You are on page 1of 13

SPE 167525

Modelling Formation Damage Induced by Propagation of Single-Phase


Non-Newtonian Polymer Solution in Porous Media
P.E.G. Idahosa, SPE, G.F. Oluyemi, SPE, M. B. Oyeneyin, SPE and R. Prabhu. The Robert Gordon University,
Aberdeen, United Kingdom

Copyright 2013, Society of Petroleum Engineers

This paper was prepared for presentation at the Nigeria Annual International Conference and Exhibition held in Lagos, Nigeria, 30 July – 1 August 2013.

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
Non-Newtonian high molecular weight polymers dissolve in water to increase viscosity for stable displacement of oil and gas
by injected water and enhanced sweep recovery. In both microscopic and macroscopic description of geological factors
controlling delivery of oil and gas from reservoir to wellbore, and hence, total eventual production, permeability is a
dominant factor. Despite the role of polymers in enhancing oil recovery, polymer adsorption on rock surfaces can also cause
permeability reduction, and hence, oil and gas productivity decline due to flow restrictions (formation damage). In the
literature, studies on selective permeability reduction (with emphasis on mobility reduction and residual resistance factors) by
polymer adsorption based on constant absolute permeability assumption are well known. However, even then, these studies
are based on mathematical expressions and formulations that are subjective. There is therefore no existing model that fully
considers a combination of the effects of polymer adsorption and its non-Newtonian behaviour on absolute permeability
during single-phase propagation in porous media. In order to describe single-phase non-Newtonian flow in porous media, a
macroscopic flow rate/pressure drop relationship that combines a mathematical description of fluid rheology must be
developed. In this paper, a new method is developed in order to investigate and characterize absolute permeability alteration
due to the propagation of single-phase polymer solutions in porous rock media. The validity of the proposed model is
demonstrated by using experimental data sets involving single-phase polymer flooding with varying concentrations and
formation permeabilites.

Keywords:
Formation damage, polymer adsorption, porous media, single-phase flow, fluid rheology.

Introduction
It is a common knowledge that reservoir flow properties within sandstone and shale constituents drastically differ due to the
complexity and sequential discontinuities in such reservoir. A major objective of the study of reservoir engineering is the
optimal recovery of hydrocarbon under various conditions and probable economic constraints. Compared to base or
conventional water flooding and other chemical EOR processes, additional oil have been recovered with polymer flooding by
increasing sweep efficiency (Wang et al. 2007, 2001), but not without its negative impact of permeability plugging
(formation damage). Although with its formation damage potential (Annie et al. 1999; Carr and Yang 1998, Chauveteau et al.
2002), polymers have been reported to have several other important functions (Sun et al. 2012; Navarrete et al. 2001, 2000;
Seright 2011, Clark 2010, Wang et al. 2008, Sydansk and seright 2007). In geosciences (and petroleum engineering in
particular), absolute permeability, relative permeabilities, capillary pressures, and formation resistivities amongst others are
regarded as quantitative prediction of the continuum flow descriptors of porous media systems (Guodong et al. 2004, Civan
2002). This partly implies that oscillatory permeability fields are needed to describe the medium for a single-phase flow
problem. A measure of the ability of porous materials to conduct flow is permeability which is dictated by the geometry of
the pore network program. Relative-permeability is the ease with which a particular fluid flow in the pore channel in the
presence of other fluids and its data on the other hand, is used to characterise multiphase flow of fluid even though there are
still challenges in so doing. The absolute or total permeability alteration due to adsorption of polymer molecules on rock
medium surfaces is one of such challenges. Relative-permeability data is also useful for diagnosing potential formation
damage expected under various operational conditions as well as in making engineering estimates of productivity, injectivity,
and ultimate recovery from reservoirs for the purpose of planning and re-evaluating production activities. It is therefore
2 SPE 167525

required in reservoir simulation studies. However, because of the interference between fluids when they share the same flow
channels, the sum of the effective permeabilities for all phases is usually less than the absolute permeability (Darcy 1856,
Honarpour and Mahmood 1988). In order to fully understand this mechanism, it is therefore very fundamental to develop
better methods for understanding single-phase non-Newtonian polymer solution flow behaviour in porous media in relation
to absolute permeability as an important flow factor. This can then be extended for better investigation of two-phase (e.g.
polymer solution/oil) or other multiphase relative permeabilities.

