You are on page 1of 19

SPE 113358

Foam Modeling in Heterogeneous Reservoirs Using Stochastic Bubble


Population Approach
F. Farshbaf Zinati, R. Farajzadeh and P.L.J. Zitha, Department of Geotechnology, Delft University of Technology,
The Netherlands
Copyright 2008, Society of Petroleum Engineers

This paper was prepared for presentation at the 2008 SPE/DOE Improved Oil Recovery Symposium held in Tulsa, Oklahoma, U.S.A., 1923April2008.

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
Foam is an attractive option in EOR for increasing oil recovery in mature water-flooded reservoirs. In this paper we use
stochastic bubble population model and complex power law rheological model, to integrate foam physics into a flow
simulator. Foam displacement is examined in layered reservoirs with and without isolating shale barrier between the layers and
in stochastically distributed permeability fields. It is demonstrated that in isolated layers foam propagates faster in the high
permeability layer and sweeps the low permeability layer only modestly. In communicating layers, sweep efficiency is
improved significantly due to cross flow. In stochastic random permeability field, foam injection increases the liquid recovery
by a factor of two in comparison to gas injection.

2 SPE 113358
Introduction
Foam is an excellent mobility control agent for Enhanced Oil Recovery (EOR) processes, such as gas (nitrogen, carbon
dioxide, etc.) or steam injection [1-8]. Due to its unique microstructure, foam dramatically reduces gas mobility and
cosiderably improves the vertical and areal reservoir sweep efficiency [1-4]. The ability to reduce fluid mobility forms the
basis of profile correction used to improve water- or steam-flooding in highly heterogeneous (layered) formations [7]. Foam
has also been used for gas blocking in oil wells [5] and of acid diversion in matrix stimulation operations [2]. In these near-
wellbore applications foam is particularly attractive because it is inherently non-damaging and low cost, allowing easily
recursive treatments in case of an unsuccessful operation [1-4].
The evaluation and implementation of foam processes rely on experiments, modeling and numerical reservoir simulations.
This work is mainly concerned with the modeling and numerical simulation of foam. Several foam models have been proposed
in the last three decades, including semi-empirical models [9], fractional flow models [10, 11], population balance model [12-
15], and percolation theory [16]. Recently, Zitha [17] examined the strengths and limitations of the available models. In order
to capture more realistically the physics of foam motion in porous media, Zitha [17] developed a new stochastic bubble
population (SBP) foam model, with the main features described below:
Foam is a complex fluid, characterized by a yield stress and, above the yield stress, by a power law behavior. Its rheology can
be described using a Herchel-Bulkley type of model where the constitutive parameters depend mainly on the bubble density,
also referred to as the foam texture. Since on a large scale foam generation occurs over a large number of randomly
interconnected pores, bubble generation can be treated as a stochastic process. The kinetic of foam generation can thus be
described by a simple exponential growth law involving only two unknown parameters that can be determined from core flow
experiments. The SBP foam model was applied successfully to simulate different core-flow experiments [18, 19]. The
simulation results showed a rather good match with fluid partitioning images obtained by X-ray computed tomography (CT)
[18, 19]. Both experiment and simulations in layered cores showed that foam propagates faster in the high permeability layers
[18].
In this paper, we build a novel multi-phase multi-dimension reservoir simulator based on the SBP foam model and provide a
numerical analysis of foam flow in heterogeneous porous media. We examine several scenarios concerning the co-injection of
liquid (surfactant solution) and gas (N
2
, CO
2
, etc.) in layered and randomly heterogeneous reservoirs. We show that foam
performance, i.e. the amount of fluid recovery during foam injection, is higher with foam regardless of reservoir heterogeneity.

1. Stochastic Bubble Population Model

1.1 Physical Aspects

1.1.1 Foam rheology
As mentioned above, an important feature of SBP foam model is the rheology of foam and its dependence on bubble density.
In the SBP model [17] foam is treated as a complex threshold fluid in its own right, with a power-law character. Accordingly,
foam is characterized by a yield stress
y
such that: a) when
y
< foam does not shear, b) when
y
> the foam shears with
a power-law behavior. Hence, foam rheology can adequately be described by a Herschel-Bulkley type of model as follows:

1 1
1 2
0,
,
f y
m
f g f f y y
u for
K nu K u for



=

= + + >

(1)

where,
f
and
g
are the apparent viscosity of flowing foam and the gas viscosity, respectively, n is the bubble density,
f
u =
f
u is the norm of foam velocity
f
u , and
1
K and
2
K are constants. These constants depend on the properties of the
porous medium and can be estimated from a capillary model for the porous medium.
y
is a function of the bubble density n .
Equation (1) also indicates that when 0
f
u then
f
(solid behavior) whereas
f
u implies that
f g
. Zitha
[17] argued that during liquid displacement by foam the condition
y
> is readily achieved in the entire flow domain. Thus,
assuming the term
y
to be negligible, the viscosity of the flowing foam reduces to:

1
1
m
f g f
K nu

= + (2)

The above equation is reminiscent of the Hirasaki and Lawsons foam viscosity formula [21], which is based on Bretherthon
model [22]. It stresses that foam viscosity depends mainly on the bubble density n and foam velocity
f
u . Foam viscosity
increases with the bubble density and reduces to gas viscosity when 0 n . Taking 2 3 m = , according to Hirasaki and
Lawson, we observe that foam is also shear-thinning.

SPE 113358 3
1.1.2 Relative permeabilities
The mobility of the i
th
phase is given in the standard way, by the ratio between the (partial) phase permeability
i
k and the
corresponding viscosity
i
, i.e.
i i i ri i
k kk = = , where
ri
k is the relative permeability of the phase i. The SBP foam
model assumes that the phase relative permeabilities depend only on water saturation, ( )
ri ri w
k k S = , and provided that
y
>
over the entire flow domain, they are identical to those in the free (unfoamed) gas case. This means that the relative
permeability functions ( )
ri ri w
k k S = are hardly modified when free gas is replaced by foam. Therefore common Brooks-Corey
type of correlations are used assuming that endpoint relative permeabilities for water and foam are 0.5 and 1, respectively. The
rock sorting factor and connate water saturation are 5 and 0.25 correspondingly.

1.1.3 Capillary Pressure
Likewise, according to a common combination of Corey and Leverett expressions, the capillary pressure
c
p is a unique
function of the liquid saturation
w
S . An important modification with respect to free gas-liquid mixtures is the lowering the
surface tension due to the adsorption of surfactant molecules at the liquid-gas interface. Moreover, the non-uniqueness of the
capillary pressure functions, associated with the hysteresis of capillary pressure during successive drainage and imbibition
cycles is ignored. Furthermore, foam collapse at a certain limiting capillary pressure,
*
c
p , is disregarded. Khatib et al. [20]
have shown that foam is inherently unstable for capillary pressures above
*
c
p . This parameter depends on various factors
including surfactant type and concentration (
*
c
p increases with increasing surfactant concentration), the pore diameter and
wettability. Several authors, e.g. [20], proposed that foam always flows near
*
c
p under steady-state conditions. In the present
study we are primarily concerned with foam generation and propagation during the primary drainage so that the liquid
saturation remains always high and
c
p is always much smaller than
*
c
p ; hence ignoring the latter is acceptable.

