Está en la página 1de 12

Understanding the Heat-Transfer

Mechanism in the Steam-Assisted Gravity-


Drainage (SAGD) Process and Comparing
the Conduction and Convection Flux in
Bitumen Reservoirs
Mazda Irani,* University of Alberta, and Sahar Ghannadi, University of Alberta

Summary outward, losing its latent heat when it comes into contact with the
SAGD is one successful thermal recovery technique applied in cold bitumen at the edge of a depletion chamber. As a conse-
the Athabasca and Peace River reservoirs in central and northern quence, the viscosity of bitumen falls several orders of magnitude,
Alberta, Canada. In SAGD, steam is injected into a horizontal and the bitumen flows under gravity toward a horizontal produc-
injection well and is forced outward, losing its latent heat when it tion well located several meters below and parallel to the injection
comes into contact with the cold bitumen at the edge of a deple- well (i.e., 5 m, but drilling tolerances often leave variations
tion chamber. As a consequence, the viscosity of the bitumen falls between 3 and 7 m). As the oil flows away and is produced, the
several orders of magnitude, its mobility rises several orders of steam chamber expands both upward and sideways (see Sections
magnitude, and then it flows under gravity toward a horizontal B and C in Fig. 1a). A cross-section of the SAGD process is dis-
production well located several meters below and parallel to the played in Fig. 1a. Section A shows the circulation stage, Section
injection well. Heat-transfer mechanisms are pivotal to the SAGD B presents the early phase in which the chamber is not well devel-
process. Though heat energy is transferred from steam to reservoir oped, and Section C presents the mature steam chamber in the
by conduction and convection, heat transfer by convection is not injection phase.
considered in the classic SAGD mathematical models such as The term steam-assisted gravity drainage was first devel-
Butlers. Researchers such as Butler and Stephens (1981), Reis oped by Roger Butler and his colleagues at Canadas Imperial Oil
(1992), Akin (2005), Liang (2005), Nukhaev et al. (2006), and in the late 1970s (Al-Bahlani and Babadagli 2009). Butler and
Azad and Chalaturnyk (2010) considered conduction from steam Stephens (1981) proposed the first closed-form solution for the
to cold reservoir to be the only heat-transfer component. How- prediction of oil production rate in the SAGD process. In his
ever, because the heat capacity of water is typically two to five model, known as the Butler theory, Butler described the SAGD
times that of bitumen, convection caused by the mobile conden- process as when steam is injected, a steam-saturated zone, called
sate flow in the reservoir may contradict these studies. Farouq-Ali a steam-depletion chamber or simply a steam chamber is
(1997) was the first to criticize the assumption that there is only a formed, in which the temperature is that of the injected steam
thermal conduction mechanism in the SAGD process. He pointed (Tchamber Tsteam). The steam flows toward the interface of the
out that with so much condensate flowing, convection would be steam chamber, where it condenses and loses its latent heat by
expected to be the dominant heat-transfer mechanism, which can flashing to bitumen. The latent heat from steam is transferred by
be plausible at high temperatures. In response, Edmunds (1999a) thermal conduction into the surrounding reservoir and mobilizes
stated that on the basis of the associated change in enthalpy, the the bitumen. The steam condensate and mobile bitumen flow by
heat transfer into a cold reservoir because of convection is prob- gravity to the production well located below the injector from
ably less than 5% of that because of conduction. Ito (1999) chal- side-drained paths (see Section F in Fig. 1b).
lenged Edmunds (1999a) statement, on the basis of Ito and Suzuki The physics differs at the top and sides of the steam chamber.
(1996, 1999) and Ito et al. (1998), pointing out that this number, The top of the chamber rises because of steam fingering (ceiling
5%; i.e., ratio between convection to conduction presented by drainage) (Gotawala and Gates 2008), and the mobilized bitumen
Edmunds (1999a) is unrealistically low, (and) it should be in the falls from the sides of the steam finger, where steam rise usually
range of 50%. This study examined the relative roles of convec- impedes liquid drainage. Along the sides of the chamber, bitumen
tive and conductive heat transfer at the edge of SAGD steam mobility is controlled by both conduction and convection heat
chambers. In summary, the mathematical model developed in this transfer on the interface, and thus, condensate and mobilized bitu-
study considered both conduction and convection, and the result- men are drained by gravity drive (slope drainage) (Edmunds et al.
ant output from the model is reasonably consistent with published 1989, 1994; Nasr et al. 2000).
field data. This study supports the idea that although convection Further research includes that of Reis (1992), who modified
can dominate near the chamber edge in high-water-saturation res- Butlers model by revising the steam chamber configuration from
ervoirs, in bitumen-rich reservoirs, its contribution to heat transfer an S-curve shape suggested by Butler to a triangular shape (linear
is less than 1% and can be neglected. geometry). Akin (2005) modified Butlers model by including the
effects of steam distillation (hence, asphaltene deposition) and
Introduction considering viscosity reduction and production rate increase as
results of deasphalting. Liang (2005) presented an analytical model
SAGD is one successful thermal recovery technique applied in for cyclic steaming. Nukhaev et al. (2006) modified Butlers model
the Athabasca and Peace River reservoirs in Alberta, Canada. In to consider real-time production rate variation. Azad and Chalatur-
SAGD, steam injected into a horizontal injection well is forced nyk (2010) modified the work of Reis (1992) to suggest a semiana-
lytical model for considering geomechanical effects with a limit
*
Now with RPS Energy equilibrium model. They named this the model of slices.
Copyright V
C 2013 Society of Petroleum Engineers Though the SAGD process looks simple, there are some theo-
Original SPE manuscript received for review 23 January 2012. Revised manuscript received
retical pitfalls. All of the analytical models pertain to the flow of a
for review 13 August 2012. Paper (SPE 163079) peer approved 21 August 2012. single fluid (Farouq-Ali 1997). They assume constant steam

134 February 2013 SPE Journal


Caprock

A B C

Vertical Drained Flow

Steam Temperature = Tst

Side Drained Flow

Initial Reservoir Temperature = Tr

a. Cross section of SAGD process; Section A presents circulation phase, Section B presents early phase, and Section C presents
steam injection phase.

Caprock

D E F Steam Temperature = Tst

Drained Flow

Initial Reservoir Temperature = Tr

b. Cross section of s-shaped SAGD process in Butler model and this study; Section D presents circulation phase, Section E
presents early phase, and Section F presents steam injection phase.

Fig. 1Cross-section of SAGD process; Section A represents a real behavior, and Section B represents a behavior modeled in the
Butler model and this study. Cross section of SAGD process; Section A presents circulation phase, Section B presents early
phase, and Section C presents steam injection phase. Cross section of S-shaped SAGD process in Butler model and this study;
Section D presents circulation phase, Section E presents early phase, and Section F presents steam injection phase.

