Vous êtes sur la page 1sur 10

International Journal of Engineering & Technology IJET-IJENS Vol:16 No:03

24

Modeling and Simulation of Laboratory-Scale


Polymer Flooding
Ferreira, V. H. S. and Moreno, R. B. Z. L.
Abstract-- Polymer flooding enhanced oil recovery is
distinct from the other recovery methods for its ability of oil
production anticipation and water circulation reduction. These
enhancements are due to changes in the viscosity of the injected
fluid and the permeability of the porous media. Polymer flow
through porous media, however, is subject to complex
phenomena such as non-Newtonian viscosity, retention
mechanisms, and inaccessible pore volume. These phenomena
must be addressed in laboratory experiments which often use salt
tracers, although specific instrumentation and planning are
required to properly quantify them. The objective of this work is
the development of a cheap and easy to use computational tool to
assist experimentalists. This work develops a mathematical
modeling to one-phase flow of polymeric solutions in onedimensional porous media. The transport of fluid and
components (polymer and tracer) are modeled through three
coupled partial differential equations. Constitutive equations are
also present to model the coupled phenomena. A dimensionless
numerical solution is achieved through the finite differences
method and a fully implicit solution for pressure and
concentration behavior is proposed. The complete model is
presented and discussed to allow reproduction of the results. One
experiment aimed to estimate inaccessible pore volume,
retention, resistance factor and residual resistance factor is
selected from the literature for a history matching. The input
parameters directly taken from the experiments data were: fluid
rheology; fluid-rock interactions; rock sample geometry and
properties; and initial and boundary conditions. The only
parameter up for adjustment is the moment for injection startstop. The simulation results fitted the experimental
measurements nicely. The differences are discussed in terms of
validity and an enhanced analysis of the results is made possible
by the model. As of this works novelty, the presented model
proves itself as a simple and low-cost aiding tool for experimental
planning and analysis in single-phase one-dimensional polymer
flooding experiments in the presence of salt tracers.
Index Term- Enhanced Oil Recovery, Modeling, Numerical
Solution, Polymer Adsorption, Non-Newtonian Fluid.

I. INTRODUCTION1
The use of polymeric solutions for enhanced oil recovery
(EOR) is well known to improve oil recovery. The
incremental oil is a consequence of the reduction of the
mobility ratio of the injected fluid relative to the reservoir oil
through increase in viscosity and reduction of permeability to
the injected fluid [11], [26], [28].
The flow of polymeric solutions in porous media is subject
to some particular effects such as non-Newtonian flow,
retention and inaccessible pore volume (IAPV) that, among

others, are key to evaluating the success of a polymer flooding


project [28].
The determination of these parameters is achieved through
measurementofeffluents(polymerandoftentracer)profiles
in one-dimensional injection experiments [17]. These
experiments are often single-phase [17], [18], [24], but
sometimes also conducted at residual oil saturation [17], [18],
[22], [23]. Although, the presence of oil is reported to be of
minor impact in the measurements [20]. The data collected in
these experiments is invaluable to simulate field-scale
polymer EOR processes [23], which in turn are important to
estimate risk and profit. However, experiments of that nature
are of difficult instrumentation and planning that often leads
them to be based on similar reports available in the literature
or personal expertise.
That said, the objective of this paper is to present a
computer tool, developed by the authors in Matlab, that
simulates these laboratory experiments. This tool is capable of
giving invaluable insights to experimental planning such as: 1.
order of magnitude of the variables to be measured; 2.
variation rate expected to said variables; 3. time necessary to
reach steady-state flow regimen; 4. overall length of
experiments. More so, this tool can be used to determine
retention levels, IAPV and in-situ viscosity through history
matching.
It is intuitive to compare the proposed tool to commercial
reservoir simulators. These simulators can handle laboratoryscale simulation. However, the proposed tool offers
advantages over them from a researcher point of view. It does
not require the expensive acquisition of licenses, it allows
direct code modifications, and it provides a simple user
interface. These advantages issue an easy usage by users not
familiarized with computer simulations.
The objective is achieved through a dimensionless
mathematic model and numerical solution of a onedimensional single-phase flow of polymeric solutions and salt
tracers in porous media. This model takes into account
coupled effects of pressure and concentration and
nonlinearities such as viscosity and retention.
II. POLYMER FLOW IN POROUS MEDIA PHENOMENA
MODELING
Polymer transport in porous media is subject to many
effects that should be adequately modeled and included in the
transport equations. This section focuses on the modeling of
the phenomena considered in the model proposed by this