1.1.4 Fluid densities
In this study we will consider only foam flow under isothermal conditions in the entire reservoir [23]. Therefore, the fluid
densities are related to pressure changes using isothermal compressibility: the density of the i
th
fluid can be expressed in terms
of fluid pressure as
1
i
i i p i
c

= , where
i
c is the isothermal compressibility of the i
th
fluid. For small pressure variations
compared to a certain reference pressure, the isothermal compressibility is nearly independent of the pressure and the
integration of the equation above leads to a well-known exponential expression of fluid densities in terms of fluid pressures.
For small values of isothermal compressibility the exponential expression can be approximated and reduced to a linear
function, which consequently increases the speed of the simulation process.
1.1.5. Remarks
It is important to stress that the above discussions are only valid for
y
> . When
y
in some parts of the porous medium
foam is immobile, although fluid can still flow by the foam drainage process [17]. In an up-scaled representation, a residual
(trapped) gas term,
fr
S , should be accounted for in effective saturation expressions. Besides that the porosity should be
reduced by 1
fr
S and permeability by some power of 1
fr
S . Finally, the sorting factor can be modified as foam trapping
gives rise to a new porous medium. Fortunately, the condition
y
occurs only for steady-state foams. For the case of
transient foam that we are mainly interested in >
y
, 0
fr
S = is a good approximation.

1.2. Equations of Motion
As given in [17], the SBP foam model relies on the modification of two-phase flow equations for liquid and gas, taking into
account foam generation and foam rheology. In a compact form, the SBP model is expressed as:

( ) ( ) { }
( ) {
( ) ( )
( ) ( )
. 0, ,
, ,
0 for for
. ( )
t i i i i i
w w w
f y f f y
t f f f g d
S q i w f
P
t
or P
nS n D n S K n n K n




+ + =

= >

+ =

u
u
u
u
%
G
r (3)

where
i
,
i
S ,
i
u ,
i
and
i
P are the density, saturation, (filtration) velocity, mobility and total pressure of the i
th
fluid. G is a
function of the bubble density and is the stress. The third equation in the system (3) indicates that foam flows only when is
larger than a threshold value
y
and is immobile otherwise.
g
K and
d
K are bubble generation and bubble destruction rate
4 SPE 113358
coefficients and n

is the maximum bubble density. In principle


g
K depends on surfactant type and concentration, foam
velocity, presence and type of oil, etc. The coefficient D represents diffusive-dispersive transport contribution, which can be
neglected in comparison with convective contribution. The source or sink terms
i
q% (mass per unit volume per unit time)
represents material accumulation or depletion due to injection or production.
The total pressure
i
P of the i
th
phase is the sum of
i
p and an internal pressure
i
resulting from the action of external force
field acting upon the fluid body, that is:

i i i
P p = + (4)

An example of external force field is the gravity field. Choosing the vertical axis z in the upward direction, the associated
pressure term is
i i
gz = . To simplify the problem and focus on the effect of heterogeneity, in the present paper we neglect
the term
i
and, thus, replace
i
P by
i
p . Besides the set of equations (3) the closure conditions for our problem read:
1,
w f C f w
S S p p p + = = , (5)

which express the saturation and capillary pressure relations.

2. Numerical Simulation
We now turn to the numerical method to develop the 2D foam simulator that is based on the SBP model. To this end, we adopt
the Black Oil Model (also known as the -model), where water is the wetting phase and gas is the non-wetting phase. In
addition to the earlier assumption, i.e.,
y
>> (see section 1.1.1), we consider that
f
P >> G and 0
d
K over the entire
flow domain. Physically this means that foam is strong and has low yield stress. To comply with the -model we introduce the
formation volume factor
, i i sc i
B = and
, i i i sc
q q = % , where
, i sc
is the density of the i
th
fluid at standard conditions.
Substituting the Darcys law into the conservation equations, the set of equations (3) can be reduced to:

( ) ( )
( ) ( )
( ) ( )
1 1
1 1
.
.
( ) .
t w w w w w w
t f f f f f f
t f f g f f
B S q B p
B S q B p
nS S K n n n p




+ =

+ =

= +

(6)

Furthermore, using closure conditions (5) we express foam pressure and saturation in terms of liquid pressure and saturation.
As a result the following set of equations is obtained:

( ) ( ) ( )
( ) ( ) ( ) ( ) ( )
( ) ( ) ( ) ( )
1
1 2
1 1
1 2 2
.
. .
1 ( ) . .
w t w w t w w w w w
f t w f t w f f f w f f C f t C
t w f g f w f C
C S C p q B p
C S C p q B p B p C p
n S S K n n n p n p

+ + =

+ + = +

= + +

(7)

where

( ) ( ) ( )
1 1 1 1
1 2 1 2
, , , 1
w f
w w w w p w f f f w p f
C B C S B C B C S B

= = = =
(8)
The set of equations is expressed in terms of only three primary variables (
w
S ,
w
p and n ) and known functions of these
variables. This system of equations is to be solved for the primary variables with the appropriate initial and boundary
conditions.

2.1 Discretization
The discretization of the flow equations (7) was done using the Finite Volume Method (FVM). This method is similar to the
block centered Finite Difference Method (FDM) and it enables the boundary conditions to be treated more easily. The flow
domain is discretized in: 1) two dimensional Cartesian coordinates to N
x
grid blocks in horizontal direction and N
y
grid blocks
in vertical direction and in 2) two dimensional radial coordinates to N
r
grid blocks in radial direction and N
z
grid blocks in
vertical direction. The simulator is capable of using non-uniform grid cells in order to refine the grids in specific circumstances
and regions.

SPE 113358 5
2.2 Method of Solution
The dependent variables and their gradients are computed at the center of each grid block. The parameters involved in foam
and liquid mobilities, except for the absolute permeabilities of adjacent blocks, are upstream-weighed. The absolute
permeability in mobility or transmissibility between adjacent grid-blocks is obtained by harmonic averaging. This ensures the
convergence and stability of the numerical scheme [23].
The equations were solved using the IMPES method [24, 25], i.e. by treating the pressure implicitly and other dependent
variables (saturation, bubble density) explicitly. The first two equations of set (7) are assembled to eliminate saturation terms.
The obtained equation only involves liquid pressure as an unknown. This leads to a set of three decoupled equations involving
the three primary dependent variables:

1 2 3 4
M M M
n+1 n n
w w c
= + + P P P M (9)
5 6 7
M M
n+1 n n n+1
w w w w
= + + + S S P P M (10)
1
8 9
M
n n +
= + n n M (11)

The matrixes
1
M ,
2
M ,
3
M ,
5
M ,
6
M and
8
M describe the transmissibilities and accumulations and the vectors
4
M ,
7
M ,
9
M
correspond to the boundary conditions. The details of derivations and decoupling using IMPESS method are explained in
appendix A.
To deal with the nonlinearities (dependency of mobilities on the primary variables) and control the stiffness associated with
diffusive terms, the pressure is solved by an iterative method, based on fully implicit treatment of mobilities with respect to
pressure [23]. After a new pressure field has been calculated the transmissibility matrices are updated and the system is solved
again until it meets a certain convergence criteria.
At each time step, in each grid block, the magnitude of the vector representing the gas interstitial velocity is used to compute
foam viscosity from equation (2). The velocity is obtained by first summing the flow of foam into and out of the grid block in
the two orthogonal directions. Then the arithmetic average for each direction is calculated and the magnitude of the resultant
vector is used to calculate the shear-thinning portion of the foam effective viscosity.
The algorithm used to solve the system of equations (9)-(11) is shown in the flow diagram of Fig. 1. The pressure is obtained
by solving equation (9) for the pressures of the current time step using saturations and bubble densities of the previous time
step. Then the saturation equation (10) is solved explicitly for the current time step using the updated pressure field. The
updated pressure and saturation fields are inserted to equation (11) and new foam bubble density is calculated explicitly.

