of 12 950 m. In the vertical
direction, 95 layers of variable thickness, most of them under
1 m thickness, were used. The total cell count for the
geological grid was 3 346 280.

SPE/PS-CIM/CHOA 97742

The facies logs distinguished eight rock types that were


identified as follows:
Facies 1:
Facies 2:
Facies 3:
Facies 4:
Facies 5:
Facies 6:
Facies 7:
Facies 8:

Shoreface Sand,
Muddy Marine Sand,
Mudstone,
Mud Dominated Heterolithic Strata,
Sand Dominated Heterolithic Strata,
Sandstone,
Breccia, and
Mudstone Filled.

Petrophysical interpretations provided logs of Sg, So, Sw, and


kh for the 33 wells within the geological model parent domain.
From a detailed review and synthesis of core, core analysis,
logs and a history match of a McMurray SAGD pilot,
correlations of permeability versus porosity and vertical to
horizontal permeability ratio (kv/kh) were established for each
facies in the model. The correlations are listed in Table 1.
The facies specific permeability-porosity transforms were
used to generate well permeability logs.

Table 1:
Horizontal permeability versus porosity
correlations.

0.050
0.075
0.100
0.125
0.150
0.175
0.200
0.225
0.250
0.275
0.300
0.325
0.350
0.375
0.400

Facies
1
4.2
18.7
53.7
121.8
237.8
418.7
683.3
1052.7
1549.4
2198.0
3024.5
4056.9
5324.3
6857.9
8690.0

2
0.0021
0.0057
0.0118
0.0206
0.0324
0.0477
0.0665
0.0893
0.1163
0.1475
0.1834
0.2240
0.2696
0.3203
0.3764

3
0.0002
0.0005
0.0009
0.0013
0.0019
0.0026
0.0033
0.0042
0.0051
0.0062
0.0073
0.0085
0.0098
0.0112
0.0127

4
5.1
24.0
71.9
168.3
337.4
607.6
1011.1
1584.7
2368.6
3407.2
4748.4
6443.9
8549.2
11123
14228

5
17.3
61.0
149.2
298.6
526.5
850.4
1288.3
1858.4
2579.1
3469.1
4547.3
5832.9
7345.1
9103.2
11127

6
17.5
63.0
156.5
317.1
564.7
919.7
1403.3
2037.1
2843.3
3844.2
5062.8
6522.2
8246.0
10258
12582

7
10.8
36.2
85.5
166.5
286.9
454.6
677.2
962.5
1318.2
1752.0
2271.5
2884.5
3598.7
4421.6
5360.9

8
0.00023
0.00050
0.00087
0.00134
0.00191
0.00257
0.00333
0.00418
0.00512
0.00616
0.00729
0.00850
0.00981
0.01121
0.01269

For each facies, the vertical to horizontal permeability ratio for


the model was set to 0.65, .055, .055, 0.055, 0.300, 0.500,
0.500 and 0.005 for facies 1, 2, 3, 4, 5, 6, 7, and 8
respectively.
As described in Robinson et al. (2005), the facies and
petrophysical well logs were entered into Roxar's IRAP RMS
geostatistical modelling package and upscaled to populate the
gridblocks intersected by the wells with the following suite of
parameters: facies, Sg, So, Sw, , kh and kv. The grid was kept
relatively fine in the vertical direction and the distribution of
layer thickness ensured that the blocked well data represented
the actual log data sufficiently well.
To populate the facies at gridblocks away from the wellbores,
a Sequential Indicator Simulation15, conditioned on the facies
vertical proportion curve, was done.
Any number of
equiprobable facies realizations may be generated from this

SPE/PS-CIM/CHOA 97742

data set. The facies distribution through the reservoir for one
realization is displayed in Figure 2. After review by team
geologists, the geological model was considered to be a
reasonable reflection of the geological environment14.

Table 3:
Horizontal and vertical permeabilities:
minimum, maximum, average, and standard deviations for
each facies.
Horizontal, mD
Facies
1
2
3
4
5
6
7
8

Vertical, mD

Min Max Avg


Min
Max Avg

200 4000 1903 1000


130
2600 736
951
0.001
1
0.0678 0.1125 5(10-5) 0.055 0.0031 0.0034
0
0.005 0.0011 0.0013
0
5(10-4) 0.0001 0.0001
500 5000 2721 1239
27.5
275 122.39 64.36
750 9000 4280 2107
225
2700 1270 684
1500 10000 5563 2295
750
5000 2777 1178
750 5000 2729 1176
375
2500 1382 588
0
0.005 0.001 0.0015
0
0.005 0.0001 0.0005

As described in Robinson et al. (2005), 100 equiprobable


facies, porosity, permeability and saturation realizations were
constructed. A comparison revealed that there were small
variations in volumes among the 100 realizations and a
realization near the center of the population was chosen to
construct the working geological model from which a
reservoir simulation model could be extracted.

FACIES

Figure 2: Facies distribution in geological model.

The gas and oil saturation distributions within the model are
shown in Figure 3. The gas saturation is concentrated in the
upper marine sands whereas the highest oil saturations are
located in the central elevations of the model. The porosity
and horizontal and vertical permeability distributions are
displayed and described in Robinson et al. (2005).

To populate the permeabilities, porosities and saturation


distributions in the regions between the well locations,
Sequential Gaussian Simulation (SGS) was used. The SGS
that was used to populate gas, oil and water saturations in the
model were run independently to ensure that the underlying
statistics describing these parameters were honored in the
model. The saturations were then normalized to ensure that
Sg+So+Sw = 1 at every grid block. To obtain porosity and
permeability distributions between the well locations, a SGS
was conducted with upper and lower bound cut-offs applied to
the wellblock porosity and permeability distributions. Tables
2 and 3 list the minimum and maximum cut-offs, average
values, and standard deviation of the porosity and
permeabilities distributions.

