Vous êtes sur la page 1sur 14

SPE 153616

Streamlines Simulation of Polymer Slugs Injection in Petroleum Reservoirs


B. J. Vicente, V. I. Priimenko, A. P. Pires, SPE, North Fluminense State University

Copyright 2012, Society of Petroleum Engineers

This paper was prepared for presentation at the SPE Latin American and Caribbean Petroleum Engineering Conference held in Mexico City, Mexico, 1618 April 2012.

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
The injection of polymer solutions is an effective method of Enhanced Oil Recovery. However, the continuous injection may
be very expensive, and the injection of slugs is an alternative to improve the recovery factor. In this paper, streamlines
simulation is used to model the 2D 2-phase oil displacement by polymer slugs, considering adsorption effects.
Streamline simulation can be much faster than conventional finite difference models and its results may be used for slug
polymer injection optimization in heterogeneous reservoirs. It decouples the 2D problem into multiple 1D problems along
streamlines. The mass transport equations in the streamlines are solved analytically. The continuous polymer injection is a
Riemann problem and its solution is self-similar, but in the case of slug injection, there are interactions between waves of
different families. We analyze the results for several adsorption isotherms, like Henry and Langmuir.
Analytical solutions are free from numerical diffusion, and allow the choice of greater time steps. The simulations were run
in 2D incompressible models without gravity effects. Results obtained were compared with finite difference simulation. The
analytical solutions along streamlines allowed the exact modeling of shock and rarefaction waves as well as the interaction
between waves of same and different families. Sensitivity analysis was performed regarding the technique adopted for the cell
concentration and saturation determination between time steps.

Introduction
Waterflooding is the most widely used method for secondary oil recovery by petroleum industry. This technique became more
important in the 1950s, period in which their main weaknesses and strengths were indentified. Waterfloonding at unfavorable
mobility ratio or in strongly heterogeneous reservoirs does not show satisfactory results. Polymer flooding is an alternative to
improve oil recovery in such scenarios (Littmam 1988; Sorbie 1991).The polymer increases the aqueous-phase viscosity and
may, in addition, decrease the effective permeability to the aqueous phase. Thus, there is a greater area swept by water and a
delay in water breakthrough. The main drawback of a polymer injection project is the price of the polymer; this problem may
be overcome through the use of a polymer slug driven by water.
A fast and robust full field scale simulator that includes a polymer flooding option is of fundamental importance for
preliminary evaluation of different secondary and improved oil recovery options for a given reservoir. During the simulation
of slug polymer flooding, an important issue to be considered is the adsorption of the chemical component in porous media.
Depending on the adsorption isotherm different zones with varying concentration of polymer both ahead and behind of the
slug may appear. Since certain polymer properties depend on concentration, e.g., the solution viscosity, simulation models
should take in account this effect.
Traditionally, such simulations are performed through finite differences based numerical models, well suited to represent
the complex physics of the problem, but that lead to results with numerical diffusion. Streamline simulation is an alternative to
overcome the limitations presented by conventional simulators. This techinique decouples the 2D problem into multiple 1D
problems along streamlines. Around 30 years ago, Lake et al. (1981) introduced a hybrid streamline approach to model large-
scale micellar-polymer flooding combining an areal streamtube model with a cross-sectional finite-difference simulator.
Batycky et al. (1997) presented a new streamline simulator applicable to field scale flow. The proposed method is three
dimensional and accounts for changing well conditions, heterogeneity, mobility effects, and gravity effects. The changing in
the mobility field is accounted for when updating the streamline, and in this work the transport equation was solved
numerically. Thiele et al. (2010) extended this approach to field scale polymer flooding. Clemens et al. (2010) used streamline
simulation to efficiently manage a field polymer injection project. A new metric was introduced: the polymer injection
2 SPE 153616