pressure inside the chamber (Farouq-Ali 1997), with no steam flow that can be done to raise it. In addition, surface area is not control-
outside of it (Farouq-Ali 1997); constant oil saturation equal to the lable in the SAGD process. Because of these limitations in con-
residual inside the steam chamber (Farouq-Ali 1997); and heat duction heat transfer, convective heat transfer and controlling
transfer ahead of the steam chamber by conduction only (Farouq- convection can enhance the thermal flux rate at the edge of steam
Ali 1997). None of these models include geomechanics in closed- chambers. Despite this, it remains unclear how large convection is
form solution. [It must be noted that the work of Azad and Chala- relative to conduction. Farouq-Ali (1997) criticized the idea that
turnyk (2010) must be programmed, and cannot be considered a only thermal conduction exists in SAGD, pointing out that with
closed-form mathematical solution.] Most analytical models deal so much condensate flowing, convection would be expected to be
with slope drainage, and little (aside from Das and Butler 1996) has the dominant heat-transfer mechanism; this is plausible at high
been published dealing with the mechanisms of drainage at the temperatures. Farouq-Ali (1997) also supported his idea by refer-
ceiling of the steam chamber (Al-Bahlani and Babadagli 2009). ring to Ito and Suzukis (1996) numerical simulations, which
A key control on production rate in the SAGD process is the show that convection is more important than conduction. This
heating of oil sands at the edge of the steam chamber (Sharma convection is associated with condensate carrying heat into the
and Gates 2011a); the higher the flux rate, the hotter and mobile oil zone. In response to Farouq-Alis (1997) critique,
more mobile the bitumen. In analytical models presented by most Edmunds (1999a) stated that on the basis of the associated change
researchers (e.g., Butler and Stephens 1981; Butler 1985, 1994b; in enthalpy, 22% of total heat transfer is carried by convection
Ferguson and Butler 1988; Reis 1992, 1993; Akin 2005; Liang (i.e., 18% by liquid water and 4% by mobile bitumen). Edmunds
2005; Nukhaev et al. 2006; Gotawala and Gates 2008; Azad and concluded that because of the water streamline being nearly per-
Chalaturnyk 2010), heat conduction is the major mode of heat pendicular to the temperature gradient (and almost parallel to iso-
transfer between the edge of steam chambers and the outer cold therms), the convective heat flux perpendicular to the steam
bitumen. On the basis of Sharma and Gates (2011a), conduction chamber edge must be reduced by the sine of the angle between
heat transfer at the edge of steam chamber can be enhanced by the streamlines and the isotherms. That is, if the condensate flows
raising steam chamber temperature, increasing the thermal con- exactly along the isotherms, there is zero net convection. He
ductivity of oil sands, and increasing heat-transfer surface area. finally concluded that heat transfer into cold reservoirs caused by
Because steam chamber temperature is a function of steam pres- convection is probably less than 5% of that caused by conduction.
sure (because steam chamber is at saturation condition), it is lim- Edmunds (2000) also emphasized that heat transfer in the mobile
ited by the range of practical steam pressure. The thermal zone is dominated by conduction, except very near the liner or
conductivity of cold oil sands is nearly constant, and there is little anywhere that live steam penetrates. Ito (1999) challenged the

February 2013 SPE Journal 135


Te In simpler form, this is presented as:
mp
Co er
atu  2   
nv re @ T @2T @2T @T y @T z @T
Co ec K  q c c pc V c V c V c
nd tio @x2 @y2 @z2 @x @y @z
uc n  
tio T @T
n st qr cpr                           2
Te @t
mp
er
atu where Vc , Vcy , and Vcz are Darcys velocities in the x-, y- and z-
re
T Di direction (i.e., x-direction is the direction normal to the steam
r str
ibu chamber interface, y-direction is the direction parallel to the steam
Di tio
Steam Temperature = Tst sta n chamber interface in steam chamber plane, and z-direction is par-
Co nc
nv e allel to injector and producer wells), respectively; qc is condensate
ec
tio density; cpc is condensate heat capacity; qr is reservoir density
n
(i.e., immobile bitumen outside the steam chamber); cpr is reser-
Co voir heat capacity; and K is reservoir thermal conductivity.
nd
uc Because in commercial practices injection- and production-
tio
Condensate + Bitumen Flow n well lengths are typically in the range of 500 to 1500 m (Das
2005; Sharma and Gates 2011a) (e.g., Shell wells in Peace River
Initial Reservoir Temperature = Tr are 1500 m in horizontal length, and AOSTRA underground test
facility (UTF) Phase B wells are 500 m), in homogeneous reser-
voirs, the temperature gradient parallel to the well trajectory is
Fig. 2Illustration of conduction and convection heat transfer very small and can be neglected. It must be noted that in many
at the edge of a steam chamber. oil-sand reservoirs, this assumption needs to be explored. In heter-
ogeneous reservoirs and in fields in which injection pressure along
an injector varies rapidly as a result of frictional pressure drop in
Edmunds (1999a) statement, on the basis of Ito and Suzuki (1996, the injector annulus (McCormack 2002), the temperature gradient
1999) and Ito et al. (1998), pointing out that this number, 5%; parallel to the well trajectory can be significant and needs to be
i.e., ratio between convection to conduction presented by considered. In these cases, the barbell-shaped steam chamber is
Edmunds (1999a); is unrealistically low, (and) it should be in the created, causing thermal gradient along the well to be significant
range of 50%. Ito (1999) roughly calculated this ratio to be 56% (Shaw and Bedry 2012). This assumption is valid in homogeneous
for steam injection with injected temperature equal to 260 C. reservoirs with the steam chamber conformance control along the
Edmunds (1999b) argued that because of that simulation imple- horizontal well such as Orion field SAGD well pairs that incorpo-
mented grid size, the analysis should not focus on convective flux rate interval control valves (ICVs) (Clark et al. 2010). Also be-
near the steam interface. He also stated that the grids were not cause of normal chamber expansion, the temperature gradient
aligned with flow streamlines, and that the vertical flux through a parallel to a steam chamber edge is very small in comparison with
horizontal gridblock boundary was not to be interpreted as a local that at the normal chamber edge, and can be neglected. The heat-
vertical flow. Edmunds noted that the maximum penetration depth transfer physics presented in Eqs. 1 and 2 can be reduced to the
was small compared with the length of the steam interface, and following equation assuming one-direction heat transfer for the
that the streamline angle in the SAGD front was not clear. direction normal to the chamber interface:
Edmunds (1999b) stated that Butlers model may agree with sim-      
ulators because the convective heat transfer may balance out the @2T @T @T
K  V q c
c c pc q c
r pr . . . . . . . . . . 3
reduced relative permeability of oil because of water. He also @x2 @x @t
added that the fudge factor is in the field-data-adjusted constant in
the Butler model. Jimenez (2008) studied the nonlinear compres- where Vc is the condensate convective velocity normal to the
sibility and geomechanical behavior of oil sands and concluded steam chamber edge and x is the normal distance to the advancing
that at low pressure, SAGD heat transfer from the steam chamber front of the steam chamber. This assumption cannot be valid in
into the formation relies mainly on conduction because mean specific geologies or in cases with profound wellbore effect (i.e.,
effective stress remains unchanged or increases because of ther- when pressure drops along the well as a result of turbulent flow or
mal stresses. He concluded that at high pressure, SAGD heat large distance between wells) (Farouq-Ali 1997).
transfer has a component of convection caused by compressibility Butler (1991) solved this equation by eliminating the convec-
because the mean effective stress becomes lower. This is because tion part, with the chamber interface moving at Ux velocity (see
the pressure front travels ahead of the thermal front. Sharma and Fig. 3). Under these conditions, the x can be transformed to the
Gates (2011a) present the only closed-form solution for consider- moving distance measured from the chamber interface, rather
ing convection mechanisms from the edge of steam chambers into than from a fixed origin. This is done using the new variable (n)
reservoirs. They modified the Butler theory to consider convec- defined as
tion heat transfer, concluding that convection at the chambers
edge dominates above approximately 225 C, whereas conduction t
is the major mechanism below 125 C. This study examines the nx Ux dt x  Ux t . . . . . . . . . . . . . . . . . . . . 4
0
relative roles of convective and conductive heat transfer at the
SAGD steam chamber edge. Controlling factors in convective
heat transfer are also studied. The x value can be replaced with n, resulting in
   
@T @T
. . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
Heat-Transfer Theory Considering Convection @x @n
The schematic of the conduction and convection heat transfer  2   2 
@ T @ T
from the edge of the steam chamber into the reservoir is illustrated . . . . . . . . . . . . . . . . . . . . . . . . . . 6
in Fig. 2. According to Carslaw and Jaeger (1959), heat transfer @x2 @n2
ahead of the steam chamber occurs by convection and conduction          
@T @ T dt @ T dn @T @T
according to the following equation:  Ux
@t x @t n dt @n t dt @t n @n t
~  rT qr cpr T_
K r2 T  qc cpc V . . . . . . . . . . . . . 1                    7

