You are on page 1of 10

JOURNAL

OF

Journal of Petroleum Science and Engineering 15 (1996) 281-290

Practical considerations in developing numerical simulators for


thermal recovery
J.H. Abou-Kassem
Chemical and Petroleum Engineering Department, UAE University, P.O. Box 17555, Al-Ain, United Arab Emirates

Received 17 March 1995; accepted 15 June 1995

Abstract

Numerical simulation of steam injection and in-situ combustion-based oil recovery processes is of great importance in
project design. Development of such numerical simulators is an on-going process, with improvements made as the process
description becomes more complete, and also as better methods are devised to resolve certain numerical difficulties. This
paper addresses some of the latter, and based on the author’s experience gives useful guidelines for developing more
efficient numerical simulators of steam injection and in-situ combustion. The paper takes up a series of questions related to
simulating thermal processes. Included are: the elimination of constraint equations at the matrix level, phase change, steam
injection rate, alternative treatments of heat loss, relative permeabilities and importance of hysteresis effects, improved
solutions to the grid orientation problem and other simulation problems such as potential inversion, grid block size, time-step
size control and induced fractures.
The points discussed in the paper should be of use to both simulator developers and users alike, and will lead to a better
understanding of simulation results.

1. Introduction The results obtained for the Fourth SPE Compara-


tive Solution Problem (Aziz et al., 1987) by six
Numerical simulation of thermal recovery meth- participants varied by as much as 15%, even though
ods-steam injection and in-situ combustion-is still the problem and the input data were tightly defined.
not a standardized procedure. Choices must be made These variations were in part owing to the differ-
at various stages in the implementation of a given ences in the numerical treatment used.
solution scheme. This paper considers several related
topics and offers suggestions based on the author’s
experience with thermal simulators. A previous pa- 2. Constraint equations
per (Abou-Kassem et al., 1986) gave a detailed
discussion of the various steam injection models A constraint equation in the present context is one
proposed, and provided a comparative view of the that does not involve any convection or conduction
treatment of nonlinearities and the solution schemes. terms. Thus, it is a function only of the unknowns of
These items will not be discussed here. Another the grid block, for which it is written, independent of
paper that gives useful “tips” for fully implicit the unknowns of neighboring grid blocks. Examples
simulation is the work of Au et al. (1980). of constraint equations are:

0920-4105/96/$15.00 Copyright 0 1996 Elsevier Science B.V. All rights reserved.


PIZ SO920-4105(96)000 12-5
1. Phase saturation constraint: The sum of saturation Consider a one-dimensional reservoir having three
of all phases in a grid block add up to one. This grid blocks (i.e., I= 1, 2, 3). The process being
constraint is encountered in all multiphase simula- modeled is steam injection. Each grid block 1 con-
tors. tributes ?I, constraints (e.g., n, = 3: cx, = 1. cy,
2. Phase composition constraints: The sum of mole
fractions of all components in a phase adds up to = 1 and S, + S,, + S, = 1) and n2 40, equations
one. These constraints are usually encountered in (e.g., n2 = 4: material balance equations for hydro-
compositional simulators. carbon components 1 and 2, material balance equa-
3. Thermodynamic equilibrium equations: These tion for the water component and energy balance
equations and phase composition constraints rep- equation). Let the unknowns of each grid block be
resent flash calculations in compositional simula- ordered as elements of subvector X,, (e.g., .Y,, x1
tors using equations of state. One form of these and S,) and subvector X,, (e.g., S,>, S,., T and p).
constraints is that fugacity of any component in All constraints and flow equations are written in
the system is the same in all phases. residual form and linearized using Newton’s itera-
The constraint equations and flow equations are tion. For each grid block the constraints are written
solved simultaneously in a fully implicit simulator first, followed by the flow equations. The resulting
using Newton’s iteration. However, the constraint linearized equations for the whole reservoir are pre-
equations can be eliminated at the matrix level with- sented in Fig. 1. Submatrices A, ,, and B, (, are
out influencing the results or the stability of the fully obtained by taking the derivatives of the constraints
implicit scheme. for block 1 with respect to the elements of Xc, and
This procedure was briefly referred to by Coats X,,, respectively. The derivatives of these constraints
(1980) implemented by Abou-Kassem and Aziz with respect to the unknowns of all blocks neighbor-
(1985) and was described by Naghiem and Rozon ing blocks 1 form zero submatrices of the sizes
(1989). Here the author presents the mechanics of 11, X n, and 11, X ~1~. The derivatives of the flow
the implementation of this powerful procedure that equationsf‘or block 1 with respect to the elements of
results in considerable reduction in both the CPU X,, and X,, form submatrices C,,, and D,,,. respec-
and memory storage requirements. tively. The derivatives of the flow equations for