Table 2: Porosity: minimum, maximum, average, and


standard deviations for each facies.
Facies
1
2
3
4
5
6
7
8

Min
0.1
0.05
0
0.05
0.1
0.2
0.1
0

Max
0.32
0.25
0.15
0.3
0.4
0.4
0.36
0.15

Avg
0.2363
0.1434
0.0549
0.2140
0.2670
0.3081
0.2764
0.0366

Gas
Saturation

Oil
Saturation

0.1102
0.0567
0.0508
0.0728
0.0657
0.0454
0.0652
0.0476

Figure 3: Gas and oil saturation distributions in geological


model.

SPE/PS-CIM/CHOA 97742

Reservoir Simulation Model

East

The next step of the workflow was to upscale, extract, and


import the geological model into the reservoir simulator.
The reservoir simulation of the SAGD process was conducted
with Computer Modelling Groups (CMG) thermal reservoir
simulator STARS16. As described in Robinson et al. (2005),
an upscaled geological description of Section 30 was extracted
consisting of 516 096 cells from the original geological parent
model.
From this upscaled model, within the CMG
preprocessor, a subdomain able to accommodate two 750 m
wellpairs, with 500 m in the East-West (X=lateral) direction
by 950 m in the North-South (Y=downwell) direction, was
extracted. In the downwell direction, the subdomain was
tessellated into 12 gridblocks. The total number of gridblocks
equals 74 592. Figure 4 displays grid of the reservoir
simulation model two-thirds of the way down the wells; the
left wellpair is referred to by LP whereas the right wellpair is
referred to as RP.

North

Porosity

West

LP
RP
South

Horizontal
Permeability, mD

West

East

LP

RP

Figure 4: Dual wellpair reservoir simulation model grid.

Figure 4 shows the locations of the left and right wellpairs as


well as wells that were inserted into the gas cap to mimic the
continuity of the gas cap beyond the model. The gas cap wells
were constrained to constant bottomhole pressure equal to the
gas cap pressure (they will be referred to as pressuremaintenance wells). The grid blocks in the East-West
direction were refined to ensure that the maximum gridblock
width 50 m on each side of the left and right wellpairs was
3.845 m. The inter-wellpair spacing is 200 m.
Figure 5 displays porosity and horizontal permeability
distributions of the reservoir simulation model at planes along
the wellpairs and two lateral cross-sections. The black lines in
the downwell planes indicate the locations of the wellpairs.
In the central region of the model, the porosity ranges from
0.27 to 0.38 and it is improving from the South to the North.
Similarly, the horizontal permeability mainly lies between 2
and 6 D in the central regions of the reservoir with better
permeability in the Northern part of the reservoir.
Figure 6 shows cross-sections of the gas, oil, and water
saturations in the reservoir simulation model. Consistent with
the geological model, the gas cap, typically 3 to 4 m thick, is
located throughout the model.

Figure 5: Dual wellpair reservoir simulation model:


porosity and horizontal permeability distributions (crosssections along wellpairs and two lateral locations
downwell).

In the top gas zone, the maximum gas saturation is about 0.82
with the remainder of the pore space containing water.
Initially, the pressure of the gas cap is about 1050 kPa (the
pressure equaled 936 and 1631 kPa at the top and bottom of
the reservoir model, respectively). The pressure-maintenance
wells located in the gas cap are set to produce fluids if the
pressure exceeds 1075 kPa in order to mimic an gas cap zone
that extends beyond the model domain. It is anticipated that in
the case with an established depletion chamber in contact with
the top gas zone, if the steam injection pressure is too large,
steam will be diverted from the steam chamber into the top gas
zone raising the pressure there. Then, steam and gas will flow
out into the gas cap and out through the gas cap pressuremaintenance wells.
The oil saturation distribution displayed in Figure 6 reveals
that there is a central region of the reservoir with relatively
high oil saturation. The average thickness of the oil-rich zone
(> 0.7 oil saturation) is about 20 m. In some parts of the
reservoir, it is as high as 26 m and in others it drops as low as
14 m. The production wells of each wellpair are located just a
couple meters above the bottom of the oil-rich zones. The
injection wells are located 5 m above the production wells.

SPE/PS-CIM/CHOA 97742

East

North

Gas Saturation

West

LP
RP
South

Table 4 summarizes additional heat loss parameters, fluid


properties, and rock-fluid properties. The reservoir simulation
model fluid components consisted of bitumen, water (liquid
and vapour), and solution gas. The bubble point pressure in
the model was taken to be 1000 kPa. Since most of the model
is above this pressure, the initial solution gas to oil ratio was
constant at 4.2 m3/m3 throughout the oil layer. The K-value
relationship used in the model is listed in Table 4. The
bitumen viscosity was similar to Mehrotra and Svrceks
(1986) correlation for Athabasca bitumen17. The oil-water and
gas-liquid relative permeability curves were obtained from a
detailed history match of a McMurray SAGD pilot. The
endpoints are listed in Table 4.
Table 4: Reservoir simulation input parameters.

Oil Saturation

Water Saturation

Item
Initial Reservoir
Temperature, C
Top of model depth, m
Sorw
Swc
Sorg
Sgc
krwro
krocw
krogc
krg(Sorg)
Oilsand thermal
conductivity, kJ/m day C
Over/Underburden heat
capacity, kJ/m3 C
Over/Underburden thermal
conductivity, kJ/m day C
Methane K-value
correlation,
kv4

K-value = kv1 e T + k

v5

Value
10
277
0.2
0.3
0.15
0.05
0.197
0.48
0.8
1
149.5
1169
74.9
kv1= 3.1914x104 kPa
kv4=-330.67 C
kv5 = -277.1 C

Well Constraints

Figure 6: Dual wellpair reservoir simulation model: gas,


oil, and water saturations (cross-sections along wellpairs
and two lateral locations downwell).

The water saturation distribution shown in Figure 6 reveals


that there are relatively high water saturation regions at the top
and at the bottom of the model. The region at the top is above
the gas cap and consists of tight muddy marine sand reservoir
rock (Facies 2). The porosity and permeability of this facies
are both low (see Tables 2 and 3). The high water saturation
at the bottom of the reservoir is in Facies 6 and the water is
relatively mobile given the permeability of this part of the
reservoir.