efficiency as a function of time for each pattern. AlSofi et al. (2009, 2010) performed an analysis to streamline polymer
flooding considering both Newtonian and non-Newtonian behavior.
In our approach the equations of transport along a streamline are solved analytically. Fayers and Perrine (1959) were
among the first to analyze solutions of hyperbolic systems modeling continuous polymer injection. Exact analytical solutions
were obtained for continuous chemical flooding with one dissolved component (Fayers 1962; Claridge and Bondor 1974),
with two dissolved components (Braginskaya and Entov 1980) and by any arbitrary number of components (Johansen and
Winther 1989). Bedrikovetsky (1982) described the hydrodynamics of oil displacement by injection of polymer slugs for
different adsorption isotherms.

Mathematical Model Description


The mathematical model is a simplified water/oil two-phase flowing in a porous media with the following assumptions:
Two phase flow,
The fluid flow obeys Darcys law,
Incompressible flow,
Gravity/capillary efects neglected,
Component diffusions and dispersions are neglible,
No polymer dispersion.
Polymer adsorption at thermodynamics equilibrium and its adsorption is a function of polymer concentration.
Local equilibrium exists everywhere.

General Equations
Following the model assumptions, the pressure equation is given by

( o + w ) K P = qt (1)

where K is the permeability tensor, qt a source or sink volumetric flow rate, o = kro o1 and w = krw w1 are the oil and
water mobility respectively, krl and l are the relative permeability and viscosity of the corresponding phase.
The transport process is described by the following quasilinear equations:

s
+ u f = 0
t
(2)
( cs + a )
+ u (cf ) = 0
t

where s ( x, t ) is the water saturation and c ( x, t ) is the polymer concentration in the water phase. The adsorbed
concentration is given by a ( c ) , is the porosity of porous media, and the water fractional flow is defined by

f ( s, c ) = w ( w + o )
1
(3)

The total velocity (u) is calculated using Darcys law:

u = ( o + w ) K P (4)

Coordinate Transformation
The basis for any streamline simulation method is a sequential splitting of the coupled pressure and saturation equations. The
pressure and velocity fields are then used as parameters while advancing the transport equations a given time step. Finally, the
new saturation field is used as input parameters for a new pressure solution step, and so on.
In reservoir simulation, the most important streamline parameter () is called time-of-flight, since it can be interpreted as
the travel time of a neutral particle along the streamline (Datta-Gupta and King, 2007):
SPE 153616 3


= d (5)
|u|

Together with the bi-stream functions () and (), for which u=, the time-of-flight form an alternative set of
coordinates for 3D space. The Jacobian of the transformation from physical coordinates to time-of-flight coordinates (,,) is
equal to . From this, and the fact that u is orthogonal to and, allow us to simplify the directional gradient along u as
follows


u = (6)

This operator identity is a key point in any streamline method, allowing 3D transport (Eqs. 2) to be transformed to a family
of 1D transport equations,

s f
+ =0
t
(7)
( cs + a ) (cf )
+ =0
t

where the functions s(,t) and c(,t) characterize the 1D flow along streamlines.

Statement of the Problem


The displacement of oil by a polymer slug driven by water is governed by Eqs. 7, and it is described by the following initial
boundary value problem:

f =1
s = sI
t = 0: and = 0 : cJ , 0 < t < t J (8)
c = 0 c = 0, t > t
J

where sI is the initial water saturation, cJ is the concentration of injected polymer and t J is the injection time.