n. Il_ Il. tl, Il. I-l”

"1

“2

xxx xxxxooooooo
"1 Q Q xxx xxxxooooooo
xxx xxxx 0000000
x x x xxxx xxx x x x x
xxx *xxx xxx x x x x
“2 c 2.1 !J 2.1 xxx *xxx xxx x xxx
xxx x xxx xxx x xxx

"1 Q Q Aa., EL,,

“2

Fig. 1. Coupled constraint and flow equations for one-dimensional reservoir consisting of three block\.
J.H. Abou-Kassem/ Journal of Petroleum Science and Engineering 15 (1996) 281-290 283

biock 1 with respect to the elements of x’,,_, and the flow equations of other grid blocks for which
X,!_, (for the neighboring block I - 1) form C,,, and block 1 appears as a neighboring block. For example,
D,,,, respectively; whereas those taken with respect for I= 2, these submatrices are C,,, and C,,,. The
to the unknowns of the other neighboring block row operations in this step are performed on the
1 + 1 form the corresponding C,,, and D,,,. corresponding submatrices D and subvectors of the
The steps involved in the elimination of the con- right-hand side (for I = 2, these submatrices are D,,,
straint equations at the matrix level are described and D,3,, and the subvectors are the intermediate
below. values of subvectors R,, and Z?rl resulting from row
Step I. Submatrix A,,, is transformed into an operations in both Step 2 and Step 3). This step is
identity matrix Z by performing row operations. executed for every grid block I= 1, 2, 3 in that
These row operations are at the sam_e time performed order. Once this is done, the constraints are decou-
on submatrix B,, and subvector R,,. This step re- pled from the flow equations. The decoupled system
sults in I, B;,. and R’,, shown in Fig. 2. This step is of equations are shown in Fig. 2.
performed for every grid block I= 1, 2, 3. Step 4. The 3n, decoupled flow equations (in this
Step 2. Each diagonal element of the resulting example) are solved simul$neo_usly for+the reduced
identity matrix Z in Step 1 is used to zero the subvectors of unknowns, X,,, XI, and XI,.
elements of the column under it of the n2 X n, Step 5. The unknowns of subvector Xc1 are solved
submatrix C, ,, (in Fig. 1) by row operations in a explicitly by back substitution using:
procedure similar to that in Gaussian elimination.
These row operations need to be performedonly on ?c,=l?c,-B;,o?r,; 1= 1,2,3 (1)
the n2 X n2 submatrix D,, and subvector R,, on the
right-hand side. This step is performed for every grid
block 1 = 1, 2, 3. This step results in the n2 X n,
zero submatrices and _O;,a shown in Fig. 2 and 3. Phase change
intermediate values of RrI (say Z?;).
Step 3. Each diagonal element of the identity Under certain conditions, the gas, water or oil
matrix Z resulting in Step 1 is used to zero the phase may disappear or reappear from regions of a
elements in that column, using row operations, of all reservoir undergoing steam injection or in-situ com-
submatrices that are contributions of grid block 1 to bustion. Three examples are given for conditions