Non-Newtonian high molecular weight polymers dissolve in water to increase viscosity for stable displacement of oil and gas
by injected water and enhanced sweep recovery (Cao et al. 2008, Liu and Liu 2007, Urbissinova et al. 2010, Sorbie 1991,
Lake 1989, Littman 1988). Polysaccharides (with shear-thinning property) and polyacrylamides (have both shear-thinning
and shear–thickening properties) are the two main polymers used for enhanced oil recovery (EOR) applications (Lake 1989,
Delshad et al. 2008, Wu et al 2007, Masuda et al. 1992, Maerker 1984, Heemskerk et al. 1984). Despite the role of polymers
in enhancing oil recovery, polymer adsorption on rock surfaces can also cause permeability reduction (Stavland 2010,
Chauveteau et al. 2002), and hence, oil and gas productivity decline (Carr and Yang 1998) due to flow restrictions (formation
damage). Zaitoun and Kohler (1988) conducted two-phase flow through porous media to study the effect of an adsorbed
polymer layer. They observed that the dilatant character of HPAM molecule in diverging flow caused a high permeability
reduction with shear rate. In another study, Zitha et al. (1995) demonstrated the wide use of polymers for near well-bore
conformance control treatments for the purpose of permeability contrast correction between layers. They concluded that
polymers had the ability to invade deep into high-permeability layers compared with low-permeability layers which in turn
enhances resistance to flow in the high-permeability watered-out zones. From the work of Chauveteau et al. (2002), the
thickness of an absorbed polymer layer depend more on shear rate rather than the injection rate, but did not model the effect.
The literature has therefore devoted much research efforts to quantify the effects and contribution of polymer adsorption.
Most of these studies focused on the use of polymers for water shut-off for enhanced oil recovery and relative permeability
treatments without giving much attention to the potential of polymers to damage the formation or cause restriction to flow of
hydrocarbon in the reservoir; much less in relation to absolute permeability impairments. Therefore, so far, no generalized
methods have been developed to deal with issues of single-phase non-Newtonian polymer flow and their associated potential
for formation damage.

This paper considers the shear-thinning properties of polymer solutions with which a new method is developed in order to
investigate, calculate and characterize absolute permeability alteration due to the propagation of single-phase polymer
solutions in porous rock media.

Darcy Flow
Darcy’s law, an empirical equation describing the Laminar flow of incompressible fluids, is largely used for calculation of
fluid flow through porous media. It relates the macroscopic velocity (flux) of a fluid of known viscosity to the pressure
gradient by a proportionality factor called absolute permeability, expressed in Darcies. Invariably, the work of Darcy (1856)
marks the foundation of all fluid flow studies in porous media. Since then, it has been generally accepted to use Darcy’s law
to describe or model the slow flow of Newtonian fluids in reservoirs and aquifers because the low-matrix permeability results
in low velocities. In reality, Darcy’s observations imply an apparent decrease in flow capacity with increasing velocity. In his
series of experiments on the flow of water through a column of sand-pack at various pressure differentials, Darcy (1856)
showed that the pressure gradient can be related to the flow rate of the fluid and the superficial velocity through a
permeability constant given by:

k A P
q (1)
 L

From equation (1), we can write:

q
k  (2)
A (P / L)

Where,
P
= pressure gradient
L
 = apparent fluid viscosity
k = apparent permeability (an intrinsic property of porous medium)
q = volumetric flow rate,
SPE 167525 3

