SPE 185015-MS Chemical Additives and Foam Assisted SAGD Model Development

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 16

SPE-185015-MS

Chemical Additives and Foam Assisted SAGD Model Development

Ran Li, Danling Wang, and Zhangxin Chen, University of Calgary

Copyright 2017, Society of Petroleum Engineers

This paper was prepared for presentation at the SPE Canada Heavy Oil Technical Conference held in Calgary, Alberta, Canada, 15-16 February 2017.

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
Steam assisted gravity drainage (SAGD) is recognized as a profitable and stable approach to address
the exploitation of heavy oil and oil sand resources. However, the efficiency of SAGD, a close relative
of a sufficiently-expanded and uniformly-developed steam chamber, tends to be deteriorated by quick
steam movement and high heterogeneity. Chemical additives and foam assisted SAGD (CAFA-SAGD) is
a strategy proposed on this account. This study aims to analyze the mechanisms and phenomena involved.
The injection of chemical additives to promote in-situ foam generation reduces gas relative permeability
by slow-moving and stagnant bubbles trapping. Also, lamella resists bubbles flow and increases apparent
gas viscosity. The restriction of steam mobility thus favors a sufficiently-expanded steam chamber and the
nitrogen co-injected to stabilize bubbles works as a separator between steam and overburden to reduce heat
loss. Simultaneously, the interfacial tension reduction due to surfactants injection at a water/oil interface
may influence phase behavior, which further leads to the solubilisation of residual oil. CAFA-SAGD is thus
likely to increase heat efficiency and add oil output.
A homogeneous model is built to analyze CAFA-SAGD considering foam generation by snap-off and
leave-behind, foam trapping in a porous medium and foam coalescence due to both the lack of surfactants
and capillary suction. Besides, with the analysis of foam wall slip phenomena, a comprehensive foam
property model is coupled to analyze shear thinning rheology and calculate lamella viscosity as a function of
gas saturation and gas velocity. In addition, the influences generated by surfactant injection should be added.
This study also develops an analytical FA-SAGD model based on Butler's finger rising model (1987) to
show foam's effects on a steam chamber growth rate and shape. We derive the FA-SAGD model accounting
for the retarded steam movement with higher steam viscosity and lower gas relative permeability. The foam
viscosity is calculated as a function of gas saturation and a gas rate, and the modification of gas relative
permeability is reflected with a higher gas residual saturation according to Bertin et al.'s foam property
model (1998). After comparing, validating, and discussing the developed model against the SAGD model,
we find that foam injection contributes to high production efficiency with less steam consumption. A lower
steam mobility generated by stronger foam is more likely to have a lower SOR (steam-oil ratio).
The results agree well with the published high-temperature steam foam experiments and pilot tests. Strong
bubbles accumulate along the boundary of a steam chamber to restrict steam movement, while weak foam
fills inside the chamber to enhance steam trap, contributing to a higher oil recovery factor and lower SOR.
2 SPE-185015-MS