Calculate M1 to M9
Is C0nvergence
Satisfied?
Initialize Flow
Problem Parameters
Update Saturation
Solve Pressure
Equation
Update
Transmissibilities in
Matrixes
Update Bubble
Density
Is Time
Step Below
the Limit?
Has
Simulation
Time Ended?
Discard the
Calculations and
Modify the Time
Step
Calculate next time
step and Increase
the Time
No
No
Yes
Yes
No
Start
End
Yes
Calculate M1 to M9 Calculate M1 to M9
Is C0nvergence
Satisfied?
Initialize Flow
Problem Parameters
Initialize Flow
Problem Parameters
Update Saturation Update Saturation
Solve Pressure
Equation
Solve Pressure
Equation
Update
Transmissibilities in
Matrixes
Update
Transmissibilities in
Matrixes
Update Bubble
Density
Update Bubble
Density
Is Time
Step Below
the Limit?
Is Time
Step Below
the Limit?
Has
Simulation
Time Ended?
Has
Simulation
Time Ended?
Discard the
Calculations and
Modify the Time
Step
Discard the
Calculations and
Modify the Time
Step
Calculate next time
step and Increase
the Time
Calculate next time
step and Increase
the Time
No
No
Yes
Yes
No
Start
End End
Yes

Fig. 1. Algorithm of multi-dimensional foam simulator.

6 SPE 113358
2.3 Time Step Control
The stability criteria for IMPES method combined with the required accuracy in the computations impose restrictions to the
allowed time step [23]. Moreover, most of the standard expressions for maximum allowed time step are valid only for standard
two-phase flow problems. To the best of our knowledge, no stability analysis has been reported for our kind of flow problem,
i.e. for the case where the two-phase flow equations is modified by introducing bubble density equation. In this case, the
stability analysis is rather complicated, especially because foam transmissibility depends on bubble density and this
dependence is treated explicitly. Fortunately, our analysis shows that the decrease in mobility due to foam generation improves
the stability and allows large time steps to be used.
Since we are dealing with transient flow where the flow parameters change significantly over time, the selection of a constant
time step would lead to an unreasonably long simulation times. Moreover, the use of largest stable time steps will give the
smallest truncation error [23]. Note that calculating time step using the complex expression derived from stability analysis of
our flow problem would lead to a significant increase in computing time [23]. Hence, the time step is selected such that it
ensures the relative change of the primary variables (saturation, pressure and bubble density) over each time step in all grid
blocks remains below a certain value (e.g., 1%). This value is obtained from many numerical experiments and ensures that the
stability is maintained. To ensure both the stability of the solution and an acceptable truncation error, we have implemented the
following time step selection procedure.
At the beginning of each time step, the maximum of saturation, pressure and bubble density change over previous (old) time
step is calculated and the size of current (new) time step is determined according to:

min , ,
NEW OLD OLD OLD Allowed Allowed Allowed
OLD OLD OLD
MAX MAX MAX
S P n
t t t t
S P n

=



(12)

where
Allowed
S ,
Allowed
P and
Allowed
n are predefined maximum allowed values for saturation, pressure and bubble density
change and
OLD
MAX
S ,
OLD
MAX
P and
OLD
MAX
n are the corresponding maximum value of the change in saturation, pressure and bubble
density over all the grid cells in old time step. Applying this procedure maintains saturation, pressure and bubble density
changes close to the corresponding assigned maximum allowed values. Since the set of equations are non-linear and coupled,
it is possible that the saturation, pressure and bubble density changes exceed the limiting values. To tackle this problem, at the
end of each time step, if the maximum allowed change values are exceeded, the time step size is modified (decreased)
according to an iterative method and the flow problem is solved once more. A simplified formula for the modified time step in
case of instability can be written as follows:

min , ,
NEW NEW NEW NEW Allowed Allowed Allowed
Modified NEW NEW NEW
MAX MAX MAX
S P n
t t t t
S P n

=



(13)

where
NEW
MAX
S ,
NEW
MAX
P and
NEW
MAX
n are the maximum value of the change in saturation, pressure and bubble density over all the
grid cells in current time step.

2.4 Well Model
Relation between the injection or production terms with the pressure and saturation of the grid blocks for the wells located in
interior grid cells is expressed in terms of single well productivity index (WI):

( )
, , , i j i j i j wf
Q WI p p = (14)

where
, i j
Q is the production or injection rate of fluid i (water or gas) in grid block j,
, i j
WI is the well productivity index of the
well located in grid block j for fluid i and only depends on pressure, saturation and absolute permeability of that grid block,
, i j
p is the pressure of the fluid i in grid block j and
wf
p is flowing bottom-hole pressure (BHP) and is used as a general term
for both production and injection wells.
For the wells located at the boundaries, a proper boundary condition is applied to the flow problem that is consistent with the
well operation mode (e.g. constant rate or constant bottom-hole pressure).

3. RESULTS AND DISCUSSION
There are numerous scenarios that can be modeled by the developed foam simulator, ranging from different geometries and
boundary conditions to different initial conditions and injection and production scenarios. Among these we present four cases.
First, we consider foam flow in a two-layered reservoir with an impermeable shale barrier between layers. Second, we study
foam displacement in a reservoir similar to the first case with the difference that shale layer is removed and cross flow is
SPE 113358 7
allowed. Next, we simulate gas injection and foam injection in a 5 spot pattern into a reservoir with stochastically distributed
permeability field. Finally, we will illustrate the significant improvement in sweep efficiency and incremental recovery by
comparing foam injection with unfoamed gas injection.

3.1 Two Isolated Layers
Fig. 2 shows schematically the structure of the reservoir studied in the first case. The reservoir is 200 meters long and 20
meters wide. It consists of two uniform layers, with an equal width of 10 m, separated by an impermeable shale barrier with a
negligible width. The permeability and porosity of the high permeable layer are 1500 mD and 20%, respectively; those of the
low permeability layer are 150 mD and 15%, respectively. In the simulations we assume that the reservoir is initially saturated
with surfactant solution 1
w
S = .
1.5 d-20%
0.15 d-15%
Foam Quality=95%
Qinj=400 cuft/Day
Pwf=2000 psia
200 m
20 m
1.5 d-20%
0.15 d-15%
Foam Quality=95%
Qinj=400 cuft/Day
Pwf=2000 psia
200 m
20 m

Fig. 2. Geometry of the case study 1. Reservoir with two layers separated by an impermeable shale layer.