.“1 n2 “1 "2

"1

"2

"1

"2

"1

"2

Fig. 2. Decoupled constraint and flow equations for one-dimensional reservoir consisting of three blocks.
284 J.H. Abou-Kassem/ Jourtml oj’Petr&um Science und Etqqinrering 15 (1996i281-290

under which phase change takes place: (i) the gas straint Cy, = 1 is not an appropriate equation for
phase does not exist in undersaturated reservoir, and
the block because the gas phase does not exist.
it may disappear in a saturated reservoir undergoing
Case d: Sp) = 0 and S:“’ > 0 imply that the oil
repressurization, forcing steam to condense and the
hydrocarbon gas to dissolve in the oil phase; (ii) a phase does not exist and hence physical equilibrium
is not attained in the grid block. The mole fractions
large pressure drop in the injection block may cause
of components in the gas phase will be the un-
the water phase to disappear (water evaporation),
knowns that will substitute for mole fractions in the
leading to the formation of a superheated steam and
oil phase in the formulation. In this case, the con-
the vaporization of hydrocarbon gas from the oil
phase; and (iii) the oil phase may disappear from a straint equations xxi = 1 must be replaced by the
grid block because of thermal cracking in in-situ equation S, + SE L 1 (or S,, = 0). In addition, the
combustion. Therefore, proper handling of phase
gas phase composition constraint cy, = 1, is still
change in a thermal simulator is important. Two of
the methods for handling phase change are the Vari- valid in the new formulation, and v,’ is an unknown
able Substitution Method and the Pseudo-equilibrium (L’, z K,s,).
Ratio Method.
3.2. Pseudo-equilibrium rutio method
3.1. Variable substitution method
This method was introduced by Crookston et al.
This method has been discussed by Coats (1978, (1979) and further elaborated on by Abou-Kassem
1980) and Abou-Kassem (198 1). In this method, the and Aziz (1985). In this method, the disappearance
disappearance and reappearance of either the oil, and reappearance of either the oil, water or gas phase
water or gas can be handled rigorously. The main can be handled without resorting to the variable
idea is to identify the condition in a given grid block, substitution logic in the numerical solution procedure
and then determine the unknowns and the appropri- of a thermal recovery process through the use of
ate constraint equations to be used in the grid block pseudo-equilibrium ratios. These ratios are expressed
during Newton’s iteration. Any grid block in a reser- as:
voir undergoing thermal recovery will exhibit one of
!?,X,,X,; i= 1,2, . . ..Iz.
the following cases: K, = (2)
Cuse a: Sp) > 0 and SC) > 0 imply that liquid k,X,.X,; i=n,+ 1
water is in equilibrium with steam in the gas phase;
and, therefore, temperature is not an unknown since where:
it is a function of pressure.
Case b: Sp) = 0 and Sr’ > 0 imply that the water (3)
phase does not exist and the block has superheated
steam. In this case, the block temperature, which is
(4)
independent of pressure, is the unknown that will
substitute for water saturation in the formulation. s,+ 4
One must use the equation S,, + S, = 1 (or S, = 0) x, = s, + 1o-1”
as one of the constraint equations for the block.
Case c: $“) = 0 and SC) > 0 imply that the gas and E,, E, and ~~ are small numbers of the order of
phase and hence steam does not exist in the block IO-‘. The functions X,, X, and X, are nearly 1.O,
and the temperature, which is independent of pres- except when S,,, S, or S, becomes smaller than
sure, is the unknown that will substitute for gas 10p3, i.e., near phase disappearance. If Xg is set to
saturation in the formulation. Also, one must use the 1, then the above pseudo-equilibrium ratios reduce to
equation S,, + S, = 1 (or S, = 0) as a constraint those introduced by Crookston et al. (I 979). The use
equation for the block instead of c y, = 1; the con- of Eqs. (2)-(5) forces the system to act as if there
I were small amounts of non-volatile oil, non-volatile
J.H. Abou-Kassem/Journal of Petroleum Science and Engineering 15 (1996) 281-290 285