Introduction
Steam injection is a well-developed and widely used oil recovery method to carry out heavy oil extraction in
industry. A large sum of enthalpy carried by steam is transferred into cold bitumen through conduction and
convection. The highly viscous oil is thus mobilized after a specific amount of heat absorption. However,
like other gas assisted recovery processes, due to a large density difference and viscosity difference between
the gas and liquid phases, the low sweep efficiency generated by steam override and steam fingering tends
to influence the oil yield significantly. To address this issue, people make great efforts to look for gas-flow
blocking agents to reduce the mobility ratio.
Use of steam foams has been studied as a technique to promote large mobility reductions and improve
vertical and areal conformance during steam injection applied to shallow heavy oil reservoirs. Shell
conducted two steam foam pilots in Mecca Lease and Bishop Fee in the Kern River field in the 1980s.
Foam was generated by continuous injection of 50% quality steam, 4% NaCl, 0.5% surfactant and 0.06%
nitrogen after pure steam injection for several years. The Bishop pilot achieved 5.5% more oil production in
the end (Patzek amd Myhill, 1989). Another steam foam pilot was carried out in the McMurray formation
of the Athabasca Oil Sands in Alberta in 1988. To mobilize the high viscous oil, surfactant was injected at
approximately 0.5-1.4 m3/day. Foam generation contributed to the downward movement of the steam zone
and the oil production was increased by around 30% (Kular et al., 1989).
Foam is selected to carry out mobility control because of its characteristic structure. On the microscopic
scale, foam is defined as a dispersion of gas in the continuous water phase with thin films (lamella) acting
as a separation. In this way, the gas flow path is blocked to a large extent and gas relative permeability
is reduced consequently. On the other hand, lamella works to create flow resistance towards the flowing
bubbles, leading to a higher gas viscosity. Therefore, it is possible to use foam to block thief zones and reduce
channeling through decreasing steam mobility (Hirasaki, 1989; Patzek, 1996). Thermally stable surfactants
are essential to maintain the foam life to prohibit the trend of bubbles coalescence as surfactants stabilize the
lamella by decreasing the water-gas interfacial tension. Simultaneously, the interfacial tension reduction due
to surfactants injection at a water/oil interface may influence interfacial phenomenon and further leads to
the solubilisation of residual oil. CAFA-SAGD is thus likely to increase heat efficiency and add oil output.
In this study, foam is in-situ generated under high temperature and high pressure conditions to assist
SAGD production in a simulator. Foam generation by snap-off and leave-behind, foam trapping in a porous
medium and foam collapse due to high temperature and low surfactant concentration are incorporated to
determine the steam mobility by which we are able to check whether foam can preserve a sufficiently-
expanded and uniformly-developed steam chamber through conformance improvement to enhance SAGD
performance.
In addition, Butler's finger rise model is modified to account for the retarded steam movement through
introducing a higher steam viscosity and a lower gas relative permeability as a result of foam generation.
According to the previous research, bubbles strength and foam performance largely depend on foam texture,
which is a function of porosity, permeability, a gas flow rate, gas saturation, and surfactant type. On this
account, a bubble-population foam model put forward by Bertin et al. (1998) is incorporated with Butler's
finger rise model to study the effects of gas saturation and gas velocity towards foam texture and the
consequent influences towards CAFA-SAGD performance.

Steam Foam Simulation


Bubbles containing a continuous gas phase are regarded as weak foam, which, in most cases, is produced by
leave-behind. It happens when the continuous gas phase invades pores in the condition that both the capillary
pressure and the pressure gradient are not too large. At this time, stationary stable lamella parallel to the
flow direction is generated. Due to the existence of continuous gas channels, gas is able to flow through
a porous medium without displacing the trapped lamella, which reduces the gas flow cross-sectional area.
SPE-185015-MS 3