The wells placed along the length of the reservoir are perforated through entire width of the layers. At the injection boundary
gas and surfactant solution mixture is injected in total rate of 400 cuft/d with foam quality of 95%. This is translated to
separate fluid injection rates of 380 cuft/d for gas and 20 cuft/d for surfactant solution where all the rates are expressed in
standard condition. Continuity of the pressure in injection and production boundaries are maintained; therefore, although the
total flow rate is kept constant, each layer might intake different portion of the flow. The fluid is produced in the production
well at bottom-hole pressure of 2000 psia.
The simulation is done with 40 grids in horizontal direction and 8 grids in vertical direction leading to grid spacing of 5 meters
in horizontal and 2.5 meters in vertical direction.
Fig. 3 shows the simulated saturation profiles. Liquid saturations in all grid cells of each layer are averaged over the vertical
direction. The obtained liquid saturations are plotted against dimensionless horizontal distance (x/L) for both layers at different
dimensionless times (dimensionless time 1 corresponds to a time that 1 pore volume injection of the displacing fluid is
completed). Blue dashed lines correspond to low permeable layer (LPL) and solid black lines to high permeability layer
(HPL). As depicted in Fig. 3, foam propagates with a much higher speed in HPL than LPL. Foam breakthrough in HPL
happens after injection of 0.37 pore volumes (pore volume of total reservoir) while at this time it has flooded only 22% of
LPL.
Higher velocity in HPL leads to lower foam viscosity and consequently to lower flow resistance. The lower flow resistance in
HPL in comparison with LPL has two consequences, first HPL intakes larger proportion of the injected fluids and second, due
to the low mobility of foam in LPL the sweep efficiency of displacing fluid (foam) is larger. This fact becomes more visible if
we look at the residual saturation of initial fluid left behind in upstream. The high permeable layer desaturates to liquid
saturation of 0.4 (local liquid recovery of 60%) after the break through happens while this value is approximately 0.44 during
transient flow (for example 0.1 PV). Unlike in HPL, in LPL even after injection of 0.5 PV foam, the left behind liquid
saturation at upstream of flow reaches below 0.36. It is worthy to stress that although the sweep efficiency of foam in LPL is
larger than HPL, due to the fact that displacing process occurs very slowly, it seems however unfavorable in practice.
8 SPE 113358
0 0.2 0.4 0.6 0.8 1
0
0.2
0.4
0.6
0.8
1
Dimensionless Distance x/L
L
i
q
u
i
d

S
a
t
u
r
a
t
i
o
n

S
w
Saturation Profile
0.1
0.2
0.1
0.5
0.4
0.3
0.2
0.5
0.4
0.3

Fig. 3. Saturation profile for the two layers of case study 1. Blue dashed lines correspond to liquid saturation in LPL and black solid
lines to liquid saturation in HPL. The labels beside each curve indicate the corresponding dimensionless time (injected pore
volumes).

Fig. 4 depicts foam texture profiles for the same dimensionless times as saturations. The figure shows dimensionless bubble
density (defined as the bubble density divided by the maximum bauble density) as a function of dimensionless horizontal
distance. In a similar manner to saturations, bubble density is calculated by averaging over vertical sections for both layers.
Since we are not injecting any pre-generated foam the dimensionless bubble density is zero at the location of the production
well. Also since there is no foam initially in porous media, bubble density is zero at the parts that are not flooded by the foam.
The most interesting observation in Fig. 4 is the higher bubble density in LPL relative to HPL. In upstream of the flow the
foam texture in some regions in low permeable zone is even 4 times finer than the regions at the identical horizontal position
in the high permeable zone. This fact can be explained by higher resistance of the LPL to foam and lower velocity of foam in
that layer. The foam generation process occurs simultaneously in both layers, but since foam flows faster in HPL the foam
bubbles are transported with much higher speed. The bubble density increases in an approximately exponential manner from
zero to a relative maximum value at foam front. In LPL the trend is similar except for the fact that the rate of the increase is
considerably larger.
0 0.2 0.4 0.6 0.8 1
0
0.2
0.4
0.6
0.8
1
Dimensionless Distance x/L
R
a
t
i
o

o
f

F
o
a
m

B
u
b
b
l
e

D
e
n
s
i
t
y

(
n
/
n
m
a
x
)
Foam Texture Profile
0.2
0.5
0.3
0.4
0.1
0.1
0.3
0.2
0.4
0.5

Fig. 4. Foam texture profile for the two layers of case study 1. Blue dashed lines correspond to dimensionless bubble density in LPL
and black solid lines to dimensionless bubble density in HPL. The labels beside each curve indicate the corresponding
dimensionless time (injected pore volumes).

Liquid pressure profiles are shown in a manner identical to the previous two variables (saturation and bubble density) in Fig. 5.
Due to the continuity of the pressure at inlet and outlet (injection and production boundaries) as stated earlier, the pressure of
SPE 113358 9
both layers are equal for all the times at production well. As we go further towards the injection well in flow direction, a
significant pressure difference is observed between the two layers. This pressure difference exceeds 2500 psia after 0.5 pore
volume of injection. Unfortunately, the impermeable shale barrier prevents this pressure difference from causing any cross
flow. In LPL, the high foam viscosity as a result of high bubble density, low foam velocity and low permeability dominates the
effect of velocity in Darcy's law and creates a very steep pressure profile.

0 0.2 0.4 0.6 0.8 1
2000
2300
2600
2900
3200
3500
3800
4100
4400
4700
Dimensionless Distance x/L
L
i
q
u
i
d

P
r
e
s
s
u
r
e
Pressure Profile
0.1
0.3
0.4
0.2
0.5


Fig. 5. Pressure profile for the two layers of case study 1. Blue dashed lines correspond to liquid pressure in LPL and black solid
lines to liquid pressure in HPL. The labels beside each curve indicates the corresponding dimensionless time (injected pore volumes)

Fig. 6 displays the fraction of the initial liquid displaced from each layer as a function of time. The blue line corresponds to
recovery factor from LPL, the black line to recovery factor from HPL and the red line to the overall recovery factor. Before
0.37 DT the production from HPL increases linearly with a sharp slope of 62%. After injection of 0.37 PV breakthrough
occurs in HPL and production from this layer declines dramatically. From breakthrough until injection of 1 PV the fraction of
the recovered liquid increases slightly with a minor slope. Nonetheless, in LPL production profile is completely different. The
production process in LPL is much slower and the production curve increases with a much smaller slope than HPL. After 1 PV
injection, the fraction of initial liquid produced hardly reaches above 22%. As a result, the obtained overall recovery after 1
DT from the entire reservoir is approximately 45%.

0 0.2 0.4 0.6 0.8 1
0
0.2
0.4
0.6
0.8
1
Dimensionless Time (Injected PVs)
F
r
a
c
t
i
o
n

o
f

I
n
i
t
i
a
l

W
a
t
e
r

P
r
o
d
u
c
e
d
Production Profile


HPL
LPL
TOTAL


Fig. 6. Production profiles of case study 1. Fraction of initial liquid displaced as function of dimensionless time (injected pore
volumes)
Psi
3.2 Two Communicating Layers
The geometry of the reservoir for this case study is shown in Fig. 7. Note that initially both layers are fully saturated with
surfactant solution. As demonstrated, boundary conditions and well injection and production modes are identical to the
previous case of isolated layers i.e. gas and liquid mixture is injected by rate of 400 cuft/d at standard condition and volume
fraction of 95% for gas and 5% for liquid. At the other side of the reservoir fluids are produced at bottom-hole pressure of
2000 psia. Although the reservoir is geometrically identical to the previous example, this time the impermeable shale layer has
been removed and thus, two layers are in communication and cross flow is allowed between the layers.

1.5 d-20%
0.15 d-15%
Foam Quality=95%
Qinj=400 cuft/Day
Pwf=2000 psia
200 m
20 m
1.5 d-20%
0.15 d-15%
Foam Quality=95%
Qinj=400 cuft/Day
Pwf=2000 psia
200 m
20 m

Fig. 7. Geometry of the case study 2; Reservoir with two communicating layers.

First we consider the liquid saturation profiles depicted in Fig. 8. The profiles are strikingly different than the isolated layers.
Foam propagation speed in LPL has dramatically increased in comparison with previous example. Foam front in both layers
are traveling in almost the same speed but with a shift that remains nearly unchanged during the displacement. Breakthrough
occurs after 0.5 PV injection in HPL and shortly after that (at 0.53 DT) in LPL. The late breakthrough in HPL compared to
previous example is explained by the fact that some portion of the foam is diverted to LPL.