Solution of the Problem


The solution of Eqs. 7-8 is discussed in details in many books, see, e.g., Rhee et al. (1986). Such problem may present two
kinds of rarefaction waves (Bedrikovetsky 1993):
s-wave - constant concentration and varying saturation;
c-wave - varying concentration and saturation.
The hyperbolic system Eqs. 7 also admits shock waves, satisfying the following Rankine-Hugoniot condition:

f+ f
V (t ) = + = (9)
s + [a ][c]1 s + [a ][c]1

where [ A] = A A , and A and A are the limiting values from the left and right at the point (t , 0 (t )) along any
+ +

discontinuity curve 0 (t ) , and V is the shock speed. In the case of continuos injection the solution is self-similar, i.e.,
s = s ( ), c = c( ), = t .
4 SPE 153616

Case 1. Linear adsorption isotherm. The linear adsorption isotherm may be represented by Henrys law:

a ( c ) = c (10)

where is the Henrys law constant. In this case the jump of concentration is defined by the discontinuity of the boundary
condition Eq. (8) at the point (0, t J ) . There is an interaction of the jump of concentration with the s-wave of the self-similar
solution for t > t J . The path of the discontinuity 0 (t ) is built using the following system of transcendental equations:

tJ tJ ( s + ( 0 ) , cJ )
= ( s ( 0 ) , cJ ) ,
+
=
0 ( t ) f s' ( s + ( 0 ) , cJ )
(11)
t

where ( s, c ) = f ( s, c ) ( s + b ) f s ( s, c ) , b = [a][c]1 .
'

Fig. 1 presents the solution for this case.

(a) ( s, f ) - plane (b) ( , t ) - plane

(c) Saturation - s ( , t ) (d) Concentration - c( , t )

Fig. 1 - Solution of the 1D problem, case 1.

The structure of the solution can be divided into four regions, see Table 1 for details. Velocities Vk , k = 0,1, 2,K are
calculated by the following expressions:

f ( s0 , 0 ) f ( s1 , 0 ) f ( s2 , cJ )
V0 = , V1 = f s' ( s1 , 0 ) , V2 = f s' ( s2 , cJ ) = 1
= (12)
s0 sI s1 + [a][c] s2 + [a][c]1
SPE 153616 5

Table 1: Structure of the solution, case 1


Area Concentration Saturation Description
I. > V0 t c ( , t ) = 0 s ( , t ) = s I Zone of displaced oil.
The corresponding path in ( s , f ) plane, represented by a part of curve
II. V1t < < V0 t c ( , t ) = 0 s 0 < s ( , t ) < s1
f ( s , 0) .

III. V2 t < < V1t c ( , t ) = 0 s ( , t ) = s 2 Oil bank.


Slug polymer .The corresponding path in ( s , f ) plane, represented by
+
IV. 0 ( t ) < < V2 t c ( , t ) = c J s ( 0 ) < s ( , t ) < s 2
a part of curve f ( s , c J ) .

Water-drive zone. The corresponding path in ( s , f ) plane, represented


V. 0 < < 0 (t ) c ( , t ) = 0 s J < s ( , t ) < s ( 0 )
by a part of curve f ( s , 0) .

Case 2. Convex adsorption isotherm. Langmuirs adsorption isotherm represents the convex case ( a (c ) < 0 ):
''

1c
a (c) = (13)
1 + 2c

where n , n = 1, 2 are the Langmuir constants. In this case, the discontinuity in (0, t J ) produces a c-wave that will iterate
with a s-wave of the self-similar solution. The discontinuity 0 (t ) is described by Eq. (11) with b = a ( cJ ) . For t > t1 (Fig.
'

2b) the trajectory of the front of polymer slug 1 satisfies the following transcendential equations:

d 1 f ( s ( 1 ) , c ( 1 ) ) f ( s + ( 1 ) , 0 )

= = + = f s' ( s ( 1 ) , c ( 1 ) )
dt s ( 1 ) + [a][c] 1
s ( 1 ) + [a][c] 1
(14)

1 ( t ) a ( c ( 1 ) ) a ' ( c ( 1 ) ) c ( 1 ) = cJ

The solution is constructed using the hodograph transform in the regions where the interaction between rarefaction waves
occurs. Thus, the path s J s2 s3 sJ in the ( s, f ) -plane transforms to t J s2 s3 t J in the ( , t ) -plane.
The velocity after the polymer slug is defined by the right-hand side of the equation

d 2 f ( s ( 2 ) , 0 )
= (15)
dt s ( 2 ) + a ' ( 0 )

where 2 is the position of the rear part of the slug.