Therefore, leave-behind is the dominant foam generation mechanism when the gas velocity is low and the
weak foam mainly reduces the gas relative permeability. In contrast, strong foam contains a gas phase that is
made discontinuous by snap-off and lamella division. Without a continuous flow path, the generated lamella
needs to be transported. In this way, discrete bubbles work to block the gas channels and provide resistance
to the flowing gas, leading to a large reduction in the gas mobility (Schramm, 1994).
Since bubbles always have a tendency to shrink a surface area to achieve the lowest free energy, it
happens frequently that gas diffuses from small bubbles to large bubbles. The enlarged bubbles are likely to
collapse as long as a film becomes unstable due to the lack of average surfactants distribution. Besides, foam
coalescence occurs where the water saturation drops below the critical water saturation, which produces a
large capillary pressure draining the film to be thinner and thinner until foam collapses. Previous research
considered that the capillary suction governs foam coalescence in strong foam, which transforms into weak
foam abruptly around the critical capillary pressure (Schramm, 1994). The last and most severe reason for
bubbles rupture is the defoaming activity of oil droplets. Farajzadeh et al. (2012) concluded that surfactant
depletion, the penetration and spreading of oil globule into a foam film and the bridging of the film by oil
droplets were three primary causes of film rupture in the presence of oil.
Foam flow in a porous medium is a complex and ever-changing process as its gas phase and liquid phase
adopt different flow patterns. Gauglitz et al. (2002) maintained that the liquid phase flows through lamella
and smaller pores primarily while the gas phase occupies larger throats.
In this study, the foam property data comes from high-temperature foam displacement experiments
conducted by Friedmann et al. in 1991. They generated nitrogen foam in Berea sandstone cores at high
temperature and determined the foam model parameters through history match. The primary conditions
to have foam were the presentation of gas, an oil saturation higher than a limiting value and a surfactant
concentration that was high enough to maintain the lamella stability. To simplify the problem, the foam
generation mechanisms were limited to snap-off and leave-behind. The gas velocity was also checked
to further determine the dominant mechanism as the former worked at a high velocity while the latter
occurred in the low velocity area. First, a high temperature static lamella core test was run to estimate the
reduction of gas relative permeability due to the formation of a stagnant gas phase under a low pressure
gradient. It was found that the gas relative permeability was reduced in dozens of times at 0.05 wt%-0.1 wt%
surfactant concentration and it changed little with an increase in gas saturation. Second, in terms of flowing
bubbles, modifications went to both apparent gas viscosity and gas relative permeability. In accordance
with the population balance approach, they measured the influence of a gas rate towards gas viscosity by
testing the pressure gradient under multiple gas frontal advance rates. In the meantime, the trapping fraction
was investigated with a gas tracer. Even though the trapping fraction was a function of gas velocity, gas
saturation, and capillary pressure, it was treated as a constant in experiments. Finally, the foam generation
rate constant and coalescence rate constant were determined through history match without account for the
temperature effect (Friedmann et al., 1991).
Based on the SAGD experiment carried out by Oskouei et al. (2012), a grid system of 17*1*20 with a grid
size of 2 cm*14 cm*2 cm is utilized to simulate the SAGD process. The heterogeneity along the horizontal
wells is ignored and the model is simplified to a 2D problem. Only half of the reservoir is considered. For
the homogeneous model, the porosity is 0.396 and the permeability is set to be 300 D to satisfy the scaling
requirement of the gravity drainage process. The steam with the quality of 0.99 is injected at the pressure
of around 4,000 kPa. The steam trap is set to be 5 °C approximately to prevent steam escaping. The injector
is drilled 12 cm above the producer. The basic reservoir parameters are summarized in Tables 1 and 2.
To compare the performance between SAGD and CAFA-SAGD, 0.015% (mole fraction) surfactants and
nitrogen are injected together with steam after 60 minutes' SAGD production, when the steam chamber is
sufficiently expanded. The foam property parameters illustrated above are incorporated in the CAFA-SAGD
model to simulate foam generation, bubbles collapse and foam behavior. Stationary lamella reduces the gas
relative permeability by 5-10 times and the trapping fraction is set to be 0.4 in terms of flowing bubbles.
4 SPE-185015-MS

As for the apparent gas viscosity, the effect of a gas rate is incorporated because the foam is regarded as
a shear-thinning fluid in the model. According to Friedmann et al. (1991), the dependence of viscosity on
the gas rate is around the power of -0.3. The temperature effect towards foam stability is also introduced
through a thermal decomposition reaction of surfactants. Besides, the oil threat to foam's life is added into
the model using another reaction with a larger foam collapse rate constant in the presence of oil.

Table 1—Basic reservoir parameters (Oskouei et al., 2012)

Parameter Value Parameter Value

Porosity 0.396 Rock Thermal Conductivity 0.0533 J/(cm*min*°C)

Permeability 300 D Rock Heat Capacity 0.99 J/(cm3*°C)

Initial Temperature 52 °C Water Thermal Conductivity 0.372 J/(cm*min*°C)

Initial Pressure 3550 kPa Oil Thermal Conductivity 0.09 J/(cm*min*°C)

Gas Thermal Conductivity 0.018 J/(cm*min*°C)

Table 2—Oil properties (Oskouei et al., 2012)

Temperature, °C Oil Viscosity, mPa·s Oil Density, g/cm3

25 44500 0.997

40 7550 0.988

55 1965 0.979

According to Figure 1, the steam chamber in CAFA-SAGD has a smaller expansion rate as foam restricts
steam flow in the lateral direction at the overburden. The contact area for heat loss is reduced, and more
energy is thus diverted to heat the cold bitumen in the lower area laterally. In contrast, the steam chamber in
SAGD spreads quickly and achieves an enormous heat exchange with overburden, which exposes the steam
chamber to experience more heat loss and achieve lower efficiency. The oil reserved in the bottom zone is
bypassed, which leads to a low sweep efficiency. It resembles a comparison between a quickly developed
fingering advance and a slow piston-like displacement.
SPE-185015-MS 5