water and non-condensable gas. In effect, any phase condensation in well blocks on injectivity is pre-
is allowed to approach disappearance without achiev- sented. The rate of steam injection as predicted by
ing it. this scheme is always the maximum rate possible
The minute amount of non-condensable gas in the under the prevailing conditions in the well blocks,
pseudo-equilibrium ratio method affects the flow and it increases until the maximum steam injection
behavior of the gas phase, especially if gas hysteresis mass rate set by the user (I,) is reached. For simplic-
is not simulated. This deficiency is removed by ity, the case of a single-block well is treated here.
setting the relative permeability of the gas phase to Extension to multi-block wells is straightforward.
zero, when S, < Sgt. The value of Sa: can be found The mass rate of steam injection is affected by the
by accepting 1% deviation in Xg from 1, i.e., Sg: = prevailing conditions in the well block in the follow-
100~,, as shown by Sato (1987). ing manner. In the early stages of injection, steam
The success of the pseudo-equilibrium ratio condenses and the block pressure ( p,) as well as the
method depends on (i) an appropriate initialization of bottomhole flowing pressure (p,,) begin to build up
saturations that is consistent with the pseudo-equi- because in-situ fluids may have low mobility ini-
librium ratio equations, (ii) a fully implicit treatment tially. As steam continues to be injected at the
of the pseudo-equilibrium ratios, and (iii> an appro- specified rate Z,, the bottomhole pressure continues
priate reassignment of negative saturations encoun- to build up until it reaches a maximum value related
tered during iterations. The following equation is to the pressure of the generated wet steam (P,,~,~~).
recommended: Beyond this point in time the injected steam will
SC”’= y“E. only serve to replace voidage in the well block, the
P p p’ p=o,w,g (6) rate of which may be smaller than !,. As total
All saturations are checked at every iteration u. mobility improves, the steam injection rate starts to
Any negative saturation is assigned a positive value increase until it reaches Z, again, after which time
calculated from the above equation. This assigned the pwf needed to inject Z, drops below pwf,,,,
saturation value is used in the following iteration. because of continued improvement in the total mo-
bility.
3.3. Other methods To model the behavior of steam injection rate as
described above, the mass injection rate is viewed as
These methods have limited use, and they aim at a function controlled by two independent variables:
handling the gas phase change only through intro- steam availability (I,> and maximum well injectivity.
ducing a “penalty source” term in the inert gas Since these two variables are not related, this func-
equation (Forsyth et al., 1981), or generating a small tion is always determined by the limiting variable.
amount of an inert gas in the system through low The mass of injected steam according to steam
temperature cracking reaction (Grabowski et al., availability (option 1) is:
1979).
I, =Z, (7)
The pseudo-equilibrium ratio method has been the
most widely used technique to handle the phase whereas that according to well injectivity (option 2)
change in compositional simulators because it al- is:
ways uses one set of constraint equations, and its
implementation is straightforward compared with that Z2 = i p,,, (8)
for the variable substitution method which requires
where pmix is the density of wet steam at well block
complicated logic.
conditions, i.e., after the injected wet steam is con-
densed or flashed, and i is the volumetric injection
rate at the well block conditions. Therefore:
4. Steam injection rate

In this section, an implicit injection scheme that


takes into account the effect of steam flashing and
ered, however, if a simulator is used to simulate a
(“1 laboratory experiment.
In modeling heat loss in thermal simulation. it is
where X, is total mobility, G, is well geometric assumed that (i) the medium is impermeable and
factor, p, and p, are the density of water and steam hence heat conduction is the only active energy flow
at well block reservoir conditions. S, is the portion mechanism in the adjacent formations, (ii) the lateral
of the total gas saturation S, that is steam, given by: heat conduction in the adjacent formations has negli-
gible effect on temperature, and (iii) thermal rock
(12) properties are independent of pressure and tempera-
ture. These are the justifications for modeling heat
In practice, option 1 is usually chosen. The injec- loss as a source term in the finite difference equa-
tion scheme allows this option to be followed as long tion, and for using a linear heat conduction equation
as: in the vertical direction to describe the temperature
distribution in the adjacent formations.
‘2 = p,,,,G,h,( P,, -P,,) 2 1, (13)
There are four different methods for estimating
or so long as: heat loss. These are: the numerical method (Coats et
al., 1974); the semi-analytical method (Vinsome and
(‘4) Westerveld, 1980), the analytical method (Abou-
Kassem, 1981) and the analytical solution with nu-
For light and intermediate crude oil reservoirs, op- merical approximation to the convolution integral
tion 1 can be operational throughout the life of the (Souza et al., 1987) and others. The author’s experi-
steam injection project. Option 2 can be chosen from ence (Hansamuit et al., 1992) with these methods
the start, but it is automatically invoked when: leads us to conclude that the numerical method is
sensitive to grid size in the overburden. the semi-ana-
(‘5) lytical method is the simplest to program yet its
prediction is the least accurate, the analytical method
For heavy crude oil reservoirs, option 2 continues is the most expensive (in CPU time and storage
to be operational until adequate steam mobility is requirements), whereas the analytical-numerical
attained in the injection block. This scheme has to be method (Souza et al., 1987) combines both accuracy
implemented in a fully implicit fashion. For a and acceptable storage requirements, and therefore it
single-block well, if pwf in Eq. (9) and Eq. (13) is is recommended. It should be mentioned that the
treated explicitly at the old time level, the stability of temperature of a boundary block, which appears in
the scheme will not be seriously affected (Abou-Kas- the heat loss term, has to be treated implicitly to
sem, 1981). The time levels have been omitted in the avoid time step instability.
above equations, and could be the old or the new
times, depending on the solution scheme used.
6. Relative permeabilities