0 0.2 0.4 0.6 0.8 1
0
0.2
0.4
0.6
0.8
1
Dimensionless Distance x/L
L
i
q
u
i
d

S
a
t
u
r
a
t
i
o
n

S
w
Saturation Profile
0.1
0.2 0.3
0.4 0.5

Fig. 8. Saturation profile for two layers of case study 2. Blue dashed lines correspond to liquid saturation in LPL and black solid lines
to liquid saturation in HPL. The labels beside each curve indicates the corresponding dimensionless time (injected pore volumes)

The high permeable zone is invaded by injected gas and liquid earlier and due to the foam generation and flow the pressure
builds up faster than the low permeable zone where the fluid invasion and displacement lags behind. The resultant pressure
difference leads to cross flow from HPL to LPL. Pressure profiles are depicted in Fig. 9. Again blue dashed line corresponds to
LPL and black solid line to HPL. It is noticeable that the pressure difference between layers has decreased significantly and the
pressures in both layers at upstream from the production well to near the foam front are equal. The selective characteristic of
the foam is noticeable: Since cross flow is allowed in this scenario, foam minimizes its flow resistance, e.g., when the local
SPE 113358 11
pressure gradient increases in one layer foam diverts and pushes the liquid to the other layer. This leads to a foam front
advancing in both layers with approximately equal speed.

0 0.2 0.4 0.6 0.8 1
2000
2300
2600
2900
3200
3500
3800
4100
4400
Dimensionless Distance x/L
L
i
q
u
i
d

P
r
e
s
s
u
r
e
Pressure Profile
0.1
0.4
0.3
0.2
0.5

Fig. 9. Pressure profile for the two layers of case study 2. Blue dashed lines correspond to liquid pressure in LPL and black solid
lines to liquid pressure in HPL. The labels beside each curve indicates the corresponding dimensionless time (injected pore
volumes).

Fig. 10 shows foam texture profile Two major differences with the previous case study can be explained by the effect of cross
flow: first, unlike the case of isolated layers, bubble density is decreasing in LPL from production well to foam front; second,
at foam front in LPL which is close to foam front in HPL, bubble texture is approximately similar in both zones and the drastic
difference disappeared as a result of cross flow.

0 0.2 0.4 0.6 0.8 1
0
0.2
0.4
0.6
0.8
1
Dimensionless Distance x/L
R
a
t
i
o

o
f

F
o
a
m

B
u
b
b
l
e

D
e
n
s
i
t
y

(
n
/
n
m
a
x
)
Foam Texture Profile
0.4
0.1
0.3
0.2
0.5

Fig. 10. Foam texture profile for the two layers of case study 2. Blue dashed lines correspond to dimensionless bubble density in LPL
and black solid lines to dimensionless bubble density in HPL. The labels beside each curve indicates the corresponding
dimensionless time (injected pore volumes).

Fig. 11 depicts production profiles. Interestingly, initial liquid recovery from the low permeable layer has increased
significantly whereas the high permeable layer shows more or less the same recovery as previous example. After injection of 1
PV, the recovery factors from LPL and HPL are 55% and 63% respectively. In total 60% of the initial liquid in the reservoir
has been produced.


Psi
12 SPE 113358
0 0.2 0.4 0.6 0.8 1
0
0.2
0.4
0.6
0.8
1
Dimensionless Time (Injected PVs)
F
r
a
c
t
i
o
n

o
f

I
n
i
t
i
a
l

W
a
t
e
r

P
r
o
d
u
c
e
d
Production Profile


HPL
LPL
TOTAL

Fig. 11. Production profiles of case study 2. Fraction of initial liquid displaced as function of dimensionless time (injected pore
volumes).

3.3 Foam Simulations in a stochastic permeability field
To show the capability of the developed simulator to predict the foam performance in realistic geological settings, we simulate
simultaneous injection of gas and surfactant solution to one quarter of a confined conventional five-spot pattern. A
permeability field where the permeability is log-normally distributed is generated (the logarithm of the permeability is
normally distributed). The field has a Dykstra-Parsons coefficient of VDP = 0.7 and an expected permeability of 1000 mD.
Fig. 12 shows the geometry of the simulated domain with the injection and production parameters for the foam scenario. It is
assumed that the reservoir is confined at all four boundaries by sealing faults and therefore no flow boundary condition has
been applied in the simulator. The original reservoir is assumed to be square with dimensions of 200 meters by 200 meters.
Thus, the simulated quarter is 100 meters long and 100 meters wide. The injection well is located at South-West (bottom-left)
corner of the reservoir and production well at North-East (top-right) corner.


Foam Quality=90%
Qinj=200 cuft/Day
1 2 3 4 5 6 7 8 9 10
1
2
3
4
5
6
7
8
9
10

-1.5
-1
-0.5
0
0.5
100 m
100 m
Pwf=2000 psia
L
o
g
(
K
)
Foam Quality=90%
Qinj=200 cuft/Day
1 2 3 4 5 6 7 8 9 10
1
2
3
4
5
6
7
8
9
10

-1.5
-1
-0.5
0
0.5
100 m
100 m
Pwf=2000 psia
L
o
g
(
K
)
1 2 3 4 5 6 7 8 9 10
1
2
3
4
5
6
7
8
9
10

-1.5
-1
-0.5
0
0.5
1 2 3 4 5 6 7 8 9 10
1
2
3
4
5
6
7
8
9
10

-1.5
-1
-0.5
0
0.5
100 m
100 m
Pwf=2000 psia
L
o
g
(
K
)

Fig. 12. Geometry and permeability field for case study 3. Color map shows the logarithm of permeability field where the
permeabilities are in Darcy.

For simplicity porosity is assumed to be uniform over the entire domain and harmonic averaging has been applied to absolute
permeabilities between adjacent cells. To provide contrast between highly efficient foam flooding with conventional gas
SPE 113358 13
injection, first we have run gas injection simulation for the described five-spot pattern and the results are compared with foam
injection. In this simulation, gas is injected at a rate of 200 ST cuft/d. The major modification in the model of the current case
is that the viscosity of the displacing fluid (unfoamed gas) is constant rather than being dependent upon the local bubble
density and velocity.
Fig. 13 shows the liquid saturation profiles at different injection times. The gas breakthrough occurs fast. The gas fingers that
make several diagonal paths to the production well are observable after day 50. Gas finds its way through a path connecting
the high permeability blocks and reaches the production well after only 100 days. It does not flow into the low permeability
blocks and these are bypassed.
Even the regions contacted by gas are poorly swept and the liquid saturation remains above 50% in these regions. After
breakthrough, very little desaturation occurs because pressure gradients remain rather low. We can clearly observe that the
saturation of the low permeability parts do not change specially the region at top-left corner (North-East) remains un-swept for
a long time and saturated 100% by aqueous phase.