Figure 1—Temperature comparisons between SAGD and CAFA-SAGD

Figure 2 depicts that the apparent gas viscosity in CAFA-SAGD is much larger compared with that in
SAGD. After surfactants injection and foam generation, the gas viscosity increases by dozens of times
inside the steam chamber. Moreover, the gas flow in CAFA-SAGD seems to be further resisted as it gets
closer to the steam front. Strong foam mainly accumulates along the inner border of the steam chamber,
developing a low gas rate circle surrounding the steam chamber. The reason is that the injected nitrogen,
which is lighter than the steam, fills in this area. Compared to steam, nitrogen is much more stable and
beneficial to maintain the foam strength. Inside the steam chamber is a certain amount of weak foam with
a smaller resistance factor.
6 SPE-185015-MS

Figure 2—Gas viscosity comparisons between SAGD and CAFA-SAGD

Consequently, the oil production is increased by around 15% due to a higher sweep efficiency and a
lower heat loss (Figure 3). Also, the injection of surfactants also gives rise to the water-oil interfacial tension
reduction. A larger capillary number is created and the oil adhesion force at the rock surface is decreased.
Oil droplets are more likely to be taken away and the oil displacement efficiency is improved in this case.

Figure 3—Cumulative oil production comparisons between SAGD and CAFA-SAGD


SPE-185015-MS 7

CAFA-SAGD is verified to be more efficient over SAGD by means of controlling steam movement
and improving oil production. Next, we continue to study how CAFA-SAGD behaves in a heterogeneous
situation. It is obvious that the low-permeable formations have little chance to be exposed to steam, and most
of steam may escape though a high-permeable area. Likewise, the intercepted water layers and surrounding
water zones may further disturb the temperature field and transform the steam chamber. It is speculated
that heterogeneity is also likely to convert distribution pattern of chemical additives and impact CAFA-
SAGD's final yields.
A 2D cross-sectional model is cut out from a 3D heterogeneous model built according to the Firebag
Project. Well Pair 7, because of a longer production history and a better performance, is chosen to inject
chemical additives to investigate how CAFA-SAGD performs. There are several low-permeable layers in
the upper part and a top water zone which works as a potential thief zone (Figure 4). We choose to inject
chemical additives once the steam chamber is relatively expanded and the steam front approaches a low-
permeable area to prevent the steam from bypassing the oil in the upper part and losing heat in time (after
conducting SAGD production for 6 months).

Figure 4—Properties distribution of 2D heterogeneous model (Well Pair 7)


(A: vertical permeability; B: porosity; C: water saturation; D: oil saturation)

Water has higher thermal conductivity and heat capacity. The existence of top water acts as a thief zone,
adsorbing heat and deteriorates heat efficiency when the steam chamber runs into it. Due to the existence
of top water, the steam chamber of SAGD grows quickly towards the enthalpy sink vertically instead of
horizontally at the lower part of the reservoir.
Once the chamber contacting the overburden, it stretches and extends at a large lateral rate along the
overburden immediately, achieving an enormous heat exchange surface and forming a plate-shape steam
8 SPE-185015-MS

chamber. The low-permeable layers which are less likely to be swept by steam lead to even worse results
as most of the steam tends to escape though the high-permeable area. A direct outcome is the excessive
lateral growth of the chamber along the high conductive area severely. Thus there are two recessions at the
contour (Figure 5). Consequently, the residual oil saturation is high in the area with poor pore structures
as it is not fully heated. Also, owing to the rapid vertical movement of steam, the oil reserved in the lower
reservoir is neglected (Figure 6).