5. Heat loss Three-phase relative permeabilities are the


Achilles’ heel of reservoir simulation. This is more
The energy lost to surrounding formations influ- so for thermal simulation. Relative permeabilities
ences the efficiency of a thermal recovery process. measured in the laboratory are questionable in the
Heat loss across the base and cap rocks is significant light of the instability theory. Frequently, relative
because of the large area1 extent of reservoirs, through permeabilities become a history match parameter.
which heat loss takes place, whereas heat loss across The laboratory values of end points are still of value.
the reservoir lateral boundaries is usually and justifi- In thermal simulation, the temperature dependence of
ably ignored because of the limited vertical exten- relative permeability should be taken into account.
sion of the reservoirs. Such heat loss must be consid- The same can be said of hysteresis. Often, two-phase
J.H. Abou-Kassem / Journal of Petroleum Science and Engineering 15 (1996) 281-290 287

relative permeability data (oil-water and oil-gas) Their scheme introduces mobility values from nearby
are used in some variation of Stone’s model (Di- points that lie in the true upstream direction rather
etrich and Bondor, 1976; Aziz and Settari, 1979) to than a single grid line as is the case in ordinary
obtain three-phase relative permeabilities. Myhill one-point upstream weighting. Though the method is
(1980) has referred to the problem of obtaining low extendible to multi dimensions, it has a few short-
oil saturations behind the steam front using Stone’s comings. It is only first order correct, it does not
model. Once the end points are obtained, it is easy to easily extend to fully implicit simulators, and its
use the Naar-Wygal-Henderson equations (Naar et behavior with phase change is not known. The au-
al., 1961). Kasraie (1987) gives a modified version thor’s experience shows that the nine-point scheme,
of these equations, suggesting that one way of in- in two dimensions, alleviates the problem of grid
cluding temperature dependence is to make the end orientation. This is usually extended to an eleven-
points in the equations temperature dependent. Rela- point formulation in three dimensions (a more rigor-
tive permeability hysteresis should be included in ous treatment calls for a 27-point scheme). Abou-
simulation of cyclic steam stimulation. This can be Kassem (1981) has given a generalized extension of
done in several ways; a simple approach was dis- Yanosik and McCracken’s nine-point scheme
cussed by Bang (1984). Sato (1987) has discussed (Yanosik and McCracken, 1979), and has shown that
the role of relative permeabilities on thermal simula- for the cases investigated, the results for the parallel
tion. and diagonal grids were close. Coats and Modine
(1983) have given generalized equations for calculat-
ing nine-point transmissibilities for heterogeneous
7. Grid orientation reservoirs. Ostebo and Kazemi (1991) have proposed
a split difference operator scheme for IMPES simula-
Under certain situations, vastly different results tors which could be extended to fully implicit simu-
are obtained for a steam flood, depending on whether lators. This scheme produces identical results to the
the grid lines are parallel or diagonal to the line nine-point difference scheme and can be easily im-
joining an injector-producer pair. The steam break- plemented in simulators using five-point scheme.
through time in the latter case may be three times Only recently Brand et al. (1991) have carried out
that in the former. This effect may even be more linear stability analysis to the difference equations
severe in in-situ combustion simulation. Grid orienta- for immiscible, incompressible two-phase flow and
tion was first noticed in the displacement of oil by showed that grid orientation is a numerical manifes-
water at an unfavorable mobility ratio. Details of the tation of physical instability. They concluded that
grid orientation problem can be seen in the works of grid orientation cannot be overcome with grid refine-
Todd et al. (1972) and Aziz and Settari (1979). In ment, and they provided a technique for estimating
steamflood simulation, both mobility weighting and block size for which errors remain small and limited
discretization scheme affect grid orientation by local truncation error.