163503-8484-IJET-IJENS June 2016 IJENS

IJENS

International Journal of Engineering & Technology IJET-IJENS Vol:16 No:03

25

work. In this section, all equations are presented in their


dimensional form.
A. Viscosity of Polymeric Solutions Flowing through Porous
Media
Polymeric solutions behave as a non-Newtonian fluid. The
rheologic behavior of a polymeric solution is characterized by
a Newtonian plateau at low shear rates, followed by a shear
thinning (pseudoplastic) region. For rod-like polymers, such
as xanthan, the high shear rate region is characterized by
another Newtonian plateau (close to the solvent viscosity).
While for random coil polymers, such as polyacrylamide, a
shear-thickening (dilatant) region is observed in high shear
rates due to the viscoelastic nature of these polymers [28].
These rheological behaviors can be modeled in a plenitude
of ways [2], [13], [27], [28], with the simplest one being the
Ostwald-de Waele (power-law) model. This model, however,
can exhibit a poor representation of the rheological behavior
of polymeric solutions in the low and high shear rate regions,
which can be remedied by the introduction of two plateaus of
viscosities, as of ( 1 ).

(1)

This modified Ostwald-de Waele model brings it close to


the Carreau model A [6].
Fig. 1 exemplifies the match of this model with
experimental data, while comparing it with the Carreau model.
The experimental data was published by [22] and the same
author did the Carreau fit.
As only the Darcy (apparent or superficial) velocity can be
calculated by the pressure profile, the need for a way to
estimate the actual shear rate of the polymeric solution arises.
This relation can be estimated through several models [5], [8],
[21], [24]. Equation ( 2 ), presented by [8], is used in this
work.
(2)

Equation ( 2 ) is a modified version of the equation for the


shear rate at the wall of a capillary, which a derivation is
presented in [14]. A valueof=1,7is valid for packs of large
spheres with equal diameter [8], while the value of =2,5 is
considered appropriate for porous media with angular particles
[29].

Fig. 1. Viscosity curves for various polymer concentrations as modeled by


( 1 ), based on [22]

Another primary factor influencing the viscosity behavior


of polymeric solutions is the concentration of polymer in the
solution. The concentration of polymer in the injection front is
diffuse due to actual diffusion and retention and so the
viscosity behavior due to alterations in polymer concentration
must be addressed. In the present work this is achieved
through interpolation between the viscosity curves at known
concentrations.
B. Retention
The flow of polymer through porous media has a strong
relation to its retention. This phenomena is due to adsorption,
mechanical entrapment and hydrodynamic retention [28]. The
first one is considered irreversible and contributes the most to
the mass retention of polymer, the two later ones can be
reversed by changing the flow conditions but they play a much
minor part on mass retention if the solution is correctly
screened. So, polymer retention is often modeled as an
adsorption therm.
The retention of polymer physically changes the porous
medium, effectively reducing its permeability. In a waterpolymer-water injection experiment, this phenomena is
observed in laboratory as an increase in the pressure gradient
of the post-flush water relative to the pre-flush water. This is
named the residual resistance factor (RRF).

163503-8484-IJET-IJENS June 2016 IJENS

IJENS

International Journal of Engineering & Technology IJET-IJENS Vol:16 No:03


1) Adsorption: The work [3] have divided the isotherms for
physical adsorption into five classes. The isotherms for
adsorbents in which the pore size is not very much greater
than the molecular diameter of the sorbate, which is the case
for this work, are of type I [25]. So, various authors, such as
[1], [12], [19], and [28], model the adsorption as a Langmuir
isotherm ( 3 ), the simplest theoretical model for type I, singlelayer adsorption.
(3)

The dynamic of the adsorption can be modeled as


presented by [28]. Nevertheless, the dynamics of physical
adsorption at a surface is so fast that the mass transport always
dictates the overall adsorption rate [25].
2) Residual Resistance Factor: Polymer retention in porous
media results in a permanent reduction in its permeability
[28]. This effect is measured as the RRF. The RRF increases
as the molecular weight of the polymer increases although it
can be successfully modeled as a function of the locally
adsorbed amount through use of laboratory data. The model (
4 ) is used in this work [4].
(4)

C. Inaccessible Pore Volume


The paper [9] was the first to report a phenomenon that
polymer molecules are transported through the porous
medium faster than the much smaller tracer molecules, which
was named IAPV. This condition is reported as the result of
large polymer molecule size, forcing its transport through
larger pore throats and issuing a portion of the pore volume as
inaccessible [7], [10], [28].
(5)

III. MODEL
The modeling is composed of the phenomenological
equations presented in section 2 along with three partial
differential equations (PDE): a hydraulic diffusion equation
for pressure solution and two reaction-advection-dispersion
equations for concentrations of salt and polymer solutions. All
equations are solved in their dimensionless form and
discretized by finite difference method for numeric solution.