Figure 5—Temperature distribution comparison of 2D heterogeneous model (A.1:


SAGD, 18 months after preheating; A.2: SAGD, 30 months after preheating; B.1: CAFA-
SAGD, 18 months after preheating; B.2: CAFA-SAGD, 30 months after preheating)

Figure 6—Oil saturation distribution comparison of 2D heterogeneous


model (42 months after preheating) (A: SAGD; B: CAFA-SAGD)

The generation of foam, especially the strong bubbles concentrating around the margin of the steam
chamber helps to slow down the steam rising velocity during CAFA-SAGD, consistent with the foam
SPE-185015-MS 9

distribution pattern of the homogeneous model. This phenomenon delays the vertical steam movement,
controls the trend of steam chamber lateral extension in the early days, reduces the heat consumption towards
the top water, and saves more energy to mobilize oil in the lower part. Also, it is speculated that the strong
bubbles generated in the high-permeable area force the steam to penetrate into the low-permeable regions
under a higher pressure. A certain number of weak bubbles accommodate in the low-permeable layers. In
this way, bubbles manage to push the steam front uniformly and control the steam chamber circumference
into a bowl-shape without severe fingering phenomenon (Figures 4 to 7). Finally, the performance of SAGD
and CAFA-SAGD is listed in Table 3. A smaller steam chamber volume signifies a lower expansion rate.
The performance is greatly improved in terms of oil production and SOR.

Figure 7—Lamella mole fraction distribution of 2D heterogeneous


model (18 months after preheating; B: 42 months after preheating)

Table 3—Performance comparison for heterogeneous SAGD and CAFA-SAGD cases

Case Cumulative Oil Production, m3 SOR, m3/m3 Steam Chamber Volume, m3

SAGD 5.57e4 3.48 8.66e4

CAFA-SAGD 1.27e5 2.47 5.62e4

Steam Foam Model


A steam front is not flat but composed of many small fingers. According to previous experiments SAGD
relies on those fingers protruding into a cold reservoir to transfer heat and further drain oil. Under
the assumptions that fingers rise at a constant velocity and there is no heat convection, Butler (1987)
developed an analytical model to predict the steam finger rising velocity and the corresponding oil flux.
Simultaneously, as illustrated above, with a lamella acting as a separator between the gas and aqueous
phases, gas movement is blocked and steam breakthrough to the overburden is controlled. In consequence,
the steam chamber expansion rate decreases and oil production increases. To incorporate the effects of foam
towards gas mobility, Li et al. (2006) put forwards to calculate foam viscosity as a function of gas saturation
and the gas rate and reflect the modification of gas relative permeability with a higher gas residual saturation.
With the above two models, we measure the influences of foam during SAGD.
According to Butler (1987), oil is assumed to flow parallel to the steam chamber boundary by gravity
while impeded by the pressure differential pressure of the gas phase. The oil kinematic viscosity increases
with an increasing distance away from the injector (Figure 8). At point P, where the boundary has an angle
θ, Darcy's law is represented by Equation (1).
10 SPE-185015-MS

Figure 8—Rising finger plot (Butler, 1987)

(1)

where Qo is oil volumetric flow across the normal plane; K is the reservoir permeability; Kro is the oil relative
permeability; g is the gravity acceleration; ρo is the oil density; is the pressure gradient; ε is the distance
away from the point P; ε1 is the distance between point P and the symmetry plane; ϑo is the oil kinematic
viscosity.
The pressure gradient blocks oil flow, but pushes the steam upward (Equation (2)).

(2)

where Qg is the gas volumetric flow; Krg is the gas relative permeability; x is the distance away from the
center on a finger; μg is the gas viscosity; ρg is the gas density.
After substituting Equation (2) into Equation (1) and eliminating gas density, oil flow is

(3)

where R is the ratio of the gas flow rate to the oil flow rate.
The oil kinematic viscosity varies with location mainly due to the temperature difference at different
locations. Equations (4) and (5) are combined to build relationship between oil viscosity and location
(Equation (6)).

(4)

(5)

(6)

where TR is the reservoir temperature; TS is the steam temperature; ϑOS is the oil kinematic viscosity at TS;
m is a constant depending on oil properties; α is the thermal diffusivity; t is time.
With the help of Equation (6), Equation (3) can be further transformed into Equation (7).
SPE-185015-MS 11

(7)

At a particular point at the interface when two fingers just interfere, oil flow becomes

(8)

where Qio is the oil flow from point P to the point of interference; x1 is half distance between the two fingers;
xi is the interface position.
The oil flow rate at a particular point can also be calculated with the reduction of oil content.
(9)
where Ø is the porosity; ΔSo is the alteration of oil saturation during production.
Finally, after the substitution of Equation (9) into Equation (8), the finger rising velocity is

(10)

where Xi = xi/x1 is the dimensionless finger width and R′ = Rρg/ρo.