A = cross-sectional area of medium


Coupled Equation for Non-Newtonian Flow Characterisation
Knowing the solution properties of a polymer solution is the first step in predicting its mobility and associated effects in a
permeable reservoir. The viscosity of polymer solutions in rock media can be characterized by the Power-law (Bird et al.
2002) and Carreau (1972) models amongst others. Probably due to its simplicity, the Power-law model is the most
encountered analytical form of viscosity-shear rate relationship which describes the flow of most polymer solutions in rock
media. However, it does not account for neither the lower nor upper Newtonian plateaus of the viscosity curve. For the
power-law model (applicable to Newtonian fluids), k and n (see equation 3) assume reasonable constant only over a narrow
interval of shear rate range. For non-Newtonian fluids such as polymers (whose behaviour is different from Newtonian),
viscosity is not constant for all rates of shear; therefore, the computation of pressure drop is not a straightforward task. The
Power-law (Bird et al. 2002) model which describes the non-Newtonian (or pseudoplastic) region is given by the expression:

 ( )  K  n 1 (3)

Where, K = consistency index (or power-law coefficient (cp.secn-1)),


n = flow index or power-law exponent (dimensionless constant ( 0 .4  n  1 )).

The Carreau (1972) model is written as (Cannella et al. 1988):

    ( 0p   )[1  (eff ) ]( n 1) /  (4)

  
Or  [1  (eff ) ]( n 1) /  (5)
 p  
0

Where,
 = apparent shear viscosity in porous media
 0p = polymer viscosity at zero shear rate
  =  w = viscosity at infinite shear rate
 = time constant (i.e. relaxation time for realignment of polymer rods in a shear flow field) is found from bulk viscosity
measurements
eff = rate of deformation; called effective (or in-situ) shear rate in single-phase shear flow.
n = dimensionless constant known as the shear-thinning index that depends on the concentration. 0 .4  n  1 for
pseudoplastic or shear-thinning fluids
 = shape or tuning parameter used to improve the viscosity match; and is usually taken to be 2 (Canella et al. 1988).
By substituting equation (4) into (2) gives:

k
q
A (P / L)

  (  0p   )[1  (eff ) 2 ]( n 1) / 2  (6)

Equation (6) simply implies that the apparent permeability is a function of shear rate. The same conclusion is implicit in
equations (1), and (2). The justification for using equation (4) instead of (3) in (2) is that the Carreau model covers and
combines the power-law region and the two Newtonian regions of the viscosity curve. Therefore, it has a better application
compared to the power-law model.

Equation (6), however, cannot be used to model or quantify absolute or apparent permeability alteration due to the flow of
polymer solution in porous media. This is because the best-fitting parameters of equation (4) used in (6) are viscometer-
measured data; and that the in-situ shear rate in the rock is not known. In the literature, however, models exist for converting
the viscometer-measured viscosity to apparent viscosity of polymer solution applicable in permeable porous medium. In this
study, we use the in-situ shear rate model for single-phase flow (Canella et al. 1988, wreath et al. 1990) given by:
4 SPE 167525

n /  n 1  u 
 3n  1
eff  C   w  (7)
 4n   kaq 

Where,
u w = Volumetric flux or Darcy velocity of the polymer-containing water phase,
k aq = water-phase permeability,
 = porosity
C = a correction factor or constant (called the shear rate coefficient) = 6.0 (fits very well a wide variety of coreflood data;
Canella et al. 1988).