26

so a rheological study of said solution in the experimental


salinity and temperature conditions is invaluable for
correlating the impact of the polymer concentration and shear
rate in the viscosity of the fluid.
A. Dimensional Model
The hydraulic diffusion equation ( 6 ) is derived from the
fluid mass balance and considers a one-dimensional porous
medium completely saturated with one slightly compressible
fluid under Darcian flow and neglects the quadratic pressure
term.
(6)

The reaction-advection-dispersion of components ( 7 ) is


derived from the components mass balance and consider a
one-dimensional porous medium with the diffusion coefficient
considered constant in space.
(7)

The equation ( 7 ) is valid for both polymer and tracer.


Note that if the permeability of the medium, the viscosity
and the velocity of the fluid are invariants and there are no
reactions ( 6 ) and ( 7 ) are reduced to the classic diffusion and
advection-dispersion equations in the literature [15], [28].
B. Dimensionless Model
A dimensionless model has some advantages such as: 1.
enhancing the vision over key parameters relations; 2.
reduces the overall number of parameters on a system; 3. the
normalization of variables reduces the numerical
approximation errors related to operations between tiny and
huge numbers.
The dimensionless model is achieved by normalizing the
variables in regard to reference values, as summarized in
Table 1. This process results in the appearance of some
dimensionless numbers that indicates the relative importance
of some phenomena modeled, as presented in Table 2.
The dimensionless model arises by the introduction of the
variables in Table 1 in ( 6 ) and ( 7 ).
The dimensionless forms of the hydraulic diffusion and the
reaction-advection-dispersion of components equations are ( 8
) and ( 9 ) respectively.

Note that the proposed model only considers effects that


are dominant in laboratorial conditions. The user must have
the expertise to analyze the results when working under
conditions not modeled in this work, such as high temperature
(thermal degradation) or high flow rate (shear degradation).

(8)

(9)

The viscosity of the polymeric solution is especially


important for the correct history match of the pressure profile,

163503-8484-IJET-IJENS June 2016 IJENS

IJENS

International Journal of Engineering & Technology IJET-IJENS Vol:16 No:03

The retardation factor (FR) is a factor that slows-down the


transport of a component the higher the adsorption rate is. It
can be modeled as ( 10 ).

TABLE II
Dimensionless variables and reference values

Dimensionless
Number

( 10 )

Pclet Number

TABLE I
Dimensionless variables and reference values

Variable

Dimensionless
Form

Reference Value

Time

Length

27

Langmuir
Number

Description

Definition

Ratio between
advective and
diffusive transport
of component i
Ratio between rock
adsorptive capacity
and component i
concentration

Pressure
( 11 )

Permeability
a

Viscosity
Concentration
of component
i

( 12 )

( 13 )

Adsorption

( 14 )

Shear rate
Langmuir
equilibrium
constant
System
hydraulic
diffusivity
Volumetric
flow per area
unit
Mass flow
per area unit
for
component i

C. Numerical Model
A finite difference approach was used to derive a numeric
model. Spatial derivatives of diffusive nature were discretized
by second-order central differences, spatial derivatives of
advective nature were discretized by backward differences and
time derivatives were discretized by forward differences.
These choices were made to ensure model coherence and
minimize errors.

a Viscosity of the fluid that initially saturates the porous medium; b In case of successive
polymer injections the maximum inlet concentration is used as the reference.

Mass transport and constitutive equations also take part in


the model in their dimensionless form, that is: dimensionless
modified Ostwald-de Waele model ( 11 ), dimensionless
Langmuir model ( 12 ), dimensionless Darcyslaw(13 ), and
dimensionless mass transport of components per area unit
( 14 ).