The heat conservation law is applied to determine the steam-oil ratio. Butler (1987) maintained that the
heat released by the injected steam mainly went to the steam chamber, residual oil and connate water, and
the oil flow left the fingers as well as the oil saturated reservoir.
(11)
(12)

(13)

where Hc is the heat inside the steam chamber; ρcCc is the volumetric heat capacity of the steam chamber
together with the residual oil and connate water; Ho is the heat in oil flow leaving between the steam
chambers; ρoCo is the heat capacity of oil; Tm is the mixing temperature of oil; HR is the heat in the reservoir;
ρRCR is the heat capacity of the reservoir.
The mixing temperature Tm is

(14)

In total, heat consumption HT is

(15)

Supposing that all the heat is provided by steam latent heat λ, the required steam ms is

(16)

Therefore, the steam-oil ratio is finally expressed by Equation (17).

(17)
12 SPE-185015-MS

Finally, the steam rising rate ends to be

(18)

where

(19)

(20)

It is obvious that the steam velocity changes with the dimensionless width of fingers. The maximum
finger rising velocity is obtained after differentiating Xi.
Both gas relative permeability Krg and gas viscosity μg are modified to incorporate foam behavior.
According to Li et al. (2006), a larger gas residual saturation is introduced to recalculate the gas relative
permeability Krgf. In addition, the gas viscosity is calculated as a function of gas saturation Sg. Li et al.
(2006) maintained that a change in gas saturation controlled the lamella density, further determining the
foam viscosity (Figure 9). Equation (22) is simplified by neglecting the influence of gas velocity.

Figure 9—Lamella density and gas saturation relationship (Li et al., 2006)

(21)

(22)

where Srgf is flowing gas saturation; Fg is a geometric factor; nf is lamella density, Cnf is a coefficient related
to porosity and permeability; Sgm is the critical gas saturation; d and e are constants.
The data listed in Tables 4-6 is applied to compare Butler's model and the modified model to observe
the effects of foam generation.
SPE-185015-MS 13

Table 4—Data used for comparison (Gotawala and Gates, 2008)

Parameter Value Parameter Value

TR, °C 4 ρoCo, J/m3·°C 1,879,600

Ts, °C 200 k, D 3

P, kPa 1,553 2
α, m /s 7.06e-07

λ, J/kg 1,941,000 Soi 0.8

Ø 0.39 Sor 0.15

ρo, kg/m 3
1,005 kro 0.65

ρcCc, J/m3·°C 373,000 krg 0.4

ρRCR, J/m3·°C 2,184,000 m 4

Table 5—Fluid data used for comparison (Athabasca Oil, m=4) (Murtaza et al, 2014)

Ts νos, m2/s νgs, m2/s ρg, kg/m3

100 0.0003 1.71e-05 0.598

200 1.84e-05 2.04e-06 7.86

250 1.06e-05 8.76e-07 20

300 2.86e-06 4.26e-07 42.6

Table 6—Foam property (Li et al., 2006)

Parameter Value Parameter Value

Srgf 0.2 Sgm 0.8

e 2 d 5

Cμf 4500

According to Figures 10 and 11, there is no doubt that fingers rise more quickly at a higher temperature for
the two models due to the oil viscosity reduction. It is also obvious that the finger rising rate is significantly
reduced due to the generation of foam at the same injection pressure. In addition, the difference is larger
along with the increasing temperature. The finger rising velocity is reduced from 50 cm/day to less than
6.5 cm/day at the temperature of 300 °C. Figure 11 shows that with an increase in gas saturation between
Srgf and Sgm, the foam generation rate overtakes the foam coalescence rate. The steam expansion rate slows
down as a result of higher lamella density. Once the gas saturation reaches Sgm, foam collapses immediately
due to a large capillary pressure, leading to a quickly expanded steam chamber.
14 SPE-185015-MS

Figure 10—Finger rise velocity without foam

Figure 11—Finger rise velocity with foam

Conclusions
The SAGD process can be enhanced by foam generation due to higher sweep efficiency and lower heat loss
as foam is conducive to control steam mobility. The results agree well with the published high temperature
steam foam experiments and pilot tests. Strong bubbles accumulate along the boundary of a steam chamber
to restrict steam movement, while weak foam fills inside the chamber to enhance steam trap. An increase
in gas saturation improves foam performance through increasing lamella density. But, once it exceeds the
limit gas saturation, foam collapses quickly as a result of a large capillary pressure.