0.2 0.4 0.6 0.8 1
0.2
0.4
0.6
0.8
1
x/L
y
/
W
0.2 0.4 0.6 0.8 1
0.2
0.4
0.6
0.8
1
x/L
y
/
W
0.2 0.4 0.6 0.8 1
0.2
0.4
0.6
0.8
1
x/L
y
/
W
0.2 0.4 0.6 0.8 1
0.2
0.4
0.6
0.8
1
x/L
y
/
W
0.2 0.4 0.6 0.8 1
0.2
0.4
0.6
0.8
1
x/L
y
/
W
0.2 0.4 0.6 0.8 1
0.2
0.4
0.6
0.8
1
x/L
y
/
W
0.2 0.4 0.6 0.8 1
0.2
0.4
0.6
0.8
1
x/L
y
/
W
0.2 0.4 0.6 0.8 1
0.2
0.4
0.6
0.8
1
x/L
y
/
W
25
Day
50
Day
100
Day
200
Day
600
Day
500
Day
400
Day
300
Day
0.2 0.4 0.6 0.8 1
0.2
0.4
0.6
0.8
1
x/L
y
/
W
0.2 0.4 0.6 0.8 1
0.2
0.4
0.6
0.8
1
x/L
y
/
W
0.2 0.4 0.6 0.8 1
0.2
0.4
0.6
0.8
1
x/L
y
/
W
0.2 0.4 0.6 0.8 1
0.2
0.4
0.6
0.8
1
x/L
y
/
W
0.2 0.4 0.6 0.8 1
0.2
0.4
0.6
0.8
1
x/L
y
/
W
0.2 0.4 0.6 0.8 1
0.2
0.4
0.6
0.8
1
x/L
y
/
W
0.2 0.4 0.6 0.8 1
0.2
0.4
0.6
0.8
1
x/L
y
/
W
0.2 0.4 0.6 0.8 1
0.2
0.4
0.6
0.8
1
x/L
y
/
W
25
Day
50
Day
100
Day
200
Day
600
Day
500
Day
400
Day
300
Day

Fig. 13. Liquid saturation profile for unfoamed gas injection. Liquid saturation profiles are given after 25, 50, 100, 200, 300, 400, 500
and 600 days after foam injection.

By co-injection of gas and surfactant solution foam generates in porous media and the results differ dramatically. In this
simulation gas-liquid mixture is injected with a rate of 200 cuft/d and a foam quality of 90% leading to an injection rate of
180 cuft/d for gas and 20 cuft/d for gas. The production well operates at constant bottom-hole pressure of 2000 psia. The
reservoir is assumed to be initially saturated with 100% liquid. The initial pressure of the reservoir is 2000 psia.
Fig. 14 depicts the liquid saturation profiles for different times as in the previous gas injection scenario. Foam is displaced in a
semi-spherical manner at the beginning but as it propagates further in the reservoir and faces resistance to flow from the low
permeable regions, diverts to zones that show less resistance (high permeability zones). Foam propagation to high permeable
regions causes transportation of more bubbles into these regions which then leads to higher foam viscosity and consequently to
higher flow resistance. This fact is visible in snap-shots corresponding to 100 and 200 days in Figs. 14 and 15: foam has been
displaced more in the north direction rather than in the east direction. If we look back at permeability field, in a radius of
approximately 50 meters from the injector in the north direction, more blocks of high permeability are located. Most of the
high permeable blocks in north direction are adjacent with other high or medium permeability blocks while in the east
direction high permeability regions are interrupted by very low permeability blocks.













14 SPE 113358



0.2 0.4 0.6 0.8 1
0.2
0.4
0.6
0.8
1
x/L
y
/
W
0.2 0.4 0.6 0.8 1
0.2
0.4
0.6
0.8
1
x/L
y
/
W
0.2 0.4 0.6 0.8 1
0.2
0.4
0.6
0.8
1
x/L
y
/
W
0.2 0.4 0.6 0.8 1
0.2
0.4
0.6
0.8
1
x/L
y
/
W
0.2 0.4 0.6 0.8 1
0.2
0.4
0.6
0.8
1
x/L
y
/
W
0.2 0.4 0.6 0.8 1
0.2
0.4
0.6
0.8
1
x/L
y
/
W
0.2 0.4 0.6 0.8 1
0.2
0.4
0.6
0.8
1
x/L
y
/
W
0.2 0.4 0.6 0.8 1
0.2
0.4
0.6
0.8
1
x/L
y
/
W
0.2 0.4 0.6 0.8 1
0.2
0.4
0.6
0.8
1
x/L
y
/
W
0.2 0.4 0.6 0.8 1
0.2
0.4
0.6
0.8
1
x/L
y
/
W
0.2 0.4 0.6 0.8 1
0.2
0.4
0.6
0.8
1
x/L
y
/
W
0.2 0.4 0.6 0.8 1
0.2
0.4
0.6
0.8
1
x/L
y
/
W
0.2 0.4 0.6 0.8 1
0.2
0.4
0.6
0.8
1
x/L
y
/
W
0.2 0.4 0.6 0.8 1
0.2
0.4
0.6
0.8
1
x/L
y
/
W
0.2 0.4 0.6 0.8 1
0.2
0.4
0.6
0.8
1
x/L
y
/
W
0.2 0.4 0.6 0.8 1
0.2
0.4
0.6
0.8
1
x/L
y
/
W
25
Day
50
Day
100
Day
200
Day
600
Day
500
Day
400
Day
300
Day

Fig. 14. Liquid Saturation profile for foam injection in case study 3. Liquid saturation profiles are given after 25, 50, 100, 200, 300, 400,
500 and 600 days after foam injection.

Foam texture profile for 100 and 200 days in Fig. 15 also indicates that foam is mainly generated and transported to upper
parts. Increasing bubble density in these regions causes the reduction of foam mobility, and foam tends to displace the liquid in
east direction. The process of foam diverting to regions where it faces less resistance continues until flow converges to the
injector.


0.2 0.4 0.6 0.8 1
0.2
0.4
0.6
0.8
1
x/L
y
/
W
0.2 0.4 0.6 0.8 1
0.2
0.4
0.6
0.8
1
x/L
y
/
W
0.2 0.4 0.6 0.8 1
0.2
0.4
0.6
0.8
1
x/L
y
/
W
0.2 0.4 0.6 0.8 1
0.2
0.4
0.6
0.8
1
x/L
y
/
W
0.2 0.4 0.6 0.8 1
0.2
0.4
0.6
0.8
1
x/L
y
/
W
0.2 0.4 0.6 0.8 1
0.2
0.4
0.6
0.8
1
x/L
y
/
W
0.2 0.4 0.6 0.8 1
0.2
0.4
0.6
0.8
1
x/L
y
/
W
0.2 0.4 0.6 0.8 1
0.2
0.4
0.6
0.8
1
x/L
y
/
W
25
Day
50
Day
100
Day
200
Day
600
Day
500
Day
400
Day
300
Day
0.2 0.4 0.6 0.8 1
0.2
0.4
0.6
0.8
1
x/L
y
/
W
0.2 0.4 0.6 0.8 1
0.2
0.4
0.6
0.8
1
x/L
y
/
W
0.2 0.4 0.6 0.8 1
0.2
0.4
0.6
0.8
1
x/L
y
/
W
0.2 0.4 0.6 0.8 1
0.2
0.4
0.6
0.8
1
x/L
y
/
W
0.2 0.4 0.6 0.8 1
0.2
0.4
0.6
0.8
1
x/L
y
/
W
0.2 0.4 0.6 0.8 1
0.2
0.4
0.6
0.8
1
x/L
y
/
W
0.2 0.4 0.6 0.8 1
0.2
0.4
0.6
0.8
1
x/L
y
/
W
0.2 0.4 0.6 0.8 1
0.2
0.4
0.6
0.8
1
x/L
y
/
W
25
Day
50
Day
100
Day
200
Day
600
Day
500
Day
400
Day
300
Day

Fig. 15. Foam texture profile for case study 3. Dimensionless bubble density profiles are given after 25, 50, 100, 200,
300, 400, 500 and 600 days after foam injection.