The PDEs are coupled through fluid velocity, viscosity and


permeability, thus issuing a computationally heavy fully
implicit solution. The proposed solution scheme solves only
the pressure and concentrations implicitly while all the other
variables are evaluated explicitly. The solution flowchart,
implemented by the authors in Matlab , is described in Fig.
2.
The numeric forms of the dimensionless hydraulic
diffusion equation and the dimensionless reaction-advectiondispersion of components are ( 15 ) and ( 16 ), respectively.
By taking this approach to the problem, each PDE
constitutes a tridiagonal matrix for each time step and can be
solved by the Thomas algorithm.

163503-8484-IJET-IJENS June 2016 IJENS

IJENS

International Journal of Engineering & Technology IJET-IJENS Vol:16 No:03

( 15 )

28

( 16 )

Fig. 2. Flowchart of the problem solution, in the present case there are two components: polymer and tracer

163503-8484-IJET-IJENS June 2016 IJENS

IJENS

International Journal of Engineering & Technology IJET-IJENS Vol:16 No:03

The transport equations are also discretized using the same


approach as the PDEs. The numeric forms of the
dimensionless Darcys law and dimensionless mass transport
of components per area unit are ( 17 ) and ( 18 ), respectively.

29

TABLE III
Input data for history matching the experiment of [22]

Variable

Value

Core lenghta
( 17 )

Core cross-section areaa


Porositya,d

( 18 )

IV. EXPERIMENT HISTORY MATCHING


This section focuses on matching the experiment reported
in [22], on water-wet Bentheim sample to estimate IAPV,
retention, resistance factor (RF) and RRF using partially
hydrolyzed polyacrylamide (HPAM). The experiment
consisted of four sequential injections in an initially oil
saturated (water at swi) medium. Note that, once the residual
oil saturation is reached during the first injection, the behavior
should be identical to a single-phase system as no significant
oil production was observed in subsequent injections. The
input values are all based on [22], and their values are
presented in Table III.
The work [22] analyzed the in-situ viscosity of the
polymer for this core by performing a multi rate flooding. The
effective viscosity was estimated and shear rate was calculated
using equation ( 2 ). Comparing this study with the rheometer
measurements (Fig. 1), it seems that the solution lost a fraction
of its viscosity, as in Fig. 3.

Effective water
permeability @ sor a
Rock specific massb
System total volumetric
compressibilityc
Viscosity of the solventa
Diffusion coefficient of
polymer in free solutionc
Inaccessible pore volume
of polymera
Residual resistance
factora
Adsorption saturation
limit for polymera
Langmuir equilibrium
constant for polymerc
Low shear-rate plateau
for polymer viscosityb
High shear-rate plateau
for polymer viscosityb
Injection flow ratea
Production/Initial
pressurea
1st inj.
2nd inj.

Polymer
concentratio
n injecteda

3rd inj.
4th inj.

a Data given by [22]; b Calculated from data by [22]; c Based on [22] conditions;
d Effective porosity added to residual oil saturation.

This reduction can be explained by either depletion-layer


effects or deep bed filtration.
Fig. 3. Comparisson between the fit of viscosities for an HPAM (3630) at
400ppm, estimated by rheometry and in-situ by [22]

By Fig. 1 and Fig. 3 analysis, we are able to say that the


viscosity model proposed do not fit the viscosity in the high

163503-8484-IJET-IJENS June 2016 IJENS

IJENS

International Journal of Engineering & Technology IJET-IJENS Vol:16 No:03

shear rate region in this case, which is not a problem since the
in-situ shear rate for this experiment is 25.9s-1 (after RRF
effect). The viscosity model used for the experiment history
matching was the estimated in-situ viscosity.
That way, the only parameter free for adjustment is the
time for switching injections, as its not clear in the
experimental work. After running the simulation and adjusting
for cumulative polymer production, the results can be seen in
Fig. 4, 5, and 6. Note that [22] presented polymer
concentration data as pressure differentials on a capillary tube.

30

This data were converted to polymer mass, by the authors,


through usage of calibration curves provided in [22].
Regarding the pressure fit (Fig. 4), the first injection does
not quite adjust because the experiments core is saturated
with oil, but the other injections fit quite well. The polymer
concentration for the second injection (first polymer injection)
shows a steeper increase in the simulation. Also, the
simulation does not show the slight polymer production during
the third injection. However, even with those differences, the
concentration fits nicely.