136 February 2013 SPE Journal


the interface [basically, the relative permeability of water at residual

e
tur
oil saturation (ROS); on the basis of the Butler model, the oil satura-

era
T tion inside the chamber at its residual]; and lint
st Te w is viscosity of water

mp
Di mpe at the steam interface [at steam temperature (on the basis of the But-

Te
str
ibu ratu ler model, the chamber including the interface temperature is equal
tio re to the steam temperature). Sharma and Gates (2011a) explained that
n
T their correlation is indirectly a function for the pressure difference
r
Steam Temperature = Tst Di required for convective flow, because given an original reservoir
sta
nc pressure, the higher the steam pressure is, the greater the corre-
e
Ux sponding steam temperature is, the lower the bitumen viscosity is,
and, consequently, the higher the oil rate and Ux are.
On the basis of Coreys relative permeability correlation, the
relative permeability of the water inside the reservoir can be cal-
Drained Flow
culated using the following equation:
 r n
res ro Sw  Swc
Initial Reservoir Temperature = Tr krw krw . . . . . . . . . . . . . . . . . . . 12
1  Swc  Sor

where n (the Corey coefficient) sets the curvature of the relative


Fig. 3Illustration of lateral advance of steam front modified
permeability curves, Srw is water (or condensate) saturation in the
from Butler (1991).
reservoir ahead of the front, Swc is connate-water saturation in the
reservoir, and Sor is ROS in the reservoir.
Substitution of Eqs. 5, 6, and 7 into Eq. 3 yields Sharma and Gates (2011a) assumed linear distribution of oil
 2        saturation beyond the steam chamber edge and confirmed this
@ T @T @T @T using simulation results from Ito and Suzuki (1996). Sharma and
K  V c q c c pc q r c pr  U x
@n2 @n @t @n Gates (2011a) suggested presenting reservoir oil saturation ahead
                   8 of the front in the following equation:
 
and then T  Tr
Sro Sor Sio  Sor 1  . . . . . . . . . . . . 13
 2      Tst  Tr
@ T @T @T
K 2
 Vc qc cpc  Ux qr cpr qr cpr : where Sio is initial oil saturation in the reservoir and Sor is ROS in
@n @n @t
the reservoir. On the basis of Butler (1997), the temperature distri-
                   9 bution ahead of the steam chamber is presented in the following
equation:
Following Butler (1985, 1991) for the quasisteady-state stage
(where heat ahead of the front is run over by the advancing  
T  Tr Ux
steam chamber front and the temperature at distance n from the exp  n . . . . . . . . . . . . . . . . . . . . . 14
Tst  Tr a
front is constant with time), the right-hand side of Eq. 8 can be set
equal to zero, and the equation becomes Finally, Eq. 11 can be presented as
 2     
@ T @T kro =lres Ux
K  V q c
c c pc  U q c
x r pr 0 . . . . . . . . 10 Vc Ux rw w
exp  n n . . . . . . . . . . . . . . . . 15
@n2 @n int =lint
krw w a

Condensate Velocity Calculation The problem with Model 1 relates to condensate velocity eval-
uation, which is the main controller of convective heat transfer.
Because water mobility is substantially higher than that of bitumen
The Model 1 presented by Sharma and Gates (2011a) deals with
[the mobility ratio between water and bitumen is greater than 1,000,
the moving velocity (Ux) of a steam chamber interface as a moving
which means water mobility is three orders of magnitude greater
boundary problem, such as water and gas sweeping problems. The
than bitumen mobility (this will be discussed in detail in the begin-
steam chamber interface moving velocity (Ux) is mainly controlled
ning of the Model 3 subsection] and water condensate also flows at
by the rate of exposing a new oil sands element to steam hot inter-
higher rates than mobile oil on the surface edge [steam/oil ratio
face (ablation), which is, in turn, controlled by thickness and flow
(SOR) in SAGD projects are mostly above 3], condensate steam is
rate of the bitumen/condensate layer. The other debate about this
the major component of convection and mobile bitumen is less im-
model regards linear distribution of oil saturation beyond the
portant. It must be noted that the part of condensate velocity that is
steam chamber edge, which is not supported by any analytical
parallel to the temperature gradient (perpendicular to isotherms) is
work. Inconsistency is the main disadvantage of this model:
that responsible for convective heat transfer. Condensate velocity
Although Eq. 10 must be solved to evaluate temperature variation,
normal to the steam chamber edge (Vc) is the major controller and
in oil saturation calculations, the Butler model temperature varia-
only unknown in the convective heat-transfer evaluation. Conden-
tions were used, though these are calculated on the basis of con-
sate density and condensate heat capacity are known and nearly
ductive heat transfer only.
constant for the duration of SAGD processes. In this section, three
models are presented for condensate velocity evaluation.
Model 2. The second model is implemented in this study on the
basis of heat energy conservation. Heat conservation at the steam
Model 1. The first model is implemented by Sharma and Gates chamber interface is formulated as
(2011a). They suggest the following equation for evaluation of
condensate velocity normal to the steam chamber edge (Vc): Conductive Heat Flux Convective Heat Flux
kres =lres Total Heat Transferred Into Cold Bitumen      16
Vc Ux rw w
int =lint
. . . . . . . . . . . . . . . . . . . . . . . . . . 11
krw w
where
res
where krw is relative permeability of water (condensate) in the res-
ervoir, which is at connate water saturation; lres @T @T
w is viscosity of water Conductive Heat Flux K K . . . . . . . . . 17
int
(condensate) in the reservoir; krw is relative permeability of water at @x @n

February 2013 SPE Journal 137


there is a great difference in the viscosities of water and bitumen
under initial reservoir conditions. For example, water viscosity
Ux under reservoir conditions is approximately 1 cp, whereas bitu-
men viscosity in the Athabasca deposit is greater than 106 cp [i.e.,
Tr in the range of 2 to 5  106 cp (Collins 2005)]. Because the bitu-
Drained Bitumen men is at nearly its maximum saturation (i.e., water connate satu-
ration is nearly at irreducible saturation), its relative permeability
Tst Native Bitumen is roughly equal to unity. Meanwhile, the water is nearly at irre-
ducible saturation and, because of coreflood relative permeability
experiments (Bennion et al. 2006), the relative permeability of
water is low (i.e., ranging between 0.001 and 0.01). The following
Drained Flow equation shows that the water mobility is substantially higher than
that of the bitumen:
Fig. 4Illustration of stored heat in bitumen from initial case in  
reservoir to drained case in interface. kro 1
6 106
lo 10 cp
Convective Heat Flux Vc qc cpc Tst  Tr . . . . . . . . . 18  
krw 0:001 to 0:01
<< 103 to 102        25
lw 1 cp
As shown in Fig. 4, the stored heat in a unit volume of bitumen
from its initial case in the reservoir condition [i.e., with tempera- As a result, the bitumen is largely immobile, and the water is
ture equal to reservoir temperature (Tr)] to the mobilized bitumen the only mobile phase, behaving as a single phase within a reser-
in drained [i.e., with temperature equal to steam temperature (Tst)] voir medium; this is supported by field evidence. Aherne and
is equal to Maini (2006) show that water moves horizontally through the bi-
tumen zone by interpreting the pressure data from piezometers
Stored Heat in Unit Volume of Bitumen
located in the bitumen zone in three nearby observation wells.
qr cpr Tst  Tr . . . . . . . . . . . . . . . . . . . . . . . . 19 This was achieved during a cold-water injectivity test performed
at the Underground Test Facility (UTF) Phase A SAGD pilot.
Considering the moving velocity of the steam chamber inter- This implies that connate water saturation is higher than irreduci-
face (Ux), the rate of reservoir heating and bitumen mobilizing ble water saturation.
can be evaluated in the following equation: The water is mobile ahead of the steam chamber front, and
there is a large pressure gradient normal to the steam chamber
Total Heat Transfered Ux qr cpr Tst  Tr . . . . . . . . . 20 interface because of high steam pressure inside the chamber and
comparatively low reservoir pressure. This pressure difference
On the basis of Butler (1997), the temperature variation ahead across the chamber interface drives steam condensate into the oil
of the steam chamber can be calculated as sands ahead of the chamber edge, and is the major controller of
  the convective heat transfer.
T  Tr Ux The single-phase pressure variation normal to the chamber
exp  n . . . . . . . . . . . . . . . . . . . . 21
Tst  Tr a interface can be formulated as follows:
 2   
By differentiating Eq. 21, the conductive heat transfer (Eq. 17) @ P /lc @P
can be presented as follows: . . . . . . . . . . . . . . . . . . . . . . 26
@x2 kkr @t
@T where P is fluid pressure at x, which is the distance normal to the
Conductive Heat Transfer K
@n advancing front of the steam chamber; / is porosity of the reser-
 