Fig. 16 depicts the pressure profiles for the same times as for saturation and bubble density profiles. This figure illustrates how
fast the reservoir gets pressurized as the foam is generated and flows within the porous media. The pressure at the injection
well reaches approximately 4000 psia after 500 days of injection.













SPE 113358 15


0.2 0.4 0.6 0.8 1
0.2
0.4
0.6
0.8
1
x/L
y
/
W
0.2 0.4 0.6 0.8 1
0.2
0.4
0.6
0.8
1
x/L
y
/
W
0.2 0.4 0.6 0.8 1
0.2
0.4
0.6
0.8
1
x/L
y
/
W
0.2 0.4 0.6 0.8 1
0.2
0.4
0.6
0.8
1
x/L
y
/
W
0.2 0.4 0.6 0.8 1
0.2
0.4
0.6
0.8
1
x/L
y
/
W
0.2 0.4 0.6 0.8 1
0.2
0.4
0.6
0.8
1
x/L
y
/
W
0.2 0.4 0.6 0.8 1
0.2
0.4
0.6
0.8
1
x/L
y
/
W
0.2 0.4 0.6 0.8 1
0.2
0.4
0.6
0.8
1
x/L
y
/
W
25
Day
50
Day
100
Day
200
Day
600
Day
500
Day
400
Day
300
Day
0.2 0.4 0.6 0.8 1
0.2
0.4
0.6
0.8
1
x/L
y
/
W
0.2 0.4 0.6 0.8 1
0.2
0.4
0.6
0.8
1
x/L
y
/
W
0.2 0.4 0.6 0.8 1
0.2
0.4
0.6
0.8
1
x/L
y
/
W
0.2 0.4 0.6 0.8 1
0.2
0.4
0.6
0.8
1
x/L
y
/
W
0.2 0.4 0.6 0.8 1
0.2
0.4
0.6
0.8
1
x/L
y
/
W
0.2 0.4 0.6 0.8 1
0.2
0.4
0.6
0.8
1
x/L
y
/
W
0.2 0.4 0.6 0.8 1
0.2
0.4
0.6
0.8
1
x/L
y
/
W
0.2 0.4 0.6 0.8 1
0.2
0.4
0.6
0.8
1
x/L
y
/
W
25
Day
50
Day
100
Day
200
Day
600
Day
500
Day
400
Day
300
Day

Fig. 16. Liquid pressure profile for case study 3. Pressure profiles are given after 25, 50, 100, 200, 300, 400, 500 and 600 days after
foam injection (pressure in psia).

3.3.1 Comparison between foam and unfoamed gas performance in porous media
In Fig. 17, the fraction of produced liquid is plotted against the time both for foam and gas. After 500 days approximately 60%
of the initial liquid is produced by foam injection whereas after the similar time gas has swept less than 24% of the initial
liquid. Production decline occurs after breakthrough for both cases i.e. 360 days for foam injection and only 100 days for gas
injection. After breakthrough production profile curve for foam is much steeper than the one for gas. This is another
illustration of the fact that even after breakthrough production continues in foam injection scenario while in gas injection case
almost no production occurs.
0 100 200 300 400 500
0
0.2
0.4
0.6
0.8
1
Time (Days)
F
r
a
c
t
i
o
n

o
f

I
n
i
t
i
a
l

W
a
t
e
r

P
r
o
d
u
c
e
d
Production Profile


Foam
Unfoamed Gas

Fig. 17. Comparison of production profiles for injection of foam and unfoamed gas.


Conclusions
We have built a robust and flexible simulator for simulating foam displacement in heterogeneous reservoirs. Foam physics
is directly inserted into the reservoir simulator using stochastic bubble population balance which includes a foam rheology
model.
We have shown that it is practical to model foam propagation in reservoir scale and in multi-dimensions without
sacrificing macro-scale physics and mechanisms of foam generation and propagation.
Simulated foam injection scenarios in heterogeneous two layered reservoir initially filled with surfactant solution has
shown that in the presence of an impermeable shale barrier between layers, foam diverts mainly to the high permeable
layer where it displaces the liquid at a higher speed with a large sweep efficiency while low permeable layer shows a
significant resistance to foam flow resulting in a very low recovery factor.
Psi
16 SPE 113358
In the absence of impermeable shale the cross flow plays a dominant role in diverting the fluids from HPL to LPL and
increasing the total recovery factor of the reservoir.
Foam injection into a five-spot pattern illustrates the drastic sweep efficiency of foam in liquid production from a
heterogeneous reservoir. While in simple gas injection, gas fingers through the high permeable zones leaving most parts
of the reservoir un-swept, foam sweeps the reservoir more uniformly. The liquid recovery factor for foam injection is 60%
while for gas injection it is 30%, i.e. 2 times less than foam injection.
The simulator based on the stochastic bubble population model is a valuable tool to simulate foam EOR processes and
predict feasibility of various EOR scenarios involving foam.

ACKNOWLEDGEMENTS
The authors gratefully acknowledge Alexey Andrianov and Mohand Talanana, both from Shell, for their critical reading of the
manuscript. We thank Shell for sponsoring this study.

NOMENCLATURE:
= Porosity
n = Bubble density
w
S = Saturation of liquid phase (water or surfactant
solution)
f
S = Saturation of foam
t = Time
D = Bubble diffusion coefficient
g
n = Number of generated bubbles
u = Darcy velocity
g
K = Bubble generation coefficient
n

= Maximum bubble density


d
K = Bubble destruction (decay) coefficient
= Density (of each phase)
q% = Mass depletion or production per unit volume per unit
time
= Mobility (of phases)
P = Total pressure (external and internal)
p = Pressure
= External pressure
C
P = Capillary pressure
k = Permeability
= Viscosity (of phases)
r
k = Relative permeability
c = Isothermal compressibility
wc
S = Connate water saturation
0
r
k = End-point permeability
B = Formation volume factor
M= General matrices of discretized flow equations
t = Time step size
Q= Injection or production rate in a grid
WI = Well index
wf
p = Flowing Bottom-hole pressure
DT = Dimensionless time
PV = Number of pore volumes
SC = Standard condition
Subscripts:
i = phase (i.e. aqueous or gas-foam)
j = number of the grid cell
Superscript:
n = Time step

SI Metric Conversion Factors
ft 3.048 E-01 = m
psia 6.894757 E+00= kPa
ft
3
2.831685 E-02 = m
3