At the injection wells, the steam injection pressure is


constrained to a maximum bottomhole pressure. The steam
quality at sandface equaled 0.8. At the production wells, the
steamtrap constraint was used with 5C temperature setting.
The CMG steamtrap control algorithm does not impose the
temperature difference between the injection and production
well16. Rather, a 5C subcool in CMGs algorithm means that
the bottomhole pressure in the wellblock is set corresponding
to the pressure 5C above the saturation temperature of the
gridblock. This means that no live steam can be produced
from the well. The subcool temperature difference often
referred to from field data as the difference between the steam
injection temperature and the produced fluids temperature can
be calculated from the results of the simulation.

Model Initialization
To model steam circulation, line heaters were positioned in the
locations of the injection and production wells of each
wellpair. The heating rate corresponded to the heat delivered
by 250 m3/day CWE of 0.8 quality steam. In the location of
the production wells, the wells were opened with a maximum
bottomhole pressure equal to the initial reservoir pressure.
The reason for this is to relieve pressure buildup due to
thermal expansion of the fluids near the wellbore. Similarly,
temporary production wells were inserted into the same
locations as the steam injectors so that pressure was relieved
along the injectors as the reservoir fluids near the wellbore
heated up. The circulation period lasted three months. When
the wellpairs were changed to SAGD mode, the line heaters
were turned off, the temporary production wells positioned in
the steam injection well locations were removed, and steam
injection commenced at the target rate or pressure.
Results: Optimization of Performance: From Dual
Wellpair to Single Wellpair Models

SPE/PS-CIM/CHOA 97742

injection rate and pressure and production rates, cumulative


volumes produced, and cSOR for the LP wellpair,
respectively. Figures 9 and 10 show the same plots for the RP
wellpair.

Figure 7: Steam injection strategy in the LP wellpair


operated at constant 2000 kPa injection pressure.

The simulation runs of the dual-wellpair model took up to 20


hours to complete a 12 year forecast with a 2.45 GHz dualprocessor computer workstation with parallel-enabled STARS.
It was recognized early in the study that to carry out a large
number of simulations would be prohibitive and would not be
possible on the available 5 workstations.
To obtain simulation run times in reasonable runtimes without
compromising the geological description and physics of the
SAGD process, two single wellpair models, identified as RP
and LP, were created. Each single wellpair model consists of
half the dual-wellpair model. Each half has a 250 m by 950 m
areal footprint. There are 37 296 blocks in each of the single
wellpair models and 12 year forecast runs lasted under 9 hours
on a 2.45 GHz dual processor workstation with parallelenabled STARS. It was decided that first, the operating
strategy of each single wellpair model would be optimized and
second, the individually optimized operating strategies would
be applied in the dual wellpair model. It was recognized that
due to inter-wellpair communication, the results of the single
wellpair operating strategies would have to be adjusted once
introduced into the dual wellpair model. However, it was
expected that given the presence of the gas cap, the operating
strategy would have to be gentle on the reservoir (otherwise
excessive steam would be lost to the gas cap), and therefore
communication issues would not be too hard to resolve. The
same circulation preheat strategy as was described for the dual
wellpair model was used for the single wellpair models. Many
cases were run or partially run to improve the cSOR after six
and twelve years of SAGD operation. These runs were in
sequence and in parallel and the operating strategy was
modified after carefully reviewing and analyzing the results of
each run to further improve the cSOR.

Figure 8: Production rates, cumulative volumes, and


cSOR of LP wellpair operated at constant 2000 kPa
injection pressure.

Constant Pressure Injection


First, the results for constant steam injection pressure at 2000
kPa will be presented. Figures 7 and 8 show plots of the

Figure 9: Steam injection strategy in the RP wellpair


operated at constant 2000 kPa injection pressure.

SPE/PS-CIM/CHOA 97742

Without steam placement control, injected steam takes the


path of least resistance to better quality reservoir at the toe and
bypasses reservoir of poorer quality at the heel.
By
introducing control of steam placement, steam is concentrated
at the heel section at higher pressures for a longer time and so
the steam chamber grows more uniformly at the heel. Also,
injection rates and pressures at the toe section are lowered to
delay the onset of the chamber contacting the top gas zone.
Robinson et al. (2005) discusses steam placement control in
more detail.

Figure 10: Production rates, cumulative volumes, and


cSOR of RP wellpair operated at constant 2000 kPa
injection pressure.

The results in Figures 8 and 10 reveal high cSOR profiles that


are the result of excessive steam losses from the steam
chamber to the gas cap. Because steam rates are high, very
little latent heat is being delivered to the bitumen and so
bitumen rates are low.
Optimized Steam Injection
With manual optimization, improved operating strategies were
determined for both LP and RP wellpairs. The strategies for
both wellpairs are similar and comprise a high steam injection
pressure until each steam chamber establishes contact with the
top gas. High injection pressure implies a relatively high
saturation temperature that leads to favourable bitumen
viscosities in the early stages of SAGD. After the top gas is
encountered, the top gas pressure dictates the steam chamber
operating pressure which according to the optimized strategy
is maintained constant thereafter at or just below the gas cap
pressure. This ensures convective losses of steam after
breakthrough to the top gas are avoided or at least minimized.
This operating strategy is consistent with the results of Gates
and Chakrabarty13.
In uniform steam injection pressure simulations, the Northern
end of the LP reservoir was depleted more rapidly than the
Southern part of the reservoir. This is because the geology
along the LP wellpair has a large contrast in reservoir quality
going from toe to heel. The toe section of the injection well
has better vertical permeability than the heel section and
consequently breakthrough to the top gas zone happens
quickly and so makes favorable thermal management of the
steam chamber more problematic. On the other hand the
reservoir quality along the well pair of the RP model is more
uniform and so makes the breakthrough time more uniform
along the injection well, which makes favorable thermal
management prior to and subsequent to breakthrough to the
gas cap easier. In order to promote more uniform formation of
the depletion chamber in the reservoir, steam injection into the
toe of the LP and RP wells were periodically stopped.