Ux Ux voir; l is viscosity of the fluid; c is compressibility of the oil
K Tst  Tr exp  n              22 sand; k is absolute permeability of the reservoir; and kr is relative
a a
permeability of the fluid in the reservoir. Because the water is the
For the steam chamber interface, setting n equal to zero yields only mobile phase and behaves as a single phase within a reser-
the following for conductive heat transfer at the edge: voir medium, the condensate-velocity profile inside the reservoir
depends on the waters relative permeability. Its viscosity varia-
Conductive Heat Transfer at Chamber Edge tions normal to the chamber interface are independent of the bitu-
Ux mens relative permeability and viscosity, and Eq. 26 can be
K Tst  Tr . . . . . . . . . . . . . . . . . . . . . . . . 23 presented in the following format:
a
 2   
The total equation can be presented as @ P /lw c @P
. . . . . . . . . . . . . . . . . . . . . 27
@x2 kkrw @t
Ux
K Tst  Tr Vc qc cpc Tst  Tr Ux qr cpr Tst  Tr
a where P is condensate pressure at x, which is the distance normal to
                   24 the advancing front of the steam chamber; lw is viscosity of the con-
densate (water); c is compressibility of the oil sand; and krw is rela-
When a is replaced with K/(qr cpr), the condensate velocity tive permeability of the condensate in the reservoir. The solution is
normal to the steam chamber edge (Vc) equals zero. This shows that used by Butler (1985, 1994b) for conductive heat transfer if the
we cannot calculate condensate velocity considering heat energy steam-chamber interface is moving with velocity Ux normal to the
conservation by using any model-evaluated temperatures that do steam chamber edge (see Fig. 5). The quasisteady-state solution for
not consider convection, such as the Butler model (Butler and Ste- pressure variation ahead of the chamber is presented as
phens 1981). In these cases, condensate velocity must be eval-  
uated beforehand. P  Pr /lw c
P exp  Ux n . . . . . . . . . . . . . 28
Pst  Pr kkrw

Model 3. If water saturation is greater than irreducible water sat- where n is distance normal to the advancing front of the steam
uration, water mobility in Athabasca and Peace River bitumen chamber at any time. The condensate velocity normal to the steam
reservoirs is many times greater than that of bitumen because chamber interface is formulated as

138 February 2013 SPE Journal


(A)

Evaluated Temperature, C
ur
250
P

s
es
st
Pr

Pr
Di essu 200
str
ibu re
tio 150 krw = 0.05
n krw = 0.02
P
r 100
Di krw = 0.01
Steam Pressure = Pst sta
nc 50
e Butlers
Co Model
0
nv 0 2.5 5 7.5 10 12.5 15 17.5 20
ec
tio Distance Normal to Interface, m
Vc nD (B)
ire

Temperature Difference, C
cti 1E+2
on
1E+1
1
Initial Reservoir Pressure = Pr 1E1
1E2
1E3
Fig. 5Illustration of pressure distribution vs. advance of

krw = 0.001
krw= 0.0001

krw = 0.01

krw = 0.1
steam front. 1E4
1E5
1E6
kkrw @P 0 2.5 5 7.5 10 12.5 15 17.5 20
Vc  . . . . . . . . . . . . . . . . . . . . . . . . . . 29 Distance Normal to Interface, m
lw @n
By differentiating Eq. 28, the condensate velocity normal Fig. 6Comparison of temperature distribution estimated on
to the steam chamber interface (Eq. 29) can be presented as the basis of Model 3 for different relative permeabilities. Section
follows: B shows temperature difference error caused by neglecting the
  convective term. See Table 1 for fluid and reservoir properties.
kkrw @P /l c
Vc  Ux / c Pst  Pr exp  w Ux n
lw @n kkrw the temperature profile with convection included) vs. Eq. 21 (i.e.,
Ux / c P  Pr                        30 the temperature profile that is based on the Butler model excluding
convection) is illustrated in Section A of Fig. 6; here, the differ-
For the steam chamber interface we can set n equal to zero and ence between Eq. 33 and Eq. 21 is shown for the possible range of
formulate the condensate velocity normal to the steam chamber reservoir properties. Section B of Fig. 6 shows the error in temper-
interface at the edge: ature profile including the condensate convection (i.e., using Eq.
33) and the Butler model (i.e., using Eq. 21) for several orders of
Vc Ux / c Pst  Pr . . . . . . . . . . . . . . . . . . . . . . 31
magnitude of the relative permeability (i.e., from 0.0001 to 0.1).
For solving the heat transfer, the Vc is substituted by its value As shown in Section B of Fig. 6, the temperature difference,
from Eq. 30: including convection, is only noticeable for relative permeabilities
 2     larger than 0.01. Accordingly, for the range of reservoir properties
@ T /lw c (i.e., relative permeabilities from 0.0001 to 0.01), the difference
K  U x / c P st  P r exp  U x n qc cpc between Eqs. 21 (Butler model) and 33 is negligible.
@n2 kkrw
 
@T
Ux qr cpr 0             32 Conductive and Convective Heat-Transfer
@n
Calculation for Different Presented Models
On the basis of Appendix A, the solution for Eq. 32 is sug- The main focus of this study is to compare conductive and con-
gested in Eq. A-19, as follows: vective heat transfer. For this purpose the conductive and convec-
  tive heat-transfer equations will be presented for different models.
Ux
n
X1
expj n ja
gn Model 1. Because Sharma and Gates (2011a) do not present the
T  Tr n 0
n!  Ux =ja n
T ( ) solution for temperature variation, Butlers model is used for tem-
Tst  Tr ja X
1
1 perature variation in the conductive heat-transfer calculation:
gn
Ux n 1
n!  Ux =ja n @T Ux
  Conductive Heat Flux K K Tst  Tr
Ux @n a
X
1 ja n  
P Ux Ux
gn  exp  n K T  Tr           36
n 0
n!  Ux =ja n a a
( )
ja X1
1       33 The convective heat transfer can be calculated using the conden-
gn
Ux n 1 n!  Ux =ja n sate velocity normal to the steam chamber edge (Vc) from Eq. 15:

where, Convective Heat Flux Vc qc cpc T  Tr


ro
 
Ux Pst  Pr qc cpc kkrw Ux Pst  Pr kkrw krw =lres
w Ux
g . . . . . 34 Ux int int exp  n n qc cpc T  Tr
K lw ac lw krw =lw a
ro
 n
/lw c krw =lres T  T r
j Ux . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35 Ux int wint qc cpc T  Tr
kkrw krw =lw Tst  Tr
  n
kro =lres 1
where ac is the simplified condensate (water) diffusivity known in Ux rw
int
w
int
qc cpc T  Tr n1        37
this study as convective diffusivity. The variation of Eq. 33 (i.e., krw =lw Tst  Tr

February 2013 SPE Journal 139


@T
TABLE 1DOVER UTF PHASE B FLUID AND RESERVOIR Conductive Heat Flux K
@n
PROPERTIES   
Ux Pst  Pr qc cpc kkrw /lw c
KC0 exp  exp  Ux n
Parameter Model 1 Model 3 K lw kkrw