(Vinsome and Au, 1981; Potempa, 1985). Therefore
both accurate numerical procedure and correct mo-
bility weighting are needed to alleviate grid orienta-
tion (Potempa, 1985). Abou-Kassem and Aziz (1982) 8. Other simulation problems
have discussed three approaches for solving the grid
orientation problem: (i) use of a nine-point, as op- The preceding discussion touched on a few items
posed to five-point difference operator (for two di- that need attention in thermal simulation. Several
mensions); (ii) use of a two-point upstream mobility other problems may occur, where a clear solution is
instead of single point; and (iii) use of harmonic total still not available. Some of the difficulties are related
mobility average for calculation of flow across a to potential inversions that may occur for a variety of
boundary. Frauenthal et al. (1985) have proposed a reasons. The usual solution is to check the flow
generalized one-point upstream mobility weighting direction after each inner iteration. Coupled well-
scheme which reduces grid orientation significantly. bores are seldom used, but the errors resulting from
uncoupled wellbore and reservoir flow equations can many physical phenomena present in thermal pro-
be significant, as shown by Kasraie and Farouq Ali cesses still elude mathematical description.
(1988).
Grid block size can be serious impediment in
simulating processes where steep fronts occur. Steam 10. Notation
injection in very viscous oils and in-situ combustion
are examples of this type. In the latter case, the A, B, C. D submatrices in Fig. 1
problem is particularly serious. To obtain meaningful B’, D' submatrices in Fig. 2
results from the measured reaction kinetics, the grid I,, 12 steam injection rates for options
block size would have to be of the order of half-a- 1 and 2, respectively
meter (Chu, 1981). specified steam mass injection
Simulation of fractures induced by air or steam rate
injection is a complex problem. The proper approach index or volumetric injection rate
to use is to couple the geomechanics of the surround- at reservoir conditions
ing rocks with the reservoir heat and fluid flow k,, K, physical equilibrium ratio and
equations (Tortike and Farouq Ali, 1987). At pre- pseudo-equilibrium ratio for
sent, while progress is being made in this direction, component i, respectively
the usual procedure is to presuppose the length, I index for block number in an
direction and orientation of the induced fracture. The ordering scheme
fracture width is normally much greater than the II old time level
actual width because of numerical considerations. An n,, 112 number of constraints and flow
example of such an approach is provided by the equations per block, respectively
work of Pethrick et al. (1986). If such a procedure is number of hydrocarbon compo-
employed, it is important to remember that the cus- nents in oil phase
tomary use of harmonic averages in simulation would block pressure
tend to distort the high permeability of the fracture. flowing BHP and maximum BHP
To minimize this effect, perhaps three blocks should of injection well
be used to represent the total width of the fracture. well geometric factor
This type of problem was noted by Huppler (1970). right-hand-side subvectors for
Time step selection is a matter of considerable con- block 1
cern in thermal simulation. Sammon and Rubin s,>S”> s,, s,, phase saturation for gas, oil,
(1983) discussed this aspect of reservoir simulation water and phase p, respectively
in details and developed a time step selection scheme imposed critical gas saturation
that was based on minimizing truncation errors. steam saturation
functions defined by Eqs. (3)-(S)
subvectors of unknown vector for
block 1
-\-, mole fraction of component i in
oil phase
9. Conclusions
?‘,. .v, mole fraction of component i
and steam in gas phase
Numerical simulation of thermal recovery meth- gas phase and steam compress-
ods is still in a state of flux. This paper considered a ibility factors
few selected topics that are of considerable impor-
Greek letters:
tance, for which guidelines were given from the
author’s experience. Still other difficulties remain, constants of the order of 1W5
which are addressed according to the preferences of total mobility of all phases in
the person doing the simulation. Apart from this, injection block
J.H. Abou-Kassem/ Journal of Petroleum Science and Engineering 15 (19961 281-290 289