Day 8.64 E+04= s
Darcy 9.869233 E-13 = m
2

cp 1.0 E-03 = Pa.s
SPEJ

References
1. Schramm, L.L. and Wassmuth, F., Foam: Basics Principle, in Foams: Fundamentals & Applications in the Petroleum Industry;
Scharamm, L.L. (ed), ACS Advances in Chemistry Series 242, Chapter 1, 1, 1994.
2. Rossen, W.R., Foams in Enhanced Oil Recovery, Chapter 12 in Foams: Theory, measurements and applications, Prud'homme,
R.K.and Khan, S. (Eds.), Marcel Dekker, 1996.
3. Kovscek, A. R. and Radke, C. J.: Fundamentals of foam transport in porous media, in Foams: Fundamentals and applications
in the Petroleum Industry, ACS Advances in Chemistry Series, N.242, American Society, 1994.
4. Patzek, T. W., Modelling of Foam Flow in Porous Media by the Population Balance Method, Surfactant based mobility
control progress in miscible-flood enhanced oil recovery, ACS Symposium Series 373, Washington DC, Chemical Society,
1988.
5. Hanssen, J. E. and Dalland, M., Gas blocking foams, in Foams: Fundamentals and applications in the petroleum industry,
Schramm, L. L. (Ed), ACS, 1994.
6. Heller, J. P. Critical Review of the foam rheology literature, Industrial & Engineering Chemistry Research, 26, 318, 1987.
SPE 113358 17
7. Patzek, T.W. and Koinis, M.T.: Kern River Steam-Foam Pilots, JPT, 42(4), 496-503, 1990.
8. Mamun, C.K., Rong, J.G., Kam, S.I., Liljestrand, H.M., and Rossen, W.R. : Extending Foam Technology from Improved Oil
Recovery to Envioremental Remediation, paper SPE 77557, Annual Technical Conference, San Antoni, USA, 2002.
9. Islam, M.R. and Farouq Ali, S.M. , Numerical Simulation of Foam Flow in Porous Media, JCPT, 29, 47, 1990.
10. Zhou, Z.H. and Rossen, W.R.: Applying Fractional-Flow Theory to Foams for Diversion in Matrix Acidizing, SPE Production
& Facilities, 1994.
11. Rossen, W.R. and Zhou, Z.H., Modelling Foam Mobility at the Limiting Capillary Pressure, SPE Advanced Technologies
Series, 3, 1, 1995.
12. Patzek, T.W., Modelling of Foam Flow in Porous Media by the Population Balance Method, Surfactant-Based Mobility
Control, ACS Symposium Series 373, Washington, D.C., Chapter 16, 1998.
13. Kovscek, A.R., Patzek, T.W., and Radke, C.J., Mechanistic Prediction of Foam Displacement in Multidimensions: A
Population-Balance Approach, SPE J., 71, 1994.
14. Kovscek, A.R., Patzek, T.W., and Radke, C.J., Mechanistic Foam Flow Simulation in Heterogeneous and Multidimensional
Porous Media, SPE J., 2, 511, 1997
15. Falls, A.H., Hirasaki, G.J., Patzek, T.W., Gauglitz, D.A., Miller, D.D., Ratulowski, T., Development of a Mechanistic Foam
Simulator: The population Balance and Generation by Snap-off, SPE Reservoir Engineering (August), 884, 1988.
16. Rossen, W.R. and Gauglitz, P.A.: Percolation Theory of Creation and Mobilization of Foams in Porous Media, AIchE J., 40, 6,
1994.
17. Zitha, P.L.J., A New Stochastic Bubble Population Model for Foam in Porous Media, SPE 98976, 2006.
18. Farshbaf Zinati, F., Farajzadeh R., and Zitha, P. L. J., Modelling and CT-scan Study of the Effect of Core Heterogeneity on
Foam Flow for Acid Diversion, SPE 107790, 2007.
19. Carretero-Carralero, M.D., Farajzadeh, R., Du, D.X. and Zitha, P.L.J., Modeling and CT-Scan Study of Foams for Acid
Diversion, SPE 107795, 2007.
20. Khatib, Z.I., Hirasaki, G.J. and Falls, A.H.: Effects of Capillary Pressure on Coalescence and Phase Mobilities in Foams
Flowing Through Porous Media, SPERE, 3(3), 919-926, 1988.
21. Hirasaki, G.J., Lawson, J.B., Mechanisms of foam flow in porous media: apparent viscosity in small capillaries, SPE J., 178,
1995.
22. Bretherton, F.P.: The Motion of Long Bubbles in Tubes, J. Fluid Mech., 10, 166-168, 1961.
23. Aziz, K., Settari, A., Petroleum Reservoir Simulation, Elsevier Applied Science Publishers LTD, 1979.
24. Stone, H.L., Iterative Solution of Implicit Approximations of Multi-dimensional Partial Differential Equations, SIAM J., 3,
530-558, 1968.
25. Stone, H.L. and Gardner, A.O., Analysis of Gas-cap or Dissolved Gas Reservoirs, Trans. SPE of AIME, 222, 92-104, 1961.

18 SPE 113358
Appendix 1: Equations
In this appendix we present the details of elements of matrices in equations (9-11). In the derivations to be followed, the
notations are adapted from the notations used in reference [23]. Discretized form of set of equations (19) for each block, in a
flow domain with uniform grid cells, can be written as follow:

( )
11 12 w w t w t w w
T p c S c p Q = + + (A.1)
( ) ( )
21 22 22 f w f C t w t w t C f
T p T P c S c p c P Q + = + + + (A.2)
( ) ( ) ( )
3
1
f w f C t w G
nT p nT P c n S K + = +

(A.3)

where:

11 12 1
,
n
w w n
w
x y z x y z
c c S b
t tB

+

= =


( )
21 22 1
, 1
n
w f n
f
x y z x y z
c c S b
t tB

+

= =


( )
1
3
, 1 ( )
n n
G w g
x y z
c K x y z S K n n
t

+


= =


,
w w f f
Q x y zq Q x y zq = =

The symbol is a difference operator, rather than Laplacian or etc. Therefore:

( ) ( ) ( )
( ) ( )
1 1
2 2
1 1
2 2
1, , 1, , , ,
, 1 , , 1 , , ,
i j i j i j i j i j i j
i j i j i j i j i j i j
T p TX p p TX p p
TY p p TY p p
+ +
+ +
+
+ +


where TX and TY are respectively x-direction and y-direction transmissibilities, for a relevant phase. IF i represents number
of the column and j number of the row of a grid block we can express the transmissibilities as follow:

1 1
2 2
1 1
2 2
1 1
2 2
, ,
, ,
, ,
,
i j i j
i j i j
i j i j
y z x z
TX TY
xB yB

+ +
+ +
+ +

= =



Subscripts
1
2
i + indicates upstream weighing of the parameter between grids ( ) , i j and( ) 1, i j + ; and
1
2
j + refers to upstream
weighing between grids ( ) , i j and ( ) , 1 i j + . Also one should note that:

1 1
,
n n n n
t w w w t w w w
S S S p p p
+ +
= =

where
t
is difference operator for time derivative and superscript n denotes the time step. By applying IMPES method to set
of equations (A.1-3), i.e. multiplying equation (A.2) by
11
21
c
c
and adding to the equation (A.1) and eliminating the term
22 t C
c P ,
we get:

( ) ( ) ( )
1 1 11 11 11 11
12 22
21 21 21 21
n n n
w w f w f C t w w f
c c c c
T p T p T p c c p Q Q
c c c c
+ +

+ + = + + +


(A.4)
( )
1 12
11 11 11
1
n w
t w w w t w
Q c
S T p p
c c c
+
= (A.5)
( ) ( ) ( )
1 1
3 3
1
1
n n G
t w f w f C
K
n S nT p nT P
c c
+ +
= +

(A.6)

SPE 113358 19
Rearranging and solving for unknowns of current time step leads to:

( ) ( ) ( )
1 1 1 11 11 11 11
12 22 12 22
21 21 21 21
11
21
n n n n n
w w f w w w f C
w f
c c c c
T p T p c c p c c p T p
c c c c
c
Q Q
c
+ + +

+ + = +



+ +


(A.7)
( )
1 1 1 12 12
11 11 11 11
1
n n n n n w
w w w w w w
Q c c
S T p p p S
c c c c
+ + +
= + + (A.8)
( )
( )
( )
( )
( )
( )
( )
1 1 1
1 1 1 1
3 3 3
1
1 1 1
1 1 1 1
n
w
n n n n
f w f C G
n n n n
w w w w
S
n nT p nT P K n
c S c S c S S
+ + +
+ + + +

= + +

(A.9)

The set of equations (A.7), (A.8) and (A.9) is equivalent respectively to the set of equations (9), (10) and (11) in the main text.

You might also like