Strategies to control steam placement might include a


combination of packers and chokes along the horizontal
section and/or the use of limited entry perforations as
described by Boone et al. (1999) as well as any inflow control
devices18. Flexibility will be key in the design of the injection
string because the geology along the injector will dictate the
optimal steam placement requirements.
Pressure and
temperature monitoring along the steam chamber will be
essential to effective management.
Figure 11 displays the injection well constraints of the
optimized operating strategy over the twelve year forecast
period for the LP wellpair. The plot reveals that over the first
two years the injection constraint was a maximum bottom hole
pressure limitation and beyond that, the constraint was the
steam injection rate that was lowered in steps until the end of
the forecast. The production rates and associated cumulative
volumes of the gas, oil and water, and cSOR over the twelve
year forecast period for the LP wellpair are presented in
Figures 12.

Figure 11:
wellpair.

Steam injection strategy in optimized LP

In the first year of the process, the steam injection pressure is


relatively high at 1800 kPa. The injection pressure is then
lowered to 1100 kPa and thereafter, in the rate-constrained
period, the pressure remains near 1000 kPa. The maximum
steam injection rate is short-lived at 175 CWE m3/day. Over
the majority of the process, the steam injection rate ranges
from 100 to 150 CWE m3/day.

Figure 12 shows that the cSOR settles after an initial transient


period after the onset of SAGD to about 2.6 m3/m3. The shape
of the bitumen production rate profile is similar to SAGD field
data from Athabasca reservoirs. The field rate hovers just
above 50 m3/day for most of the life of the process and after 8
years of SAGD operation begins to decline. With the
exception of the first year, throughout the process, more water
is produced than injected. This indicates that water from the
formation is being produced.

SPE/PS-CIM/CHOA 97742

0.00067 m3/day/m in the RP wellpair at six and twelve years,


respectively.

Figure 13:
wellpair.

Steam injection strategy in optimized RP

Figure 12: Production rates, cumulative volumes, and


cSOR of optimized strategy in LP wellpair.

Figure 13 displays the injection well constraints of the


optimized operating strategy for the RP wellpair. The
production rates and cumulative volumes of oil and water, and
cSOR over the twelve year forecast period for the RP wellpair
are presented in Figure 14. Similar to the LP wellpair, in the
RP wellpair over the first two years the injection constraint
was a maximum bottom hole pressure limitation and beyond
that, the constraint was the steam injection rate, which was
adjusted in a downward trend until the end of the forecast.
During most of the life of the RP wellpair, the pressure was
roughly 1000 kPa.
The rate profiles for the RP wellpair presented in Figure 14 are
similar to LP wellpair. The cSOR profile passes through an
initial transient period and then evolves to a uniform value of
about 2.4 m3/m3. The bitumen rate profile has a typical shape
and hovers at over 60 m3/day for most of the life of the
wellpair. As the steam injection rate declines, the bitumen
rate also falls. Similar to the LP wellpair, throughout the
majority of the life of the RP wellpair, more water is produced
than is injected into the reservoir. This means that formation
water is being produced from the reservoir.
Compared to the constant pressure injection cases described
above, the cSOR of the optimized cases are significantly
improved. In the constant pressure case, the cSOR after six
and twelve years are 188 and 533 m3/m3 in the LP wellpair
and are 96 and 259 m3/m3 in the RP wellpair, respectively.
The normalized average bitumen production rate (normalized
against the well length = 750 m) of the constant pressure cases
are 0.0077 and 0 m3/day/m in the LP wellpair and 0.025 and

Figure 14: Production rates, cumulative volumes, and


cSOR of optimized strategy in RP wellpair.

In the optimized LP wellpair, the cSOR after six and twelve


years equals 2.6 m3/m3 at both times. The normalized average
bitumen production rate at six and twelve years are 0.068 and
0.067 m3/day/m, respectively. In the optimized RP wellpair,
the cSOR is 2.4 m3/m3 at both six and twelve years. The
normalized average bitumen rate is 0.071 and 0.069 m3/day/m
at six and twelve years, respectively. The reason for improved
performance in the optimized strategies is that after the steam
chamber contacts the top gas zone, the injection pressure
drops to values below that of the gas cap pressure. As a
consequence, steam does not invade and penetrate the gas cap
and is not lost from the depletion chamber. This implies that
the steams latent heat is more directly transferred to the
bitumen at the edges of the chamber than lost to the gas zone
and overburden. Also, as the pressure falls, the saturation
temperature falls and some of the invested heat in the
formation and overburden rock is harvested back to the
chamber fluids.

SPE/PS-CIM/CHOA 97742

Figures 15, 16, and 17 shows the sequence of temperature, oil


saturation, and flowing steam quality cross-sections roughly
two-thirds downwell for the LP wellpair, respectively. The LP
wellpair and gas cap pressure-maintenance wells are
displayed. SAGD mode starts 2005-04-01 after circulation.
The temperature, oil and gas saturation distributions are
displayed in Figures 18 to 20 for the RP wellpair.

Figure 16: Oil saturation distribution in section two-thirds


downwell of LP wellpair (optimized strategy).

Figure 15: Temperature distribution in section two-thirds


downwell of LP wellpair (optimized strategy).

The flowing steam quality is determined by calculating the


volume of mobile water in the vapour phase, converting it to a
mass of water in the vapour phase, and dividing it by the total
mass of mobile water (both vapour and liquid) in the
gridblock. This is a novel way to visualize the steam chamber
in SAGD where heat transfer is viewed as a change in the

10

flowing steam quality. Because temperature and pressure are


nearly constant in SAGD, the flowing steam quality provides a
means to examine convective heat transfer in the steam
chamber.

SPE/PS-CIM/CHOA 97742

liquid water phase in the injection stream falls under gravity to


the region below the injection well whereas the vapour rises
into the steam chamber. In the region between the wells, as a
consequence of this mechanism, the flowing steam quality is
relatively low because the liquid water content, derived from
injected liquid water, is relatively high in this region. Above
the injection well, the flowing steam quality forms a nearly
uniform radial profile moving away from the injection well.

Figure 17: Flowing steam quality distribution in section


two-thirds downwell of LP wellpair (optimized strategy).
At the gridblock above the injection wellblock the flowing
steam quality is higher than the injected steam quality. The
reason for this is because there is a separation effect as the

Figure 18: Temperature distribution in section two-thirds


downwell of RP wellpair (optimized strategy).