The determination of the rock in-situ aqueous permeability ( k aq ) in equation (7) as an unknown is done by a different
method (i.e. calculated from water injectivity data just before polymer injection). Therefore, by using equation (7) in (6), the
absolute permeability ( k app ) alteration due to non-Newtonian polymer solution flow can be modelled or determined by the
following expression of equation (8):

 n 1
 
   n
 n 1   
2
 2


kapp 
q
A (P / L) 
    p  
0
  1   .C 
 
3n 1 
 

u w 

 
  (8)
 
 4n   kaq   
 

Anatically, for non-linear, non-Newtonian single-phase polymer solution propagation in porous media, equation (8) is simply
interpreted to mean that the rock absolute permeability damage will be higher or greater for lower aqueous permeability for
the same velocity. According to previous studies (Sun et al. 2012, Hirasaki and Pope, 1974; Mahdi et al. 2011; Denys et al.
2001; Zitha, 2001; Chauveteau et al. 1995), polymer molecules may adhere or adsorb on porous matrix pore walls and thus
cause a reduction in the aqueous or water-phase permeability. This implies that the apparent viscosity in porous media
predicted from bulk viscosity using equation (4) may be underestimated as earlier stated (Chauveteau 1984). Figure 1 is a
typical relationship between shear viscosity and polymer concentration; whilst Figure 2 shows the relationship between
normalized apparent viscosity and shear rates.

100
Viscosity (cp)

1500ppm
10 2000ppm
2500ppm

1
0.01 0.1 1 10 100 1000
Shear Rates (s‐1)

Figure 1: Polymer concentration and shear viscosity relationship (Lot# SP 0307* in 4%


0
NaCl at 23 C).
SPE 167525 5

(µ‐µ∞ )/(µp0‐µ∞) 10

1 1500ppm
2000ppm
2500ppm
0.1

0.01
0.01 0.1 1 10 100 1000
Shear rates (s‐1)

Figure 2: Dependence of (     ) /(  p    )
0
on shear rate and polymer concentration
0
(Lot# SP 0307* in 4% NaCl at 23 C).

Testing Validity of the Model with Experimental Coreflood Results


The proposed model is used to analyse the series of polymer bulk viscosity and coreflood experimental results of
Magbagbeola et al. (2008). The sets of coreflood results were conducted with outcrops made out from Berea sandstones of
varying permeabilities. The properties of the core and the schedules of polymers used in the experiments are summarised in
Tables 1, and 2 respectively. Equation (4) is used to calculate and characterize the shear-thinning behaviour of the polymer
solutions from the viscometer-measured viscosity over the entire range of rates of shear. The difference between the
laboratory-measured and predicted viscosities is calculated, and using an optimization program, the difference is reduced to a
minimum, thus obtaining the best estimate of the model adjustable matching parameters as shown in Table 3.

Table 1: Summary of basic core properties.


Property Core#-1 Core#-3 Core#-4 Core#-5
Length (cm) 20.32 20.32 20.32 20.32
Porosity 0.23 0.235 0.24 0.243
2
Area (cm ) 5.07 5.07 5.07 5.07
Brine permeability (mD) 647 552 372 260

Table 2: Summary of polymers used in the experiments.


Lot # Mol. Weight Degree of Hydrolysis Concentration Salinity
(Dalton) (%) (ppm)

1500

SP 0307* Unknown 25 – 27 2000 4% NaCl

2500

6
UE 4353 20 x 10 30 1500 SFB

1500 SFB
6
SU 0508* 20 x 10 28 - 30
SSFB
2+
w/20ppm Ca
6 SSFB
SU 0708* 26 x 10 28 - 30 1500 2+
w/20ppm Ca
(*): Lot designations are not from manufacturers.
SFB: Synthetic Formation Brine, SSFB: Softened Synthetic Formation Brine.
6 SPE 167525

Table 3: Summary of optimized Carreau model


adjustable fitting parameters.
Parameter Core#-1 Core#-3 Core#-4 Core#-5
µ∞, cp 1.02 1.02 1.02 1.02
0
µp , cp 13.76 6.540 18.086 65.877
λ, s 0.221 0.066 0.175 1.631
n 0.755 0.828 0.720 0.662