Acknowledgements
The authors thank the support from NSERC/AIEES/Foundation CMG IRC in Reservoir Simulation, AITF
(iCore) Chair in Reservoir Modelling and The Frank and Sarah Meyer Foundation CMG Collaboration
Centre.
SPE-185015-MS 15

Reference
Azom, P. N., & Srinivasan, S. (2009, October). Mechanistic modeling of emulsion formation and heat transfer during
the steam-assisted gravity drainage (SAGD) process. In SPE Annual Technical Conference and Exhibition. Society
of Petroleum Engineers.
Bertin, H. J., Quintard, M. Y., & Castanier, L. M. (1998). Development of a bubble-population correlation for foam-flow
modeling in porous media. SPE Journal, 3(04), 356–362.
Boeije, C. S., & Rossen, W. R. (2013, April). Fitting foam simulation model parameters to data. In IOR 2013-17th
European Symposium on Improved Oil Recovery.
Butler, R. M. (1987). Rise of interfering steam chambers. JCPT, 26(3), 70–75.
Butler, R. M. (1991). Thermal recovery of oil and bitumen.
Bryan, J. L., & Kantzas, A. (2007, January). Enhanced heavy-oil recovery by alkali-surfactant flooding. In SPE Annual
Technical Conference and Exhibition. Society of Petroleum Engineers.
Chen, Q. (2009). Assessing and improving steam-assisted gravity drainage: Reservoir heterogeneities, hydraulic fractures,
and mobility control foams.
Chen, Q., Gerritsen, M. G., & Kovscek, A. R. (2010, January). Improving Steam-Assisted Gravity Drainage Using
Mobility Control Foams: Foam Assisted-SAGD (FA-SAGD). In SPE Improved Oil Recovery Symposium. Society
of Petroleum Engineers.
Cheng, L., Reme, A. B., Shan, D., Coombe, D. A., & Rossen, W. R. (2000, January). Simulating foam processes at high
and low foam qualities. InSPE/DOE Improved Oil Recovery Symposium. Society of Petroleum Engineers.
Ezeuko, C., Wang, J. Y., & Gates, I. (2012, June). Investigation of Emulsion Flow in SAGD and ES-SAGD. In SPE Heavy
Oil Conference, Calgary, Alberta, Canada (pp. 12-14).
Falls, A. H., Hirasaki, G. J., Patzek, T. E. A., Gauglitz, D. A., Miller, D. D., & Ratulowski, T. (1988). Development of
a mechanistic foam simulator: the population balance and generation by snap-off. SPE reservoir engineering, 3(03),
884–892.
Farajzadeh, R., Andrianov, A., Krastev, R., Rossen, W. R., & Hirasaki, G. J. (2012, June). Foam-Oil Interaction in
Porous Media-Implications for Foam-assisted Enhanced Oil Recovery (SPE 154197). In 74th EAGE Conference and
Exhibition incorporating EUROPEC 2012.
French, T. R., Broz, J. S., Lorenz, P. B., & Bertus, K. M. (1986, January). Use of emulsions for mobility control during
steamflooding. In SPE California Regional Meeting. Society of Petroleum Engineers.
Friedmann, F., Chen, W. H., & Gauglitz, P. A. (1991). Experimental and simulation study of high-temperature foam
displacement in porous media. SPE reservoir engineering, 6(01), 37–45.
Gauglitz, P. A., Friedmann, F., Kam, S. I., & Rossen, W. R. (2002). Foam generation in homogeneous porous media.
Chemical Engineering Science, 57(19), 4037–4052.
Gotawala, D. R., & Gates, I. D. (2008). Steam fingering at the edge of a steam chamber in a heavy oil reservoir. The
Canadian Journal of Chemical Engineering, 86(6), 1011–1022.
Griffin, W. C. (1946). Classification of surface-active agents by" HLB". J Soc Cosmetic Chemists, 1, 311–326.
Hirasaki, G. J. (1989). A Review of Steam-Foam Process Mechanisms. paper SPE, 1951 8.
Islam, M. R., & Farouq Ali, S. M. (1994). Numerical simulation of emulsion flow through porous media. Journal of
Canadian Petroleum Technology, 33(3), 59–63.
Kam, S. I., & Rossen, W. R. (2003). A model for foam generation in homogeneous media. SPE Journal, 8(04), 417–425.
Kanicky, J. R., Lopez-Montilla, J. C., Pandey, S., & Shah, D. (2001). Surface chemistry in the petroleum industry.
Handbook of applied surface and colloid chemistry, 1, 251–267.
Kular, G. S., Lowe, K., & Coombe, D. (1989, January). Foam application in an oil sands steamflood process. In SPE
Annual Technical Conference and Exhibition. Society of Petroleum Engineers.
Lashgari, H., Lotfollahi, M., Delshad, M., Sepehrnoori, K., & De Rouffignac, E. P. (2014, June). Steam-Surfactant-Foam
Modeling in Heavy Oil Reservoirs. InSPE Heavy Oil Conference-Canada. Society of Petroleum Engineers.
Li, B., Hirasaki, G. J., & Miller, C. A. (2006, January). Upscaling of foam mobility control to three dimensions. In SPE/
DOE Symposium on Improved Oil Recovery. Society of Petroleum Engineers.
Murtaza, M., He, Z., & Dehghanpour, H. (2014). An approach to model three-phase flow coupling during steam chamber
rise. The Canadian Journal of Chemical Engineering, 92(6), 1100–1112.
Oskouei, P., Javad, S., Maini, B. B., Moore, R. G., & Mehta, S. A. (2010, January). Effect of Mobile Water-Saturation on
Thermal Efficiency of Steam-Assisted Gravity-Drainage Process. In SPE Latin American and Caribbean Petroleum
Engineering Conference. Society of Petroleum Engineers.
Patzek, T. W. (1996). Field applications of steam foam for mobility improvement and profile control. SPE Reservoir
Engineering, 11(02), 79–86.
Patzek, T. W., & Myhill, N. A. (1989, January). Simulation of the bishop steam foam pilot. In SPE California Regional
Meeting. Society of Petroleum Engineers.
16 SPE-185015-MS