Fig. 2 presents the solution of the problem in the case of the Langmuir isotherm.
The structure of the solution can be divided into seven regions, see Table 2 for details.
6 SPE 153616

(a) ( s, f ) - plane (b) ( , t ) - plane

(c) Saturation - s ( , t ) (d) Concentration - c( , t )


Fig. 2 - Solution of the 1D problem, case 2.

Table 2: Structure of the solution, case 2


Area Concentration Saturation Description
I. > V0 t c ( , t ) = 0 s ( , t ) = s Zone of displaced oil.
I

The corresponding path in ( s , f ) plane,


II. V1t < < V0 t c ( , t ) = 0 s 0 < s ( , t ) < s1
represented by a part of curve f ( s , 0) .

V2 t < < V1t , t < t1


III. c ( , t ) = 0 s ( , t ) = s 2 Oil bank.
+ f ( s5 , 0)(t t1 ) < < V1t , t > t1
1 '

Slug polymer. The corresponding path in ( s , f )


+
IV. 0 ( t ) < < V2 t c ( , t ) = c J s ( 0 ) < s ( , t ) < s 2
plane, represented by a part of curve f ( s , c ) .
J
+
2 ( t ) < < 0 ( t ), t < t1 0 < c ( , t ) < c J s ( 0 ) < s ( , t ) < s ( 2 ) Slug polymer. The corresponding path in ( s , f )
V. + plane, represented by a part of curve f ( s , c ) .
2 ( t ) < < 1 ( t ), t > t1 0 < c ( , t ) < c ( 1 ) s ( 1 ) < s ( , t ) < s ( 2 )

+ The corresponding path in ( s , f ) plane,


VI. 1 ( t ) < < + f ( s5 , 0 )(t t1 )
1 '
c ( , t ) = 0 s1 < s ( , t ) < s ( 1 )
s represented by a part of curve f ( s , 0) .
Water-drive zone. The corresponding path in

VII. 0 < < 2 ( t ) c ( , t ) = 0 s J < s ( , t ) < s ( 2 ) ( s , f ) plane, represented by a part of curve
f ( s , 0) .

Streamline Simulation Description


The simulator is based on an IMPES formulation, in which the pressure field is calculated using an implicite finite-difference
scheme. The water saturation and polymer concentration are calculated analytically along the streamlines. The streamlines,
launched from faces of a cell containing an injection well, are traced using Pollocks method (Pollock 1998).
The mapping of the saturation values determined from the analytical solution along the streamlines to the finite-difference
grid and calculation of the fractional flow of water in the producing well are performed using an approach proposed by
Batycky (1997). The concentration of polymer in the production well is made in a similar way.
SPE 153616 7

The saturation and concentration averages in a cell are calculated using a weighted average time-of-flight over the N
streamlines crossing the cell:

i ssl ( )i i csl ( )i
N N

scell = i =1
, ccell = i =1
(16)

N N
i =1
i i =1
i

where the saturation ( ssl ) and concentration ( csl ) averaged over the streamline i are given by:

1 ib 1 ib
ssl ( )i = s ( , t ) d , csl ( )i = c ( , t ) d (17)
i ia
i
i ia
i

Fig. 3 summarizes the main steps in streamline simulation.

Fig. 3-Flow Chart of Streamline Simulation used in this work.