Fig 7 PolymerandfictionaltracerproductionhistoriesinthesimulationofMoradis(2011)experiment

Since significant production of polymer during water postflush is highly unlikely, a discussion regarding the
inconsistencies between the simulation and experiment is
valid. The concentration measurements were made via a
capillary tube, that is, it is based solely on liquid viscosity.
Besides, although it is not reported any additional oil
production after the first injection, even a small quantity of oil
can easily contaminate this kind of measurement due to its
high viscosity, especially when compared with water. So even
though the pressure history did not suffered any significant
modifications, a quantity of oil considered negligible or
undetected may have contaminated the readings of the
polymer concentration. However, it is safe to say that this
effect is minor and do not invalidate the data nor the analysis.
Also, it is possible to observe the effects of polymer
retention and IAPV through injection of salt tracers along with
the polymer, as observed in Fig. 7. This procedure was not
performed in [22] and is presented here to illustrate the
modelscapabilitiesandenhancetheanalysis.

Fig. 4. Cumulative polymer production for simulation and experiment

163503-8484-IJET-IJENS June 2016 IJENS

IJENS

International Journal of Engineering & Technology IJET-IJENS Vol:16 No:03

31

diffusive one. The first polymer front is steeper than the


second due to retention.
As for the reference Langmuir number (NLA,pref=0.4126), it
is only able to say that the adsorption level is relevant for this
injected concentration, but with minor impact (N LA,pref<1, but
close to the unity).
V. CONCLUSIONS
This work proposed a mathematical model with coupled
effects of non-Newtonian viscosity, RRF, adsorption and
IAPV in the hydraulic diffusion and reaction-advectiondispersion of components equations. A numerical solution for
this model can be able to represent the flow of polymeric
solutions in porous media in the presence of tracers, under
laboratory conditions.

Fig. 5. Pressure differential between injection and production faces for


simulation and experiment

The model proposed was able to history match the pressure


and polymer concentration of an experiment through use of
experimental fluid and rock characterizations as input data.
This history match along with the dimensionless numbers also
allowed an enhanced analysis of the results.
The dimensionless reference numbers are available to the
user before the simulation is run, since they are calculated by
using reference properties. Through evaluation of these
numbers, the user can be able to qualitative analyze
characteristics of the polymer flooding such as the shape of
polymer front (Pclet number) and level of retention
(Langmuir number).
The history matching also showed the importance of the
in-situ rheology of a polymeric solution. Many differences are
observed between the rheology of the fluid measured in a
rheometer and estimated in-situ. If the rheometer data is used
without corrections, the matching of the experimental pressure
and polymer concentration histories by the simulator will be
unsuccessful.

Fig. 6. Polymer normalized concentration in the outlet face for simulation and
experiment

The first front of salt arrives at the production face earlier


than the polymer front due to the retardation nature of the
adsorption, dominant over the IAPV speed-enhancing effect.
The second front portrays the opposite, the polymer traverses
the porous medium faster than the salt due to the IAPV
effects, since all the adsorption was satisfied in the first
injection.
Also, the salt diffusion coefficient (considered to be that of
sodium chloride, Dss=4.8x10-4cm2/min, based on [16]) is much
higher than the polymer one (Dpp=3x10-6cm2/min, based on
[28]), so its expected the salt front to be more diffuse. But
this is not noticeable in the second front and the explanation is
given by analyzing the reference Pclet numbers. While the
Pclet number for the salt component (NPE,sref=4.2062x103) is
two orders of magnitude lower than the polymer one
(NPE,pref=7.2756x105), its absolute value is still very high,
indicating a dominance of the advective transport over the

The model and algorithm proposed in this work have high


potential for applications in experimental planning and
analysis as long as a fluid and rock characterizations are made.
ACKNOWLEDGEMENTS
The authors would like to thank the Brazilian National
Agency of Petroleum Natural Gas and Biofuel (ANP), PRHANP, CAPES, and Statoil for the financial support.