Ux
Hydraulic Properties  n  Tst  Tr
/ 0.33 0.33 a
" #
kabs (m2) 6.0  1012 6.0  1012 Ux qc cpc kkrw Ux
int
krw 0.02 N/A K C0 exp  P  Pr  n  Tst  Tr
K lw a
ro
krw 0.02 N/A
krw N/A Variesa                    41
N 3.5 N/A
This equation gives answers identical to those of Eq. 36. The
lint
w (cp) 0.1127b N/A convective heat transfer can be calculated using the condensate
(0.1296c) velocity normal to the steam chamber edge (Vc) from Eq. 30.
lres
w (cp) Variesb N/A
lw (cp) N/A 0.1127b Convective Heat Flux Vc qc cpc T  Tr Ux / c Pst
 
Thermal Properties /lw c
a (m2/s) 7.0  107 7.0  107 Pr exp  Ux n qc cpc T  Tr
kkrw
K (W/m C) 1.45 1.45
Ux /cP  Pr qc cpc T  Tr    42
cpc (J/kg C) 4532.7d 4532.7d
qc (kg/m3) 858.5d 858.5d
The conductive and convective heat-transfer flux is a function of
Chamber Properties both pressure and temperature, and both are controllers in the con-
Ux (cm/d) 1.7 1.7 ductive and convective heat-transfer-flux calculation. Because the
Pr (MPa) N/A 1.47e conductive and convective heat-transfer-flux equations are corre-
Pst (MPa) N/A 1.73f lated to pressure and temperature, both pressure and temperature at
Tr ( C) 15 15 any distance n from the steam chamber front must be calculated, and
Tst ( C) 205 205 then the pressure related for each temperature must be substituted in
a
Evaluated on the basis of Bennion et al. (2006).
the conductive and convective heat-transfer-flux calculation.
b
Evaluated on the basis of Reid et al. (1977).
c
d
Evaluated on the basis of Butler (1991) using Schmidt and Grigull (1981) data. Results and Discussion
Evaluated on the basis of Butler (1991).
e
Hydrostatic pressure evaluated on the basis of an average depth of 150 m.
Table 1 lists the properties used to analyze the theory presented
f
Saturation pressure related to 205 C evaluated on the basis of Williamson (1972). in this study. These values are typical of those found in the litera-
ture for the Dover UTF Phase B reservoir in the Athabasca reser-
voir, and those used by Sharma and Gates (2011a), which mostly
referred to ORourke et al. (1994, 1999) and Birrell (2001). In this
Eq. 37 shows that the convective heat transfer is on the order of table, thermal conductivity is assumed to equal 1.45 W/m C. On
n1 of temperature. the basis of Butler (1991), this value is estimated for saturated
unconsolidated sandstone with 0.33 porosity. Comparison of pres-
sure distribution calculated for Model 3 (Eq. 28) for different rela-
Model 2. This model considers only conductive heat transfer, tive permeabilities is presented in Section A of Fig. 7. The pressure
neglecting convective heat transfer. The conductive heat transfer variation under reservoir conditions for low water relative perme-
is easily calculated using Eq. 36. abilities (i.e., in the range of 0.0010.0001) is present in the first 1
to 2 m. Section B of Fig. 7 presents the condensate velocity normal
Model 3. The solution for temperature variation is suggested in to the interface distribution calculated by the theory for Model 3
Appendix A as (Eq. 30) for different relative permeabilities. The results reveal that
for low water relative permeabilities, the condensate velocity nor-
   mal to the interface loses velocity rapidly in the first 1 to 2 m.
@T Ux Pst  Pr qc cpc kkrw /l c
C0 exp  exp  w Ux n Fig. 8 shows the conductive and convective components of the
@n K lw kkrw
 total heat flux for Models 1 and 3. These results were compared
Ux with field data obtained from vertical observation wells in the Dover
 n  Tst  Tr             38
a Phase B SAGD pilot presented by Birrell (2001). Birrell (2001) ana-
lyzed temperature data from vertical observation wells in the Dover
where Phase B SAGD pilot in order to evaluate the relative roles of con-
ductive and convective heat transfer. The results from Model 3 for
j water relative permeability of 0.01 agrees strongly with Birrells
C0 ( ) . . . . . 39
(2001) analyzed heat flux values, unlike Model 1, which always
ja X
1
1
gn gives linear results. Model 3 captures the downgrading feature of the
Ux n 1 n!  Ux =ja n conductive heat flux. The results reveal that Model 3s theory gener-
ates trends similar to the field data for both conductive and convec-
Substituting the pressure variation from Eq. 28, Eq. 38 yields tive mechanisms. This suggests that it captures the underlying
behavior of convection at the edge of a steam chamber.
 
@T Ux qc cpc kkrw Ux The water saturation assigned to each water relative perme-
C0 exp  P  Pr  n  Tst  Tr ability is also shown in Fig. 8 and is based on the work by Benn-
@n K lw a
ion et al. (2006) for both low and high temperatures. The
                   40 reservoir water saturation is under 20%, but the condensate near
the edge of the steam chamber can be three to five times the vol-
Then, on the basis of Eq. 40, and the substitution of tempera- ume of the flowing oil (Farouq-Ali 1997). Thus, the presence of
ture variation in conductive heat transfer, conductive heat transfer high water saturation near the chambers edge leads to an increase
can be calculated as in the relative water permeability, which, in turn, results in

140 February 2013 SPE Journal


(A) psi MPa
1.90
270

krw = 0.0001
1.85

krw = 0.001
krw = 0.01
krw = 0.1
krw = 0.5
260 1.80

1.75 Chamber Pressure


250

Pressure
1.70

240 1.65

1.60
230
1.55
220
1.50
Reservoir Pressure
210 0
0 2.5 5 7.5 10 12.5 15 17.5 20
Distance Normal to Interface, m

(B) cm/day nano-m/sec

120
Condensate Velocity Normal to Interface

104
krw = 0.0001
krw = 0.001
krw = 0.01
krw = 0.1
100 krw = 0.5

80

60
5 105

40

20
105
0 0
0 2.5 5 7.5 10 12.5 15 17.5 20
Distance Normal to Interface, m

Fig. 7Comparison of pressure distribution calculated by theory for Model 3 (Eq. 28) for different relative permeabilities in Section
A; and comparison of condensate velocity normal to interface distribution calculated by theory for Model 3 (Eq. 30) for different
relative permeabilities in Section B. See Table 1 for fluid and reservoir properties.