Computational Results
In order to validate the accuracy of the streamlines simulator developed, the flooding of a 500 m x 500 m 1/4 five-spot pattern
reservoir was calculated with a 50 x 50 grid . The injection well was defined as rate constant (qJ = 30 m3/day) and the producer
as constant flowing pressure (P = 7000 kPa). For each adsorption isotherm (Figs. 4a and 4b) both homogeneous (kx = ky =
500 mD) and heterogeneous reservoirs (Fig. 4e) were analyzed. Relative permeability curves are calculated using the Corey
relations:

no nw
1 s sor s swc
kro = krowi , krw = krwor (18)
1 swi sor 1 swi sor

where sor is the residual oil saturation, swi is the irreducible water saturation, no and nw are constants of the Corey model,
and krowi and krwor are the endpoints of the relative permeability curves for oil and water, respectively. The curves of
fractional flow of water (Figs. 4c and 4d) are constructed using Eqs.18.
The relationship between the viscosity of the polymer solution and the concentration of polymer is assumed to be linear

R wI
w ( c ) = c + wI (19)
cR
8 SPE 153616

where R is the reference viscosity of the polymer solution at the reference concentration cR , and wI is the viscosity of
pure water. Table 3 presents other reservoir parameters used.

Table 3: Reservoir Parameters


Parameter R wI o cR tJ sI krowi krwor sor swc no nw
5 0.5 8 0.4 530
Value 0.2 0.5 0.7 0.2 0.2 0.2 2 2
(mPa.s) (mPa.s) (mPa.s) (Kg/m) (day)

(a) (b)

(c) (d)

Fig. 4-
(a) Henrys isotherm.
(b) Langmuirs isotherm.
(c) Water fractional flow for Henrys isotherm.
(d) Water fractional flow for Langmuirs isotherm.
(e) Permeability field for heterogeneous case.

(e)

The results were compared to a commercial finite-difference simulator (IMEX). First, the linear adsorption case with a
polymer injection concentration of 0.1 kg/m was run. Figs. 5 and 6 present the saturation and concentration maps,
respectively. For the case of adsorption governed by this isotherm the commercial simulator results show dispersion in the
vicinity of the polymer slug limits.
The saturation and concentration shocks are well modeled by the streamlines sinulation. Another important issue is the
numerical diffusion increase with time seen in the finite difference simulator. In Fig. 6 the smearing of the slug front also
caused by numerical effects of the commercial software may also be verified.
Fig. 7 compares the water fractional flow and the oil recovery factor for both simulators. The most important feature is the
shock dissipation in the finite difference results. In spite of these characteristics, both models are in good agreement, the
maximum deviation between the results in the revovery factor is about 4%.
Results for the heterogeneous case show a similar behavior (Figs. 8 and 9). The streamlines simulator was able to capture
the heterogeneity trends of the reservoir, while the commercial simulator numerical diffusion masked them. These differences
SPE 153616 9

in the saturation and concentration maps for the heterogeneous case are not so obvious when the water fractional flow and oil
recovery factor are observed (Fig. 10).

(a) 280 days (b) 280 days

(c) 720 days (d) 720 days


Fig. 5: Saturation map. Case 1 - homogeneous model and Henrys adsorption isotherm.

(a) 280 days (b) 280 days


10 SPE 153616

(c) 720 days (d) 720 days


Fig. 6: Concentration map. Case 1 - homogeneous model and Henrys adsorption isotherm.

Fig. 7: Water fractional flow and oil recovery factor. Case 1 - homogeneous model and Henrys adsorption isotherm.

(a) 280 days (b) 280 days


SPE 153616 11

(c) 720 days (d) 720 days


Fig. 8: Saturation map. Case 1 - heterogeneous model and Henrys adsorption isotherm

(a) 280 days (b) 280 days

(c) 720 days (d) 720 days


Fig. 9: Concentration map. Case 1 - heterogeneous model and Henrys adsorption isotherm
12 SPE 153616

Fig. 10: Water fractional flow and oil recovery factor. Case 1 - heterogeneous model and Henrys adsorption isotherm.

In the case of adsorption governed by Langmuirs isotherm the polymer injection concentration adopted was 0.4 kg/m. For
both cases of the isotherm the finite-difference method was not able to capture correctly the saturation and concentration
shocks (Figs. 11 and 12). The difference in the recovery factor was greater for this isotherm, getting close to 8%. Maps of
saturation and concentration present similar features as for the Hentys isotherm, and will not be presented.