SPE/PS-CIM/CHOA 97742

11

evolves. The visualizations, especially at 2008-04-01, reveal


that the gas zone has thickened just above the chamber, most
likely because heated oil just above the chamber is flowing
downwards under gravity and gas has fingered up to the gas
zone from the chamber. At this particular section of the
reservoir, the chamber reaches the top gas zone between 3 and
4 years after SAGD starts.

Figure 19: Oil saturation distribution in section two-thirds


downwell of RP wellpair (optimized strategy).

Figure 18 displays the temperature distributions of the RP


wellpair roughly two-thirds downwell. A comparison with
Figure 15 shows that the RP wellpair operates at slightly lower
temperature than the LP wellpair and that the thermal zone
grows more in the vertical direction in the LP wellpair than it
does in the RP wellpair. Figure 20 shows how the RP wellpair
gas cap zone and steam chamber interact as the process

Figure 20: Gas saturation distribution in section twothirds downwell of RP wellpair (optimized strategy).

12

SPE/PS-CIM/CHOA 97742

Results from Dual Wellpair Model


After the individual LP and RP wellpairs were optimized, the
two operating strategies were integrated into the dual wellpair
model. Figure 21 displays a comparison between the
production rates and cSOR calculated from the LP wellpair in
the dual wellpair model versus the single LP wellpair model.

Figure 21: Production rates and cSOR of the single LP


wellpair model versus the result for the LP wellpair from
the dual wellpair model.
Figure 22 shows a comparison between the production rates
and cSOR calculated from the RP wellpair in the dual wellpair
model versus the single RP wellpair model.

Figure 22: Production rates and cSOR of the single RP


wellpair model versus the result for the RP wellpair from
the dual wellpair model.

The results show that communication occurs between the two


wellpairs after about one year of SAGD mode. This
communication is in the form of the pressure fields interacting
from the two wellpairs. However, fluid communication
between the wellpairs is small and so differences of the rate
profiles between the single-wellpair and dual-wellpair models
are not significant.

Figure 23: Oil saturation distribution in section two-thirds


downwell of dual wellpair model (optimized strategy
obtained from single wellpair models).

SPE/PS-CIM/CHOA 97742

13

The results of the SAGD models demonstrate how the steam


chambers grow to the top gas zone as SAGD proceeds in each
wellpair. Due to geological differences, the steam chamber
interacts and reaches the top gas zone first with the LP
wellpair compared to the RP wellpair.
The oil production rates from each of the wellpairs are similar
but the RP wellpair has slightly lower cSOR than that in the
LP wellpair. The reason for this is explained by the earlier
interaction of the left steam chamber with the top gas zone.
After the steam chamber reaches the top gas zone, steam, that
is, latent heat, is delivered to the top gas instead of being
delivered entirely to bitumen. The gas saturation distributions
in Figure 24 reveal that the RP wellpair chamber, up to 2010,
was growing faster in the lateral (horizontal) direction than in
the vertical direction. A comparison of the cSOR profiles in
Figures 20 and 21 shows that the cSOR of the RP wellpair was
slightly lower than that of the LP wellpair. This comparison
reflects the higher thermal efficiency achieved when
breakthrough to the top gas is delayed. These results indicate
that after the chamber is near the top of the oil pay, it is
advantageous to lower the rate of vertical growth of the
chamber to prevent penetration into the top gas zone. At this
point, it would be advantageous to promote, if possible, lateral
growth of the steam chamber.
Steamtrap Control and Subcool
As was described above, at the production wells, a 5C
subcool in the CMG algorithm means that the bottomhole
pressure in the wellblock is set corresponding to the pressure
5C above the saturation temperature of the gridblock and
therefore no live steam can be produced from the well. In the
field, the steamtrap subcool temperature difference is defined
as the difference between the steam injection temperature and
the produced fluids temperature. Figure 25 displays the
temperature difference between the injection and production
wells for the LP and RP wellpairs in a section roughly twothirds downwell of the dual-wellpair model.
80

LP Wellpair

70

RP Wellpair

60
50
40
30
20
10
2017-04-01

2015-04-01

2013-04-01

2011-04-01

2009-04-01

2007-04-01

0
2005-04-01

Subcool
Temp.
Difference,
Subcool
(Inj.
T - Prd. T) C

90

Time (Date)

Figure 25: Subcool temperature difference in LP and RP


wellpairs obtained from dual-wellpair model.
Figure 24: Gas saturation distribution in section twothirds downwell of dual wellpair model (optimized strategy
obtained from single wellpair models).

In the first year and a half of SAGD, the subcools of both


wellpairs oscillates and achieves high values. From Figures

14

21 and 22, the cSOR achieves its maximum values in this time
interval. To recall, the optimized operating strategy had
relatively high pressure steam injection in the first year at
1800 kPa and lowered it to about 1100 kPa thereafter. After
one year of SAGD, the subcool in the LP wellpair climbed to
over 60C for a couple of months. In the same time interval,
the RP wellpair subcool dropped to zero for several months.
When the injection pressure was lowered after one year to
1100 kPa, the LP wellpair responded by a jump in liquid
production, which then led to cooler liquid being produced
from the LP producer and consequently a higher subcool.
After the system stabilized, the subcool reached the steady
value between 20 and 30C which persisted throughout the
remainder of the process. In the RP wellpair, after the
injection pressure was reduced to 1200 kPa, liquid production
rates did not rise immediately but initially remained roughly
constant and consequently, the produced fluids temperature
remained roughly the same. However, because the injection
pressure was reduced, its temperature also fell and the subcool
became nearly zero. Beyond about two years, the subcool of
the RP wellpair also stabilized between 20 and 30C. The
differences in the subcool behaviour revealed by the
simulation reflect the difference between the geology at each
of the wellpairs and its impact on the growth of the steam
chambers. From the gas saturation distributions displayed in
Figure 23, the LP wellpair has a smaller steam chamber than
that in the RP wellpair after one year of SAGD.