enhanced convective flow near the interface. This may be why range between 0.001 and 0.01. The latter value is within the
Birrells (2001) results for high temperatures (i.e., near the steam convection-active zone near the chamber interface. The tempera-
chamber interface) are identical to heat flux curves associated ture enhancement for water relative permeabilities ranging
with high water saturations. between 0.001 and 0.01 is plausibly displayed only in log scale (see
Convective flux in Model 1 is dominated by the relationship Section B in Fig. 6), and the error ranges between 0.2 and 0.002%.
between water dynamic-viscosity variation and temperature. These The results from Fig. 6 show that although convection can
are compared in Section B in Fig. 8 for two different correlations enhance the thermal flux rate at the edge of the steam chamber, it
(i.e., Reid et al. 1977; Butler 1991). Multiple correlations can be happens over very short distances. For practical Athabasca reser-
used, because water dynamic-viscosity variation is not very sensitive. voir parameters, its effect is lower than 1%; this supports the
The heat flux caused by convection, although at times large at work of Edmunds (1999a, b). However, it does not contradict Ito
the interface, happens over very short distances. Fig. 6 compares and Suzuki (1996, 1999) and Ito et al. (1998), which pointed out
temperature profiles beyond the edge of the steam chamber. In the that convective heating mainly occurs above 140 C but is not the
conduction-only case, these are calculated using Butlers model, dominant heat-transfer mechanism. Convective heat transfer
whereas the case that couples conduction and convection is calcu- becomes minor below 100 C.
lated on the basis of Model 3 (Eq. 33). The results reveal that in Overall, Edmunds (1999a, b) argument on convective heat
water relative permeabilities ranging between 0.001 and 0.0001, transfer is valid, and the assumptions behind Butlers theory can
the temperature difference between Model 3 and Butlers model be ignored, causing errors less than 1% (i.e., that heat is trans-
ranges between 0.004 and 0.30 C. This difference increases ferred from the steam chamber to the cold oil zone by thermal
rapidly [i.e., 31 C (36%) for krw of 0.05, 17 C (19%) for krw of conduction alone, and that no steam condensate flows ahead of
0.02, and 9 C (9%) for krw of 0.01 (see Section A in Fig. 6)]. the steam chamber).
Although there are significant differences for higher water relative It must be noted that the theory of Model 3 assumes a stable
permeabilities, this happens only over small distances. The water steam/bitumen front. Steam fingering or instability can accelerate
relative permeabilities at ROS are approximately 0.02, reaching convection by inducing steam fingers, which penetrate the cold bi-
0.05 in extreme cases, but under reservoir conditions the water tumen at the edge of the chamber and increase channeling flow. A
relative permeabilities beyond the edge of the steam chamber steam-finger length is one of main unknowns in this phenomenon.

February 2013 SPE Journal 141


(A) Model 1 (D) Model 3
80 80
UTF Field Data UTF Field Data
2

2
Conductive Heat Flux, W/m

Conductive Heat Flux, W/m


60 Model 1 60 Model 1
Conduction

Steam Temperature

Steam Temperature
40 40

krw = 0.05
krw = 0.01
krw = 0.005
krw = 0.001
krw = 0.0001
20 20

0 0
0 25 50 75 100 125 150 175 200 225 0 25 50 75 100 125 150 175 200 225
Temperature, C Temperature, C

50 100 150 200 250 300 350 400 50 100 150 200 250 300 350 400
Temperature, F
(B) (E) Temperature, F

100 160 100

40
60

50

30
UTF Field Data Sw , %
60-100 C
80 140 80 Model 1

40
85
65
50

30

20
2

2
Sw , %
Convective Heat Flux, W/m

Convective Heat Flux, W/m


var. w (Reid et al., 1977)

150-275 C
Convection

UTF Field Data

krw = 0.0001
var. w (Butler, 1991)

120

krw = 0.005
krw = 0.001
60 60

krw = 0.05
krw = 0.01
Model 1 Increasing initial reservoir
water saturation

Steam Temperature

Steam Temperature
40 100 40
const. w

20 80 20

0 60 0
0 25 50 75 100 125 150 175 200 225 0 25 50 75 100 125 150 175 200 225
Temperature, C Temperature, C

50 100 150 200 250 300 350 400 50 100 150 200 250 300 350 400
Temperature, F Temperature, F
(C) (F)
100 100
85
60 65
50 krw = 0.01
80 var. w (Butler, 1991) 80 50
40 krw = 0.005
2

var. w (Reid et al., 1977) 40


Total Heat Flux, W/m

30 krw = 0.001
2
Total Heat Flux, W/m

60 60 30 20 krw = 0.0001
const. w
Total

Sw , % Sw , %

60-100 C 150-275 C
Steam Temperature

Steam Temperature
40 40

20 UTF Field Data 20 UTF Field Data


Model 1 Model 1

0 0
0 25 50 75 100 125 150 175 200 225 0 25 50 75 100 125 150 175 200 225
Temperature, C Temperature, C

50 100 150 200 250 300 350 400 50 100 150 200 250 300 350 400
Temperature, F Temperature, F

Fig. 8Comparison of conductive and convective heat-flux components estimated from field data obtained from vertical observa-
tion wells in Dover Phase B SAGD pilot vs. predictions calculated by Models 1 and 3 at different water relative permeabilities. See
Table 1 for fluid and reservoir properties.

It is suggested to be on the order of several meters by Sharma and of the depletion steam chamber and can be used with small errors
Gates (2011b), on the order of meters by Butler (1987, 1994a) and in final results.
Ito and Ipek (2005), and on the order of millimeters to tens of
centimeters by Gotawala and Gates (2008). There is a no doubt
on the positive effect of steam-fingering, but there are many Nomenclature
unknowns associated with fingering, such as the steam finger c compressibility of the oil sand, 1/Pa
lengths and the rate of mixing and, in turn, its effect on convective cpc condensate heat capacity, J/kg C
flux. The steam fingering and its enhancing effect on convective cpr reservoir heat capacity, J/kg C
flux should be studied in detail. k absolute permeability of the reservoir, m2
krw relative permeability of water, no unit
int
krw relative permeability of water at the interface, no unit
Conclusions res
krw relative permeability of water in the reservoir, no unit
This study examines the relative roles of convective and conduc- K reservoir thermal conductivity, W/m  C
tive heat transfer at the edge of SAGD steam chambers. The results n the Corey coefficient, which sets the curvature of the rel-
demonstrate that convection can transfer a relatively large amount ative permeability curves, no unit
of heat at the edges of a steam chamber. However, temperature P fluid pressure, Pa
enhancement as a result of convection is minor, and is less than Pr reservoir pressure, Pa
1% under common Athabasca reservoir conditions. Therefore, this Pst steam pressure, Pa
study supports both Edmunds (1999a, b) argument and Butlers Sio initial oil saturation in the reservoir, no unit
conductive heat-transfer assumption, which assumes that conduc- Sor residual oil saturation (ROS) in the reservoir, no unit
tion is the only source of heat transfer to the cold formation ahead Swc connate-water saturation in the reservoir, no unit