density of wet-steam, water and Frauenthal, J.C., Di France, R.B. and Towler, B.F., 1985. Reduc-
Pmix3 Pw? Ps
tion of grid-orientation effects in reservoir simulation with
steam, respectively
generalized upstream weighting. Sot. Pet. Eng. J., 25(6):
u iteration number 902-908.
Grabowski, J.W., Vinsome, P.K., Lin, R.C., Behie, A. and Rubin,
B., 1979. A fully implicit general purpose finite difference
References thermal model for insitu combustion and steam. Sot. Pet. Eng.
54th Annu. Fall Meet., Las Vegas, NV, Pap. SPE 8396.
Abou-Kassem, J.H., 1981. Investigation of grid orientation in a Hansamuit, V., Abou-Kassem, J.H. and Farouq Ali, S.M., 1992.
two dimensional, compositional, three-phase steam model. Heat loss calculation in thermal simulation. Transp. Porous
Ph.D. Thesis, University of Calgary, Calgary, Alta., 287 pp. Media, 8: 149-166.
Abou-Kassem, J.H. and Aziz, K., 1982. Grid orientation during Huppler, J.D., 1970. Numerical investigation of the effects of core
steam displacement. 6th Sot. Pet. Eng. Symp. on Reservoir heterogeneities on waterflood relative permeabilities. Sot. Pet.
Simulation, New Orleans, LA, Pap. SPE 10497, 17 pp. Eng. J., lO(6): 381-392.
Abou-Kassem, J.H. and Azia, K., 1985. Handling of phase change Kasraie, M., 1987. Simulation of modified steam injection pro-
in thermal simulators. J. Pet. Technol., 37(10): 1661-1663. cesses applied to bottom water reservoir. Ph.D. Thesis, Uni-
Abou-Kassem, J.H., Farouq Ali, S.M. and Ferrer, J., 1986. Ap- versity of Alberta, Edmonton, Alta.
praisal of steamflood models. Rev. Tee. Ing., Univ. Zulia, Kasraie, M. and Farouq Ali, S.M., 1988. Effect of wellbore steam
9(2): 45-58. quality variation on steamflooding of thick formations. Calif.
Au, A.D.K., Behie, A., Rubin, B. and Vinsome, K., 1980. Tech- Reg. Meet. Sot. Pet. Eng., Long Beach, CA, Pap. SPE 17451.
niques for fully implicit reservoir simulation, Sot. Pet. Eng. Myhill, N.A., 1980. A check on numerical thermal simulation. 1st
55th Fall. Tech. Conf. Exhibit., Dallas, TX, Pap. SPE 9302. Sot. Pet. Eng./U.S. Dep. Energy Symp. on Evaluation of Oil
Aziz, K. and Settari, A., 1979. Petroleum Reservoir Simulation, Reservoirs, Tulsa, OK, Pap. SPE 8822.
Applied Science Publishers, London, 476 pp. Naar, J., Wygal, R.J. and Henderson, J.H., 1961. Three-phase
Aziz, K., Ramesh, A.B. and Woo, P.T., 1987. Fourth SPE com- imbibition relative permeability. Sot. Pet. Eng. J., l(6): 254-
parative solution project: Comparison of steam injection simu- 258.
lators. J. Pet. Technol., 39(12): 1576-1584. Naghiem, L. and Rozon, B., 1989. A unified and flexible ap-
Bang, H.W., 1984. Simulation study shows hysteresis effect on oil proach for handling and solving large systems of equations in
recovery during a cyclic steam process. Oil Gas J. (Feb. 271, reservoir simulation. In: K. Aziz and J.E. Heinemann (Editors),
82: 83-86. Int. Forum on Reservoir Simulation. Paul Steiner, Leoben.
Brand, C.W., Heinemann, J.E. and Aziz, K., 1991. The grid Ostebo, B. and Kazemi, H., 1991, Mixed five-point and nine-point