SPE/PS-CIM/CHOA 97742

reservoir rock directly above the steam chamber. In the period


between 2006 and 2008 the steam chamber grows roughly 20
m in the vertical direction indicating a vertical rise rate of
about 2.7 cm/day. Given the temperature profile and due to
the chamber being at saturation conditions, beyond 2008, the
pressure is largely constant across the steam chamber. Thus
the temperature and pressure gradients in the steam chamber
are very small.
Figure 28 displays the flowing steam quality profiles. At the
start of SAGD model in 2005-04-01, the quality profile
exhibits two peaks, one at the injection well elevation and the
other at the production well location. The reason for these two
peaks is because circulation was occurring in both wells and it
was sufficiently hot that there is steam in reservoir at the well
locations. After a quasi steady state has evolved, beyond
about 2008, the vertical quality profile has roughly constant
slope equal to 0.016 quality units per meter. Beyond the
steam chamber, the flow steam quality falls rapidly to zero.
This is the location where the steam is releasing all of its latent
heat to the oilsand at the edges of the steam chamber.

Energetics of SAGD
As has been described above the optimum steam chamber
operating strategy employed in all the cases we have reviewed
have a common methodology. This methodology calls for
maintaining a high steam chamber pressure early in the SAGD
process. The higher steam chamber pressures lead to faster
chamber growth and to higher chamber temperatures. This in
turn leads to a favourably higher oil production profile. In
general, the higher the chamber temperature, the higher the oil
production. Eventually the steam chamber will contact the top
gas after which the steam chamber operating pressure is
dropped in line with the prevailing top gas pressure. In this
section, the relative roles of vertical and horizontal heat
transfer are investigated.

Figure 26:
Temperature through vertical profile
intersecting the LP wellpair.

The analysis here will focus on the LP wellpair model.


Figures 26 to 28 display the temperature, pressure, and
flowing steam quality along a vertical plane intersecting the
wells roughly two-thirds downwell. The injection well is
located at 37 m whereas the production well is located at about
42 m (the distance equal to 0 m is located at the top of the
reservoir).
The plots in Figures 26 to 28 are good representatives of the
profiles at other locations along the wellpair. Beyond 2008,
the profiles at each time overlay each other suggesting that the
steam chamber is under quasi steady state conditions. Figure
26 shows that the temperature across the steam chamber is
roughly constant at 177C. The temperature profiles above
the chamber edge indicate conductive heat transfer into the

Figure 27: Pressure through vertical profile intersecting


the LP wellpair.

SPE/PS-CIM/CHOA 97742

Figure 28: Flowing steam quality through vertical profile


intersecting the LP wellpair.

15

Figure 30: Pressure through horizontal profile 2 m above


the LP wellpair injector.

Figures 29 to 31 display the temperature, pressure, and


flowing steam quality along a horizontal plane located 2 m
above the injection well. The injection and production wells
are located at 92 m. The plots in Figures 29 to 31 are good
representatives of the profiles at other locations across the
domain.
The temperature profiles in Figure 29 demonstrate the lateral
growth rate of the steam chamber. Beyond 2008, the lateral
growth rate is on average 0.7 cm/day. This is roughly 3.8
times less than the initial vertical growth rate.
The
temperature in the steam chamber is roughly constant at
177C. As expected due to saturation conditions in the steam
chamber, Figure 30 confirms that the pressure is also nearly
constant within the steam chamber.

Figure 31: Flowing steam quality through horizontal


profile 2 m above the LP wellpair injector.

Figure 31 shows the flowing steam quality profiles through


horizontal section 2 m above the injection well. In the steam
chamber, the steam quality profiles have an average slope
roughly equal to 0.02 quality units per m.
As a means to analyze the quality slopes in the vertical and
horizontal directions, the heat balance can be used. The heat
balance is derived by considering a small volume element of
dimensions dx, dy, and dz as shown in Figure 32 and is given
by the statement that the net heat into the element is equal to
the heat flux, Q , lost from the element.
The heat balance is given by:
Figure 29: Temperature through horizontal profile 2 m
above the LP wellpair injector.

(q x dydzdt ) + (q y dxdzdt ) + (q z dxdydt )


= [(q x + dq x )dydzdt ] + [(q y + dq y )dxdzdt ]
+[(q z + dq z )dxdydt ] + Q dxdydzdt

16

SPE/PS-CIM/CHOA 97742

Given the definition of steam quality, f, the heat flux in the x


direction given by Equation 4 can be re-expressed as:

qz+dqz

q x = kTH

dz
q x = kTH

qx+dqx

qx

dy
Q

qz

q y = kTH

Figure 32: Heat transfer into and out of a differential


element.

After dividing by the elemental dimensions, taking the limit as


they approach zero, the result is:
q x q y q z 
+
+
+Q =0.
x
y
z

T
+ [V uV hV + L u L hL ]x
x

Often, the enthalpy is expressed as a function of temperature


by using the specific heat capacity to express Equation 3 as:
T
+ V uV c pV T Tref + L u L c pL T Tref
x

)]

[( f + hL )mx ] + ( f + hL )m y + [( f + hL )mz ]
x
y
z
= Q

10


T
T
T
+ kTH
kTH
+ kTH

x
x y
y z
z

Within a SAGD steam chamber, the temperature and pressure


are nearly constant. This means that the heat conduction terms
in Equation 10 are very small and that the latent heat of
vapourization and liquid enthalpy are essentially constant.
Applying constant temperature and pressure to Equation 10
yields:

[( f + hL )mx ] + ( f + hL )m y + [( f + hL )mz ] = Q
x
y
z
11

or after re-arrangement and canceling terms associated with


mass continuity:


[ fmx ] + fm y + [ fmz ] = Q

x
y
z

[ ]