142 February 2013 SPE Journal


Swr water (or condensate) saturation in the reservoir ahead of Carslaw, H.S. and Jaeger, J.C. 1959. Conduction of Heat in Solids, Oxford
the front, no unit University Press, Second Edition (First Edition, 1946).
Tr reservoir temperature,  C Clark, H.P., Ascanio, F.A., Van Kruijsdijk, C., et al. 2010. Method to Improve
Tst steam temperature,  C Thermal EOR Performance Using Intelligent Well Technology: Orion
Ux velocity of the advancing front of steam chamber, m/s SAGD Field Trial. Paper SPE 137133 presented at the Canadian Un-
Vc condensate convective velocity normal to the steam conventional Resources and International Petroleum Conference,
chamber edge, m/s Calgary, Alberta, Canada, 1921 October. http://dx.doi.org/10.2118/
V yc condensate Darcy velocity in the direction parallel to the 137133-MS.
steam chamber interface in steam chamber plane, m/s Collins, P.M. 2005. Geomechanical effects on the SAGD process. Interna-
V zc condensate Darcy velocity parallel to injector and pro- tional Thermal Operations and Heavy Oil Symposium, Calgary,
ducer wells, m/s Alberta, Canada, 13 November, SPE-97905-MS.
x normal distance to the advancing front of the steam Das, S.K. 2005. Wellbore Hydraulics in a SAGD Well Pair. SPE/PS-CIM/
chamber, m CHOA International Thermal Operations and Heavy Oil Symposium,
a reservoir diffusivity, m2/s Calgary, Alberta, Canada, 13 November. http://dx.doi.org/10.2118/
ac simplified condensate (water) diffusivity, m2/s 97922-MS.
/ porosity of the reservoir, no unit Das, S.K. and Butler, R.M. 1996. Countercurrent Extraction of Heavy Oil
lw viscosity of the condensate (water), cp and Bitumen. Paper SPE 37094 presented at the International Confer-
lint
w viscosity of water at the steam interface, cp ence on Horizontal Well Technology, Calgary, Alberta, Canada,
lres
w viscosity of water inside the reservoir, cp 1820 November. http://dx.doi.org/10.2118/37094-MS.
qc condensate density, kg/m3 Edmunds, N. 1999a. On the Difficult Birth of SAGD. J. Cdn. Pet. Tech.
qr reservoir density, kg/m3 38 (1): 1424. http://dx.doi.org/10.2118/99-01-DA.
Edmunds, N. 1999b. On the Difficult Birth of SAGD. Journal of Canadian
Acknowledgments Petroleum Technology, 38 (1) 24.
A special thank you goes to our helpful supervisor Rick Chalatur- Edmunds, N. 2000. Investigation of SAGD Steam Trap Control in Two
nyk. The supervision and support that he gave truly helped the and Three Dimensions. J. Cdn. Pet. Tech. 39 (1): 3040. http://
progression of this study. Our sincere thanks also go to D.A. Red- dx.doi.org/10.2118/50413-MS
ford from Alberta Research Council who motivated us to study Edmunds, N.R., Haston, J.A., and Best, D.A. 1989. Analysis and Imple-
the heat transfer in SAGD process; the idea of this study came mentation of the Steam Assisted Gravity Drainage Process at the AOS-
from his invaluable course at the University of Alberta. TRA UTF. In the 4th UNITAR/UNDP International Conference on
Heavy Crude and Tar Sands, 712 August 1988, Edmonton, Alberta,
References Canada, Vol. 4, pp. 223242.
Edmunds, N.R., Kovalsky, J.A., Gittins, S.D. et al. 1994. Review of Phase
Aherne, A.L. and Maini, B. 2006. Fluid Movement in the SAGD Process:
A Steam-Assisted Gravity Drainage Test. SPE Res Eval Eng 9 (2):
A Review of the Dover Project. Paper CIPC 2006-153 presented at the
119-124. SPE 21529. http://dx.doi.og/10.2118/21529-PA.
Canadian International Petroleum Conference, Calgary, Alberta, Can-
Farouq-Ali, S.M. 1997. Is There Life After SAGD? J. Cdn. Pet. Tech. 36
ada, 1315 June. http://dx.doi.org/10.2118/2006-153.
(6): 2024. http://dx.doi.org/10.2118/97-06-DAS.
Akin, S. 2005. Mathematical Modeling of Steam-Assisted Gravity Drainage.
Ferguson, F.R.S., and Butler, R.M. 1988. Steam Assisted Gravity Drainage
SPE Res Eval Eng. 8 (5): 372376. http://dx.doi.org/10.2118/86963-PA.
Model Incorporating Energy Recovery from a Cooling Steam Chamber.
Al-Bahlani, A.M. and Babadagli, T. 2009. SAGD Laboratory Experimen-
J. Cdn. Pet. Tech. 27 (5): 7583. http://dx.doi.org/10.2118/97-06-DAS.
tal and Numerical Simulation Studies: A Review of Current Status and
Future Issues. J. Petrol. Sci. Eng. 68: 135150. http://dx.doi.org/ Gotawala, D.R., and Gates, I.D. 2008. Steam Fingering at the Edge of a
10.1016/j.petrol.2009.06.011. Steam Chamber in a Heavy Oil Reservoir, Can. J. Chem. Eng. 86 (6):
10111022. http://dx.doi.org/10.1002/cjce.20117.
Azad, A. and Chalaturnyk, R.J. 2010. A Mathematical Improvement to
SAGD Using Geomechanical Modelling. J. Cdn. Pet. Tech. 49 (10): Ito, Y. 1999. Discussion of On the Difficult Birth of SAGD. J. Cdn. Pet.
5364. http://dx.doi.org/10.2118/141303-PA. Tech. 38 (1): 2224.
Bennion, D.B., Thomas, F.B., Schulmeister, B. et al. 2006. A Correlation Ito, Y. and Ipek, G. 2005. Steam Fingering Phenomenon During
of the Low and High Temperature Water-Oil Relative Permeability SAGD Process. Paper SPE 97729 presented at the SPE/PS-CIM/CHOA
Characteristics of Typical Western Canadian Unconsolidated Bitumen International Thermal Operations and Heavy Oil Symposium, Calgary,
Producing Formations. Paper presented at the Canadian International Alberta, Canada, 13 November. http://dx.doi.org/10.2118/97729-MS.
Petroleum Conference Calgary, Alberta, 1315 June. http://dx.doi.org/ Ito, Y. and Suzuki, S. 1996. Numerical Simulation of the SAGD Process
10.2118/2006-092. in the Hangingstone Oil Sands Reservoir. Paper 96-57 presented at the
Birrell, G. 2001. Heat Transfer Ahead of a SAGD Steam Chamber: A Annual Technical Meeting, Calgary, Alberta, Canada, 1012 June.
Study of Thermocouple Data from Phase B of the Underground http://dx.doi.org/10.2118/96-57.
Test Facility (Dover Project). Paper 2001-88 presented at the Pet. Soc. Ito, Y. and Suzuki, S. 1999. Numerical Simulation of the SAGD Process
of CIMs Canadian International Petroleum Conference, Calgary, in the Hangingstone Oil Sands Reservoir. J. Cdn. Pet. Tech. 38 (9):
Alberta, 1214 June. http://dx.doi.org/10.2118/71503-MS. 2735. http://dx.doi.org/10.2118/99-09-02.
Butler, R.M. and Stephens, D.J. 1981. The Gravity Drainage of Steam- Ito, Y., Suzuki, S., and Yamada, H. 1998. Effect of Reservoir Parameter
Heated Heavy Oil to Parallel Horizontal Wells. J. Cdn. Pet. Tech. 20 on Oil Rates and Steam Oil Ratios in SAGD Projects. Presented at the
(2): 9096. http://dx.doi.org/10.2118/81-02-07. 7th UNITAR International Conference on Heavy Crude and Tar Sands,
Butler, R.M. 1985. A New Approach to the Modelling of Steam-Assisted 2730 October 1998, Beijing, China. Tulsa, Oklahoma: International
Gravity Drainage. J. Cdn. Pet. Tech. 24 (3): 4251. http://dx.doi.org/ Center for Heavy Hydrocarbons.
10.2118/85-03-01. Jimenez, J. 2008. The Field Performance of SAGD Projects in Canada. Paper
Butler, R.M. 1987. Rise of Interfering Steam Chambers. J. Cdn. Pet. Tech. presented at the International Petroleum Technology Conference, Kuala
26 (3): 7075. http://dx.doi.org/10.2118/87-03-07. Lumpur, Malaysia, 35 December. http://dx.doi.org/10.2523/12860-MS.
Butler, R.M. 1991. Thermal Recovery of Oil and Bitumen. Englewood Liang, L. 2005. An Analytical Model for Cyclic Steaming of Horizontal
Cliffs, New Jersey: Prentice Hall. Wells, MS thesis, Stanford University, Stanford, California.
Butler, R.M. 1994a. Horizontal Wells for the Recovery of Oil, Gas and Bi- McCormack, M. 2002. SAGD Injection WellsWhat Your Prof. Never
tumen, Petroleum Society Monograph Number 2. Westmount, Quebec, Told You. J. Cdn. Pet. Tech. 41 (3): 1723. http://dx.doi.org/10.2118/
Canada: Canadian Institute of Mining, Metallurgy and Petroleum. 02-03-DA.
Butler, R.M. 1994b. Steam-Assisted Gravity Drainage, Concept, Develop- Nasr, T.N., Law, D.H.S., Golbeck, H., et al. 2000. Counter-Current Aspect
ment, Performance and Future., J. Cdn. Pet. Tech. 33 (2): 4450. of the SAGD Process. J. Cdn. Pet. Tech. 39 (1): 4147. http://
http://dx.doi.org/10.2118/94-02-05. dx.doi.org/10.2118/00-01-03.