Fig. 11: Water fractional flow and oil recovery factor. Case 2 - homogeneous model and Langmuirs adsorption isotherm.

Fig. 12: Water fractional flow and oil recovery factor. Case 2 - heterogeneous model and Langmuirs adsorption isotherm.

Another important advantage of the streamlines simulator over the conventional finite difference model is the number of
time steps necessary to obtain the pressure convergence (Table 4).
SPE 153616 13

Table 4: Number of Time Steps


This work IMEX
Homogeneous case 50 254
Henry isotherm
Heterogeneous case 50 263
Homogeneous case 50 206
Langmuir isotherm
Heterogeneous case 50 299

Conclusions
As a result of this study, the following conclusions can be derived:
1. A streamline simulator, capable to model the injection of polymer slugs for the Henry and Langmuir isotherms
has been developed. The main feature of this simulator is the use of analytical solutions of the transport equations
along the streamlines.
2. The results showed a good agreement compared to a commercial finite-difference simulator (IMEX), both in
homogeneous and heterogeneous cases.
3. Streamlines simulation allowed a smaller number of pressure solvers and analytical solutions have enabled a
better estimative of the saturation and concentration shocks and rarefaction waves.

Nomenclature
a =adsorption concentration of polymer on solid phase, m/L, Kg/m
c =concentration of polymer solution, m/L, Kg/m
c =average concentration, dimensionles
f =water fractional flow, dimensionless
f s =parcial derivative of the water fractional flow with respect to saturation
K =absolute permeability tensor, L, mD
kro =oil phase relative permeability, dimensionless
krowi =oil relative permeability at irreducible water saturation
krw =water phase relative permeability, dimensionless
krwor =water relative permeability at residual oil saturation
no =oil Corey exponent, dimensionless
nw =water Corey exponent, dimensionles
N =streamline number in a cell
P =pressure, M/TL, kPa
qJ =injection rate, L/t, m/d
qt =injection/production total rate, L/t, m/d
s =water saturation, dimensionless
sor =residual oil saturation, dimensionless
swi =irreducible water saturation, dimensionless
s =average saturation, dimensionless
t =time, days
u =total Darcy velocity, L/t, m/d
Vk =velocity of discontinuity k, dimensionles
=Henrys isotherm constant, dimensionless
1 =Langmuirs isotherm constant, dimensionless
2 =Langmuirs isotherm constant, L3/m, m3/kg
i =time-of-flight difference between the entry and exit in a cell for streamline i, t, days
=mobility, L3t/m, mD/Pa.s
o =oil viscosity, m/Lt, Pa.s
R =reference viscosity, m/Lt, Pa.s
w =polymer solution viscosity, m/Lt, Pa.s
wI =pure water viscosity, m/Lt, Pa.s
=coordinate along streamline, t, days
=self-similar variable, dimensionless
=time-of-flight, t, days
ia =time-of-flight of entry in a cell for streamline i, t, days
ib =time-of-flight of exit in a cell for streamline i, t, days
l =path of discontinuity l, t, days
=porosity, dimensionless
=streamfunction, L, m
=streamfunction, L2/t, m2/day
14 SPE 153616

Subscripts
cell = cell
I = initial
J =injection
o= oil
r =relative
R =reference
sl =streamline
w = water

Superscripts
+ =value ahead of discontinuity
- =value behind the discontinuity

Acknowledgements
The authors wish to thank the Maca Section of SPE for sponsoring the Onepetro license.