Reme, A. B. (1999). Parameter Fitting and Calibration Study with a Commercial Foam Simulator (Doctoral dissertation,
Thesis, Norwegian university of Science and Technology).
Rossen, W. R. (2013, September). Numerical challenges in foam simulation: a review. In SPE annual technical conference
and exhibition (SPE 166232), New Orleans.
Schramm, L. L. (1992). Emulsions: fundamentals and applications in the petroleum industry.
Schramm, L. L. (1994). Foams: fundamentals and applications in the petroleum industry (Vol. 242). Washington, DC:
American Chemical Society.
Schramm, L. L. (Ed.). (2000). Surfactants: fundamentals and applications in the petroleum industry. Cambridge University
Press.
Sepehrnoori, K., & Johns, R. T. (2003). Theoretical and experimental study of foam for enhanced oil recovery and acid
diversion.
Sharma, J., & Gates, I. D. (2010). Multiphase flow at the edge of a steam chamber. The Canadian Journal of Chemical
Engineering, 88(3), 312–321.
Srivastava, M. (2010). Foam assisted low interfacial tension enhanced oil recovery.
Wang, J. (2008). Physical and Chemical Phenomena in Chemical Flooding.
Wang, J., & Dong, M. (2010). Simulation of O/W emulsion flow in alkaline/surfactant flood for heavy oil recovery.
Journal of Canadian Petroleum Technology, 49(6), 46.
Zhou, Z. H., & Rossen, W. R. (1995). Applying fractional-flow theory to foam processes at the'limiting capillary pressure'.
SPE Adv. Technol, 3(1), 154–162.

You might also like