February 2013 SPE Journal 143


  
Nukhaev, M., Pimenov, V., Shandrygin, A., et al. 2006. A New Analytical @T Ux Pst  Pr qc cpc kkrw /lw c
0
Model for the SAGD Production Phase. Paper SPE 102084 presented C0 exp  exp  Ux n
@n K lw kkrw
at the SPE Annual Technical Conference and Exhibition, San Antonio, 
Texas, 2427 September. http://dx.doi.org/10.2118/102084-MS. Ux
 n                          A-6
ORourke, J.C., Begley, A.G., Boyle, H.A., et al. 1999. UTF Project Status a
Update, May 1997. J. Cdn. Pet. Tech. 38 (9): 4454. http://dx.doi.org/
10.2118/97-08. Eq. A-6 is simplified for boundary-condition application for
ORourke, J.C., Chambers, J.I., Suggelt, J.C. et al. 1994. UTF Project Sta- the dimensionless temperature variable defined by Eq. A-7:
tus and Commercial Potential: An Update, May 1994. Paper PETSOC
T  Tr
94-40 presented at the Annual Technical Meeting, Calgary, Alberta, T . . . . . . . . . . . . . . . . . . . . . . . . . . . . A-7
Canada, 1215 June. http://dx.doi.org/10.2118/94-40. Tst  Tr
Reid, R.C., Prausnitz, J.M., and Poling, B.E. 1977. The Properties of
When this substitution is made, Eq. A-6 for the dimensionless
Gases and Liquids. New York: McGraw-Hill.
temperature variable becomes
Reis, J.C. 1992. A Steam-Assisted Gravity Drainage Model for Tar Sands:
Linear Geometry. J. Cdn. Pet. Tech. 31 (10): 1420. http://dx.doi.org/   
@T  Ux Pst  Pr qc cpc kkrw /l c
10.2118/92-10-01. C0 exp  exp  w Ux n
@n K lw kkrw
Reis, J.C. 1993. A Steam-Assisted Gravity Drainage Model for Tar Sands: 
Radial Geometry. J. Cdn. Pet. Tech. 32 (8): 4348. http://dx.doi.org/ Ux
 n                          A-8
10.2118/93-08-05. a
Schmidt, E., and Grigull, U. 1981. Properties of Water and Steam in SI-
Units, Springer, Second Edition (First Edition, 1979). The following variable change is applied for simplicity:
Sharma, J., and Gates, I.D. 2011a. Convection at the Edge of a Steam-
Assisted-Gravity-Drainage Steam Chamber. SPE J. 16 (3): 503512. Ux Pst  Pr qc cpc kkrw
g . . . . . . . . . . . . . . . . . . A-9
SPE 142432-PA. http://dx.doi.org/10.2118/142432-PA. K lw
Sharma, J., and Gates, I.D. 2011b. Interfacial Stability of In-Situ Bitumen
Thermal Solvent Recovery Processes. SPE J. 16 (1): 5564. SPE /lw c
130050-PA. http://dx.doi.org/10.2118/130050-PA. j Ux . . . . . . . . . . . . . . . . . . . . . . . . . . . A-10
kkrw
Shaw, J., and Bedry, M. 2012. SAGD Field Trial for a New Intelligent Well
Completions Strategy to Increase Thermal EOR Recoveries. Paper SPE
150477 presented at the SPE Intelligent Energy International, Utrecht, When these substitutions are made, Eq. A-8 is converted to
The Netherlands, 2729 March. http://dx.doi.org/10.2118/150477-MS.  
@T  Ux
Williamson, 1972. Chemical Engineering, May 15, 128. C0 exp g expj n n ! T 
@n a
 
Appendix A Ux
C0 exp g expj n n dn C1 . . . . A-11
Eq. 32 yields the following: a
 2     The following variable change is applied to integrate this
@ T /lw c
K  U x / cP st  Pr exp  U x n qc cpc equation over distance n:
@n2 kkrw
 
@T x expj n . . . . . . . . . . . . . . . . . . . . . . . . . . A-12
Ux qr cpr 0                     A-1
@n
In changing an integrated variable for variable x, Eq. A-11 is
This can be solved when written as follows, with small converted to
changes to Eq. A-1:  
Ux
 2      C0 ja 1
@ T Ux / cPst  Pr qc cpc /lw c T  expg x  x dn C1 . . . . . A-13
 exp  U x n j
@n2 K kkrw
 
Ux @T
 0                        A-2 This integral is solved using a Taylors series expansion of an
a @n exponential function inside the integral, as follows:
   
The following variable change is applied: Ux X1 n Ux
1 x 1
@T expg x x ja gn  x ja
k . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . A-3 n 0
n!
@n  
Ux
X
1 ja n1
Substituting this variable change, Eq. A-2 yields x
gn . . . . . A-14
    n!
dk Ux / cPst  Pr qc cpc /l c Ux n 0
exp  w Ux n  dn
k K kkrw a
                   A-4 The dimensionless temperature variable (T*) ahead of the front
is found by integrating Eq. A-14, yielding Eq. A-15:
When integrated over distance n from the steam chamber
8   9
front, Eq. A-4 yields Ux
>
< X
1 ja n >
=
   C0 x
Ux Pst  Pr qc cpc kkrw /l c T  gn C1
k C00 exp  exp  w Ux n j :>n 0 n!  Ux =ja n >
;
K lw kkrw
 8   9
Ux Ux
 n                            A-5 >
< X
1 ja n >
=
a C0 expj n
 gn C1
j >:n 0 n!  Ux =ja n >
;
From Eq. A-3, variable k shows the time derivative of the tem-
perature normal to the steam chamber interface (i.e., the tempera-
ture variation normal to the steam chamber interface).                    A-15

144 February 2013 SPE Journal


 
If n yields infinity, then the dimensionless temperature vari- Ux
X1 ja n
able (T*) yields zero. expj n
gn
n 0
n!  U x =ja n
T ( )
T  0 ! C1 0 . . . . . . . . . . . . . . . . . . . . . . A-16 ja X1
1
gn
Ux n 1 n!  Ux =ja n
 
So, C1 equals zero and, thus, the dimensionless temperature Ux
n
variable (T*) yields unity. The C0 constant in the following can be X
1
P ja
calculated as: gn
n 0
n!  Ux =ja n
( ) . . . . . A-19
( ) ja X1
1
gn
ja X
C0 1
1 Ux n 1 n!  Ux =ja n
T  1 !  gn
Ux n 1 n!  Ux =ja n
j
Mazda Irani is a reservoir engineer at RPS Energy. He is currently
j engaged in designing the oil and gas development projects,
1 ! C0 ( )
ja X
1
1 n
especially waterflooding projects. In 2012, Irani completed his
g PhD dissertation at the University of Alberta, which focused on
Ux n 1 n!  Ux =ja n the risk assessment of Weyburn Carbon Sequestration Project.
                   A-17 During his PhD studies, Irani worked on a part-time basis as a
Finite Element Analyzer in C-FER Technologies. During this pe-
riod, he dealt with designing slotted liners and thermal cement-
ing with the use of finite-element analysis specifically the
In the specific case in which g equals zero. C0 is identical to
ABAQUS stress analysis modeling program. Irani also holds two
the the coefficient in Eq. 22 for the convection-neglecting
MSc degrees from the Sharif University and Ferdowsi University,
assumption: respectively, and a BSc degree from Ferdowsi University. He is a
member of the Petroleum Society of Canada.
Ux Sahar Ghannadi is currently a PhD candidate in petroleum en-
C0  . . . . . . . . . . . . . . . . . . . . . . . . . . . . A-18 gineering at the University of Alberta. Her current research inter-
a est is the application of Electro-Magnetic SAGD (EM-SAGD)
technology in Athabasca bitumen deposits and its effect on
caprock integrity. Ghannadi holds a BSc degree in electrical
Substituting the C0 in Eq. A-15 yields the dimensionless tem- engineering and an MSc degree in control engineering, both
perature variable (T*) ahead of the front, yielding Eq. A-19: from Azad University in Iran.

February 2013 SPE Journal 145

También podría gustarte