T
+ [mVx hV + mLx hL ] .
x

T
+ [ f + hL ]m y
y

T
9
+ [ f + hL ]mz
z
After substituting Equations 7-9 into Equation 2, the result is:

where kTH is the thermal conductivity, T is the temperature, V


and L are the vapour and liquid densities, uV and uL are the
vapour and liquid phase (x-directed) velocities, and hV and hL
are the vapour and liquid phase specific enthalpies. The first
term on the right side of Equation 3 is the conductive heat
transfer term whereas the second term on the right side is the
convective term. The x-directed vapour and liquid phase mass
fluxes, mV = V uV and mL = LuL , can be substituted into
Equation 3 to give:

q x = kTH

q z = kTH

In the x direction, the heat flux in each direction is the sum of


the conductive and convective terms:

q x = kTH

T
+ [ f + hL ]mx
x

where is the latent heat of vapourization, given by


= hV hL , and mx is the total (vapour and liquid) mass flux.
Similarly, the y and z directed heat fluxes are given by:

qy

q x = kTH

or

qy+dqy

dx

T
+ [ fm x hV + (1 f )mx hL ]
x

12

Equation 12 reveals that the energy needed to account for the


heat losses, Q , are generated from a loss of steam quality in
the domain. For one-dimensional flow, Equation 12 simplifies
to:

which results in the more often found temperature-based


convective heat transfer term in the differential heat balance.

Q
f
=
m x
x

13

From observation, it can be seen that in a constant

SPE/PS-CIM/CHOA 97742

17

temperature, pressure domain where there are heat losses, the


steam quality will fall with distance. In one-dimension, if at x
= x0, the steam quality is f0, then the steam quality distribution
is given by:
14

Another way of interpreting Equation 13 is that the larger the


steam quality gradient, the larger the specific heat losses at the
edges of the steam chamber, that is, the larger the heat
transfer. On considering the steam quality slopes in the
vertical (0.016 m-1) and horizontal (0.02 m-1) directions
determined from Figures 28 and 31, because the quality
gradient is larger in the horizontal direction, more of the heat
supplied is being directed towards expanding the chamber
laterally as opposed to feeding overburden losses. This is a
key result that indicates that the optimized steam injection
strategy provided less heat flux in the vertical direction than
the horizontal direction. This means that there was less heat
transferred to the overburden and convective losses to the gas
cap zone and most of it was directed to growth at the sides of
the chamber.
In a 3D SAGD chamber, there are not only heat losses in the
plane (lateral-vertical plane) but also from heat losses in the
downwell direction. If the temperature varies in the downwell
direction, then there is also conductive heat transfer in the
downwell direction. This will further impact the steam quality
variation at any lateral-vertical slice of the steam chamber.
Enthalpy versus Temperature
Figure 33 displays a steam saturation curve and the enthalpy
versus temperature two-phase envelope. Within the envelope,
both vapour and liquid exist. The lines within the two-phase
region are constant steam quality lines. The topmost line
represents the 100% steam quality line whereas the bottommost line represents the 0% steam quality line.
Consider the situation where the steam chamber exists at state
A. The steam chamber is at nearly constant temperature and
pressure and so sits at point A throughout the chamber. If the
steam quality is known at the steam injector, then the enthalpy
of the injected steam is known. If the steam quality is 90%,
then the injected steams enthalpy is given by the value at
point B. Due to heat losses to the outer regions of the steam
chamber, the steam quality in the chamber falls moving away
from the injection well. Because the temperature and pressure
are roughly constant throughout the chamber, the enthalpy at
any point in the chamber will lie along the line from point B to
point C. If the steam loses all of its latent heat, it reaches the
point C and is liquid water. If further cooling occurs, then a
reduction of the liquid water temperature results. As the
pressure of the process evolves, given that the chamber is at

9000

Saturation Curve

8000

Pressure,
Pressure
(kPaa)kPa

which reveals that providing the ratio between the heat losses
to mass flow are constant, that the steam quality drops linearly
with distance.

10000

7000
6000
5000
4000
3000

2000
1000
0
100

150

100% Quality

Enthalpy, kJ/kg

Q
(x x0 )
mx

Specific Enthalpy, kJ/kg

f = f0

saturation conditions, point A will shift along the steam


saturation curve. For example, if the injection pressure falls as
the process evolves, point A will move in the left direction
along the saturation curve. As a consequence, points B and C
will also shift in the left direction.

3000
2750
2500
2250
2000
1750
1500
1250
1000
750
500
250
0
100

200

250
300
Temperature (C)

400

150

350

200
250
300
Temperature, Deg. C

0% Quality

350

400

Figure 33: Steam saturation curve and enthalpy versus


temperature diagram.

The extents of the vertical profiles of the flowing steam


quality can be plotted on the enthalpy versus temperature
diagram as shown in Figure 34. This is a subregion of the
diagram displayed in Figure 33.
The vertical portions of the profiles, above the dotted line in
Figure 34, represent the part of the chamber that is nearly all
steam and is at saturation conditions. Here the temperature is
roughly constant and heat tranfer throughout the chamber is
mainly due to convection and is reflected by the steam quality
variation. At the edges of the chamber, the portions of the
plots below the dotted line, steam is losing its latent heat, the
quality rapidly falls and the specific enthalpy of the wet steam
is relatively low. Also, partial pressure effects come into play
because solution gas comes out of solution from the bitumen
at the edges of the chamber. The top part of each profile
represents the enthalpy of the steam near the injection well
(highest quality and consequently enthalpy) whereas the

18

SPE/PS-CIM/CHOA 97742

Specific
Enthalpy,
Enthalpy,
kJ/kgkJ/kg

bottom parts of the curves represent the top of the vapour


chamber where the quality transitions from that at steam
chamber conditions to zero (where the plots cross the zero
quality line).

3000
2750
2500
2250
2000
1750
1500
1250
1000
750
500
250
0
120

100% Quality

1: 2006-04-01

2: 2008-04-01
3: 2010-04-01

4: 2012-04-01

5: 2014-04-01
6: 2016-04-01
130

140

150
160
Temperature, Deg. C

170

0% Quality
180

190

Figure 34: Vertical flowing steam quality profiles in form


of enthalpy versus temperature for LP wellpair.

along the wellbore to ensure full contact of the steam chamber


to the reservoir penetrated by the horizontal well.
The optimized steam injection strategy promotes heat transfer
in the lateral direction over that in the vertical direction. This
reduces heat losses to the overburden and convective losses to
the top gas zone. The amount of heat supplied by the injected
steam should be only that required to mobilize the bitumen at
the edges of the chamber, not to induce large conductive
losses to the overburden.
The dual wellpair operating strategy was determined by
optimizing the individual wellpair reservoir simulation models
separately and then integrating the optimized operating
strategies into the dual wellpair model. Due to the existence
of the top gas zone, the strategies developed for each of the
single wellpair models were designed to be gentle and not
promote extensive communication with the top gas zone. For
this reason, the two individually optimized operating strategies
could be implemented together in the dual wellpair model.
The dual wellpair model revealed that communication
between the two wellpairs did influence the performances of
the wellpairs slightly.