orientation effect in reservoir simulation. 11th Sot. Pet. Eng. finite-difference formulation of multiphase flow in petroleum
Symp. on Reservoir Simulation, Anaheim, CA, Pap. SPE reservoirs. 1 lth Sot. Pet. Eng. Symp. on Reservoir Simulation,
21228, 12 pp. Anaheim, CA, Pap. SPE 21227, 8 pp.
Chu, C., 1981. The curse of grid size. AOSTRA (Alta. Oil Sands Pethrick, W.D., Sennhauser, ES. and Harding, T.G., 1986. Nu-
Technol. Res. Auth.) Workshop on Computer Modeling. merical modeling optimization of cyclic steam stimulation in
Coats, K.H., 1978. A highly implicit steamflood model, Sot. Pet. cold lake oil sands. Pap. 37th Annu. Tech. Meet. Pet. Sot.,
Eng. J., 18(5): 269-283. CIM (Can. Inst. Min. Metall.), Calgary, Alta.
Coats, K.H., 1980. Insitu combustion model. Sot. Pet. Eng. J., Potempa, T., 1985. Mobility weighting in numerical reservoir
20(6): 533-554. simulation. Sot. Pet. Eng. J., 25(4): 565-572.
Coats, K.H. and Modine, D.A., 1983. A consistent method for Sammon, P.H. and Rubin, B., 1983. Practical control of time step
calculating transmissibilities in nine-point difference equa- selection in thermal simulation. 7th Sot. Pet. Eng. Reservoir
tions. 7th Reservoir Simulation Symp., San Francisco, CA, Simulation Symp., San Francisco, CA, Pap. SPE 12268.
Pap. SPE 12248. Sato, K., 1987. Sensitivity of steam displacement predictions to
Coats, K.H., George, W.D., Chu, C. and Marcum, B.D., 1974. three-phase relative permeability models, M.S. Thesis, Stan-
Three dimensional simulation of steamflooding. Trans. AIME ford University, Stanford, CA.
(Am. Inst. Min. Metall.), 257: 573-592. Souza, A.L., Pedrosa, Jr., O.A. and Marchesin, D.. 1987. A new
Crookston, R.B., Culham, W.E. and Chen, W.H., 1979. Numerical approach for calculating formation heat loss. 2nd Int. Symp.
simulation model for thermal recovery processes. Sot. Pet. for Enhanced Oil Recovery, Maracaibo, Pap. SRM 247.
Eng. J., 19(l): 37-58. Todd, M.R., O’Dell, P.M. and Hirasaki, G.J., 1972. Methods for
Dietrich, J.K. and Bondor, P.L., 1976. Three-phase oil relative increased accuracy in numerical reservoir simulators, Sot. Pet.
permeability models. Sot. Pet. Eng. 51st Annu. Meet., New Eng. J., 12(6): 515-530.
Orleans, LA, Pap. SPE 6044. Tortike, W.S. and Farouq Ali, S.M., 1987. Framework for multi-
Forsyth Jr., P., Rubin, B. and Vinsome, P.K.W., 198 1. The phase non-isothermal fluid flow in a deforming heavy oil
elimination of the constraint equation and modeling of prob- reservoir. 9th Sot. Pet. Eng. Symp. on Reservoir Simulation,
lems with a non-condensable gas in steam simulation, J. Can. San Antonio, TX, Pap. SPE 16030.
Pet. Technol., 20(4): 63-68. Vinsome, P.K.W. and Au, A.D.K., 1981. One approach to the grid
orientation problem in reservoir simulation. Sot. Pet. Eng. J., Yanosik, J.L. and McCracken. T.A., 1979. A nine-point finite
21(2): 160-161. difference reservoir simulator for realistic prediction of ad-
Vinsome. P.K.W. and Westerveld, A.. 1980. A simple method for verse mobility ratio displacements. Sot. Pet. Eng. .I.. 19(4):
predicting cap and ba\e rock heat losses in thermal simulators. X3-262.
J. Can. Pet. Technol.. l9(4): 87-90.

You might also like