[1]

[2]

[3]

[4]

REFERENCES
ALUHWAL, O. K. H. Simulation Study of Improving Oil Recovery by
Polymer Flooding in a Malaysian Reservoir. Johor Bahru: Faculty of
Chemical and Natural Resources Engineering, Universiti Teknologi
Malaysia. (2008)
BALHOFF, M. Modeling the Flow of Non-Newtonian Fluids in
Packed Beds at the Pore Scale. Baton Rouge: Cain Department of
Chemical Engineering, Graduate Faculty, Agricultural and Mechanical
College, Louisiana State University. (2005)
BRUNAUER, S.; DEMING, L. S.; DEMING, W. E.; TELLER, E. On
a Theory of the van der Waals Adsorption of Gases. J Am Chem Soc
62 (7), 1723-1732. (1940). doi: 10.1021/ja01864a025
BONDOR, P. L., HIRASAKI, G. L., THAM, M. J. Mathematical
Simulation of Polymer Flooding in Complex Reservoirs. Soc Petrol
Eng J 12 (5), 369-382. (1972). doi: 10.2118/3524-PA

163503-8484-IJET-IJENS June 2016 IJENS

IJENS

International Journal of Engineering & Technology IJET-IJENS Vol:16 No:03

[5]

[6]
[7]

[8]

[9]

[10]

[11]

[12]

[13]

[14]

[15]
[16]

[17]

[18]

[19]

[20]

[21]

[22]

[23]

[24]

CANNELLA, W. J.; HUH, C.; SERIGHT, R. S. Prediction of Xanthan


Rheology in Porous Media. Presented at the SPE Annual Technical
Conference and Exhibition. Houston, Texas, USA. 2-5 October 1988.
doi: 10.2118/18089-MS
CARREAU, P. J. Rheological Equations from Molecular Network
Theories. J Rheol 16 (1), 99-127. (1972). doi: 10.1122/1.549276
CHAUVETEAU, G. Rodlike Polymer Solution Flow through Fine
Pores: influence of pore size on rheological behavior. J Rheol 26 (2),
111-142. (1982). doi: 10.1122/1.549660
CHAUVETEAU, G. ZAITOUN, A. Basic Rheological Behavior of
Xanthan Polysaccharide Solutions in Porous Media: effects of pore size
and polymer concentration. Proceedings of the third European
Symposium on Enhanced oil recovery, pp. 197-212. (1981)
DAWSON, R.; LANTZ, R. B. Inaccessible Pore Volume in Polymer
Flooding. Soc Petrol Eng J 12 (5), 448-452. (1972). doi: 10.2118/3522PA
DOMINGUEZ, J. G.; WILLHITE, G. P. Retention and Flow
Characteristics of Polymer Solutions in Porous Media. Soc Petrol Eng J
17 (2), 111-121. (1977). doi: 10.2118/5835-PA
GREEN, D. W.; WILLHITE, G. P. Enhanced Oil Recovery.
Richardson: Henry L. Doherty Memorial Fund of AIME, Society of
Petroleum Engineers. (1998)
HATZIGNATIOU, D. G.; NORRIS, U. L.; STAVLAND, A. CoreScale Simulation of Polymer Flow through Porous Media. J Petrol Sci
Eng 108, 137-150 (2013). doi: 10.1016/j.petrol.2013.01.001
HIRASAKI, G. J.; POPE, G. A. Analysis of Factors Influencing
Mobility and Adsorption in the Flow of Polymer Solution through
Porous Media. Soc Petrol Eng J 14 (4), 337-346. (1974). doi:
10.2118/4026-PA
JENNINGS, R. R.; ROGERS, J. H.; WEST, T. J. Factors Influencing
Mobility Control by Polymer Solutions. J Petrol Technol 23 (3), 391401. (1971). doi: 10.2118/2867-PA
LAKE, L. W. Enhanced Oil Recovery. Englewood Cliffs: Prentice Hall.
(1989)
LIDE, D. R. CRC Handbook of Chemistry and Physics, internet version
2005. Boca Raton, FL: CRC Press. (2005). Available at:
http://www.hbcpnetbase.com
LTSCH, T.; MULLER, T.; PUSCH, G. The Effect of Inaccessible
Pore Volume on Polymer Core Experiments. Presented at the SPE
Oilfield and Geothermal Chemistry Symposium. Phoenix, AZ. 9-11
March (1985). 10p. doi: 10.2118/13590-MS
LUND, T., BJRNESTAD, E. ., STAVLAND, A., GJVIKLI, N. B.,
FLETCHER, A. J. P., et al. Polymer Retention and Inaccessible Pore
Volume in North Sea Reservoir Material. J Petrol Sci Eng 7, 25-32.
(1992). doi: 10.1016/0920-4105(92)90005-L
NORRIS, U. L. Core-Scale Simulation of Polymer Flow through
Porous Media. Stavanger, Norway: Faculty of Science and Technology,
University of Stavanger. (2011)
MANICHAND, R. N; STAATSOLIE MAATSCHAPPIJ SURINAME
N. V.; SERIGHT, R. S. Field vs Laboratory Polymer Retention Values
for a Polymer Flood in the Tambaredjo Field. Presented at SPE
Improved Oil Recovery Symposium. Tulsa, Oklahoma, USA. 12-16
April (2014). 15p. doi: 10.2118/169027-MS
MASUDA, Y.; TANG, K.; MIYAZAWA, M.; TANAKA, S. 1D
Simulation of Polymer Flooding Including the Viscoelastic Effect of
Polymer Solution. SPE Reservoir Eng 7 (2), 247252. (1992). doi:
10.2118/19499-PA
MORADI, H. Experimental Investigation of Polymer Flow through
Water- and Oil-wet Porous Media. Faculty of Science and Technology,
University of Stavanger. (2011)
OSTERLOH, W.T.; LAW, E.J. Polymer Transport and Rheological
Properties for Polymer Flooding in the North Sea Captain Field.
Presented at SPE/DOE Improved Oil Recovery Symposium. Tulsa,
OK, 19-22 April (1998). 12p. doi: 10.2118/39694-MS
PANCHAROEN, M.; THIELE, M. R.; KOVSCEK, A. R. Inaccessible
Pore Volume of Associative Polymer Floods. In: SPE Improved Oil
Recovery Symposium. Tulsa, Oklahoma, USA, 24-28 April (2010).
15p.
doi:
10.2118/129910-MS