Acknowledgements
After the chamber has matured, after about 3 years of
operation, the maximum and minimum specific enthalpies of
the steam (beyond the chamber edge, that is, above the dotted
line in Figure 34) are roughly constant throughout the
chamber. The profiles in the edge region of the chamber
(below the dotted line) become more vertical as the process
evolves due to continued conductive heating of the materials
above the edge of the chamber.

The authors would like to acknowledge Paramount Resources


Ltd. for permission to publish this study.

References

1.
Conclusions
The flowing steam quality provides a novel method to
visualize heat transfer within and at the boundaries of the
steam chamber. This is useful because the pressure and
temperature of the steam chamber are nearly constant. The
flow steam quality profiles provide a means to examine
convective heat transfer in the reservoir.

2.

3.
4.

The optimized operating strategy for SAGD in reservoirs with


top gas have high initial chamber injection rates and pressures
prior to chamber contact with the top gas. Subsequent to
breakthrough the chamber injection rates are lowered to
balance pressures with the top gas and so avoid or at the least
minimize convective heat losses of the steam to the top gas
zone. The lower the top gas pressure, the lower the chamber
pressure and therefore the lower the chamber temperature. In
addition to avoiding convective heat losses to top gas, the
lowering of chamber pressure also reduces conductive heat
losses to the over and under burden and allows harvesting of
heat stored in the chamber prior to breakthrough. The lower
chamber temperature does however lead to higher bitumen
viscosities.
To optimize SAGD, steam conformance must be managed

5.

6.

7.

8.

SINGHAL, A.K., ITO, Y., and KASRAIE, M. Screening


and Design Criteria for Steam-Assisted Gravity Drainage
(SAGD) Projects. SPE Paper 50410, 1998.
KOMERY, D.P., LUHNING, R.W., and OROURKE,
J.C. Towards Commercialization of the UTF Project
Using Surface Drilled Horizontal SAGD Wells. J. Can.
Pet. Tech. 38(9):36-43, 1999.
BUTLER, R.M. Thermal Recovery of Oil and Bitumen.
GravDrain Inc. 1997.
KISMAN, K.E., and YEUNG, K.C. Numerical Study of
the SAGD Process in the Burnt Lake Oil Sands Lease.
SPE Paper 30276, 1995.
ITO, Y., and SUZUKI, S. Numerical Simulation of the
SAGD Process in the Hangingstone Oil Sands Reservoir.
J. Can. Pet. Tech. 38(9):27-35, 1999.
EDMUNDS, N. and CHHINA, H. Economic Optimum
Operating Pressure for SAGD Projects in Alberta. J. Can.
Pet. Tech. 40:13, 2001.
ITO, Y., HIRATA, T., and ICHIKAWA, M. The Effect
of Operating Pressure on the Growth of the Steam
Chamber Detected at the Hangingstone SAGD Project. J.
Can. Pet. Tech. 43(1):47-53, 2004.
SALTUKLAROGLU, M., WRIGHT, G.N., CONRAD,
P.R., MCINTYRE, J.R., and MANCHESTER, G.J.
Mobils SAGD Experience at Celtic, Sashatchewan.
Paper 99-25
CSPG and Petroleum Society Joint

SPE/PS-CIM/CHOA 97742

9.

10.

11.

12.

13.

14.

15.
16.
17.

18.

Convention in Calgary, Alberta, Canada, 14-18 June,


1999.
SUGGETT, J., GITTINS, S., and YOUN, S. Christina
Lake Thermal Project. SPE/Petroleum Society of the
CIM Paper 65520, 2000.
SIU, A.L., NGHIEM, L.X., GITTINS, S.D., NZEKWU,
B.I., and REDFORD, D.A. Modelling Steam-Assisted
Gravity Drainage Process in the UTF Pilot Project. SPE
Paper 22895, 1991.
AED (Alberta Economic Development). Oil Sands
Industry Update. Available at Alberta Dept. of Energy
Website: http://www.energy.gov.ab.ca/com/default.htm.
March 2004.
YEE, C.-T. and STROICH, A. Flue Gas Injection Into a
Mature SAGD Steam Chamber at the Dover Project
(Formerly UTF). J. Can. Pet. Tech. 43(1):54-61, 2004.
GATES, I.D., and CHAKRABARTY, N. Optimization
of Steam-Assisted Gravity Drainage (SAGD) in Ideal
McMurray Reservoir. Paper 2005-193 presented at
Canadian International Petroleum Conference, Calgary,
Alberta, Canada, June 7-9, 2005.
ROBINSON, W., KENNY, J., HERNANDEZ-Hdez, I.L.,
BERNAL, A., CHELAK, R. Geostatistical Modeling
Integral to Effective Design and Evaluation of SAGD
Processes of an Athabasca Oilsands Reservoir, A Case
Study.
Paper 97743 presented at the 2005 SPE
International Thermal Operations and Heavy Oil
Symposium held in Calgary, Alberta, Canada, 1-3
November, 2005.
ROXAR. IRAP RMS Users Manual. 2004.
Computer Modelling Group (CMG) Ltd. STARS Users
Manual, Version 2004.10. Calgary, Alberta, Canada.
MEHROTRA, A.K. and SVRCEK, W.Y. Viscosity of
Compressed Athabasca Bitumen. Can. J. Chem. Engrg.
64:844-847, 1986.
BOONE, T.J., YOUCK, D.G., and SUN, S. Targeted
Steam Injection Using Horizontal Wells with Limited
Entry Perforations. SPE 50429, 1998.

19

Anda mungkin juga menyukai