References
AlSofi, A.M., LaForce, T.C. and Blunt, M.J. 2009. Sweep Impairment Due to Polymers Shear Thinning, Paper SPE 120321-MS presented at
the SPE Middle East Oil and Gas Show and Conference, Bahrain, Bahrain, 15-18 March. http://dx.doi.org/10.2118/120321-MS.
AlSofi, A.M. and Blunt, M.J. 2010. Streamline-Based Simulation of Non-Newtonian Polymer Flooding. SPE Journal 15 (4): 895-905. SPE
123971-PA. http://dx.doi.org/10.2118/123971-PA.
Batycky, R.P. 1997. A Three-Dimensional Two-Phase Field Scale Streamline Simulator. PhD dissertation. Stanford University, Stanford,
California (January 1997).
Batycky, R.P., Blunt, M.J. and Thiele, M.R. 1997. A 3D Field-Scale Streamline-Based Reservoir Simulator, SPE Reservoir Engineering 12
(4): 246-254 SPE-36726-PA. http://dx.doi.org/10.2118/36726-PA.
Bedrikovetsky, P.G. 1982. Displacement of Oil by a Slug of an Active Additive Forced by Water through a Stratum. Fluid Dynamics 17 (3):
409-417. http://dx.doi.org/10.1007/BF01091279.
Bedrikovetsky, P.G. 1993. Mathematical Theory of Oil and Gas Recovery with Applications to ex-USSR Oil and Gas Fields. Kluwer
Academic Press.
Braginskaya, G.S. and Entov, V.M. 1980. Nonisothemal Displacement of Oil by a Solution of an Active Additive. Fluid Dynamics 15 (6):
873-880. http://dx.doi.org/10.1007/BF01096638.
Claridge, E.L. and Bondor, P.L. 1974. A Graphical Method for Calculating Linear Displacement with Mass Transfer and Continuously
Changing Mobilities. SPE Journal 14 (6): 609-618. SPE-4673. http://dx.doi.org/10.2118/4673-PA.
Clemens, T., Abdev, J. and Thiele, M.R. 2011. Improved Polymer-Flood Management Using Streamlines. SPE Reservoir Evaluation &
Engineering 14 (2): 171-181. SPE-132774-PA. http://dx.doi.org/10.2118/132774-PA.
Datta-Gupta, A. and King, M.J. 2007. Streamline Simulation: Theory and Practice, Vol.11. Richardson, Texas: Textbook Series SPE.
Fayers, F.J. and Perrine, R.L. 1958. Mathematical Description of Detergent Flooding in Oil Reservoirs. Paper 1132-G presented at SPE Fall
Meeting of the Society of Petroleum Engineers of AIME, Houston, Texas, 5-8 October. http://dx.doi.org/10.2118/1132-G.
Johansen, T. and Winther, R. 1989. The Riemann Solver for a Two-Phase Multicomponent Process. SIAM J. Math. Anal 20 (4): 908-929.
http://dx.doi.org/10.1137/0520061.
Lake, L.W., Johnston, J.R. and Stegemeier, G.L. 1981. Simulation and Performance Prediction of a Large-Scale Surfactant/Polymer Project.
SPE Journal 21 (6): 731-739. SPE 7471-PA. http://dx.doi.org/10.2118/7471-PA.
Littmann, W. 1980. Enhanced Oil Recovery. Prentice Hall.
Pollock, D.W. 1998. Semianalytical Computation of Path Lines for Finite-Difference Models. Ground Water 26 (6): 743-750.
http://dx.doi.org/10.1111/j.1745-6584.1988.tb00425.x.
Ree K.-H., Aris R. and Amundson, N.R. 1986. First-Order Partial Equations, Vol. 1, 2. Englewwod-Cliffs, New Jersey: Prentice-Hall.
Sorbie, K.S. 1991. Polymer-Improved Oil Recovery. Blackie.
Thiele, M.R., Batycky, R.P., Pollitzer, S. and Clemens, T. 2010. Polymer Flood Modeling Using Streamline-Part 1. Paper SPE 115545-MS
presented at the SPE Annual Technical Conference and Exhibition, Denver, Colorado, 21-24 September.
http://dx.doi.org/10.2118/16011-MS.

Vous aimerez peut-être aussi