[25]
[26]
[27]

[28]
[29]

32

REUVERS, N.; GOLOMBOK, M. Shear Rate and Permeability in


Water Flooding. Transport Porous Med 79 (2), 249-253. (2009). doi:
10.1007/s11242-008-9313-x
RUTHVEN, D. M. Principles of Adsorption and Adsorption Processes.
New York: Wiley-Interscience Publication. (1984)
SHENG, J. J. Modern Chemical Enhanced Oil Recovery: theory and
practice. 1 ed. Kidlington: Gulf Professional Publishing. (2011)
SOCHI, T. Pore-Scale Modeling of Non-Newtonian Flow in Porous
Media. Department of Earth Science and Engineering, Imperial College
London (2007)
SORBIE, K. S. Polymer-Improved Oil Recovery. 1 ed. Boca Raton:
Blackie. (1991)
STAVLAND, A.; JONSBRATEN, H. C.; LOHNE, A.; MOEN, A.;
GISKE, N.H. Polymer Flooding - Flow Properties in Porous Media
Versus Rheological Parameters. Presented at the SPE
EUROPEC/EAGE Annual Conference and Exhibition. Barcelona,
Spain, 1417 June (2010). 15p. doi: 10.2118/131103-MS

163503-8484-IJET-IJENS June 2016 IJENS

IJENS

International Journal of Engineering & Technology IJET-IJENS Vol:16 No:03

List of Symbols
A
or

or

or

or

33

core cross-section area

backward coefficient in tridiagonal matrix for hydraulic diffusion or reaction advection


dispersion of component i

Langmuir equilibrium constant for component i

central coefficient in tridiagonal matrix for hydraulic diffusion or reaction advection


dispersion of component i

concentration of component i

forward coefficient in tridiagonal matrix for hydraulic diffusion or reaction advection


dispersion of component i

system total volumetric compressibility


diffusion coefficient of component i in free solution

independent coefficient for hydraulic diffusion or reaction advection dispersion of


component i

retardation factor
inaccessible pore volume

mass flow of component i per unit area

volumetric flow per unit area

flow consistency index

permeability of porous media

core length
flow behavior index
hydrostatic pressure
production or consumption of component i, negative for sink and positive for source
residual resistance factor
time
length

system hydraulic diffusivity

adsorbed quantity of solute (ratio between adsorbent mass and solute mass)

adsorption saturation limit for component i


shear rate
shear rate limit for Newtonian low shear rate plateau
shear rate limit for Newtonian high shear rate plateau

fluid dynamic viscosity

rock specific mass


porosity

polymer apparent porosity

Superscripts
-

refers to the injection face


refers to the initial time
current time-step
reference value
refers to the production face

Subscripts
-

163503-8484-IJET-IJENS June 2016 IJENS

dimensionless
component i (polymer or salt)
current space-step
polymer component
salt component

IJENS

Vous aimerez peut-être aussi