Discussion of Results
Based on the experimental design, the bulk or steady-shear viscosity for polymer lot# SP 0307* are measured in terms of
concentration of the polymer, shear rate, salinity, degree of hydrolysis (DH) and temperature. Table 3 and Figures 3a-b
through 5 show clearly how well the Carreau model characterises the shear-thinning behaviour of the various polymer
solutions (lot# UE4353, SU0508*, and SU0708*). The polymer relaxation time increases with increasing degree of
hydrolysis (see table 1). Kim et al. (2010, 2009) studied the effect of salinity on HPAM steady-shear behaviour. They
reported that cations screen the negative charges of carboxyl groups along the chain of the polymer; causing chain
contraction and leading to reduction in electrostatic repulsion. According to the study, when salts are added in polymer
solution, there is a shift in the critical shear rate (the on-set of shear-thinning) position to higher values; which in turn causes
a reduction in the slope of the shear-thinning region. Figures 3 a―b through 5 show these trends. Divalent salts have more
impact on viscosity reduction than monovalent salts. Figure 6 is in agreement with this study (which reported salts adversely
degrade solution viscosity); it compares the effect of salinity, hardness and molecular weight on shear viscosity for all
polymers (1500ppm in 4% NaCl), and shows that as molecular weight decreases, polymer relaxation time decreases (also
shown in Table 3). Figure 6 also shows that if all other conditions remain equal, the polymer with the larger molecular weight
would have higher viscosity (compare lot# SU0508 and Lot# SU0708). The effect of calcium ion on relaxation time which is
greater for Lot# SU0708 with a higher molecular weight compared to lot# SU0508 is also observed (Tables 2 and 3). Figure
7 shows the effect of Salt on Carreau model fit to 1500ppm for lot# SU 0708 in SSFB w/20ppm Ca2+ and lot# SU 0508 in
SFB, while Figure 8 shows the effect of molecular weight on Carreau model fit to 1500ppm for lot# SU 0708 in SSFB
w/20ppm Ca2+ and lot# SU 0508 in SSFB w/20ppm Ca2+ respectively.

100 100
Viscosity (cp)

Viscosity, cp

10 10

1 1
0.1 1 10 100 1000 0.1 1 10 100 1000
Shear Rates (1/s) Shear Rates (s‐1)
µ_msd µ_msd
Carreau fit Carreau fit
(a) (b)
Figure 3: Carreau model fit to 1500ppm (a) lot# UE 4353 in SFB. (b) lot# SU 0508 in SFB.
SPE 167525 7

Viscosity (cp) 100

10 µ_msd
Carreau fit

1
0.01 0.1 1 10 100 1000
Shear rate (s‐1)

2+
Figure 4: Carreau model fit to 1500ppm lot# SU 0508 in SSFB w/20ppm Ca .

100
Viscosity, cp

µ_msd
10 Carreau fit

1
0.01 0.1 1 10 100 1000
Shear Rates (s‐1)

2+
Figure 5: Carreau model fit to 1500ppm lot# SU 0708 in SSFB w/20ppm Ca .

100

Lot# UE4353 in SFB
Vsicosity (cp)

Lot# SU0508 in SFB
10
Lot# SU0508 in
SSFB w/20ppmCa+
Lot# SU0708 in
SSFB w/20ppmCa+

1
0.01 0.1 1 10 100 1000
Shear Rates (s‐1)

Figure 6: Combined effect of salinity, hardness and molecular weight on shear viscosity for all polymers
(1500ppm in 4% NaCl).
8 SPE 167525

100

Lot# SU0508 in
2+
SSFB w/20ppm Ca
Viscosity (cp)

µ_msd
10
Carreau fit
Lot# SU0508 µ_msd
in SFB
Carreau fit

1
0.01 0.1 1 10 100 1000
Shear rates (s‐1)

2+
Figure 7: Effect of Salt on Carreau model fit to 1500ppm lot# SU 0708 in SSFB w/20ppm Ca and
lot# SU 0508 in SFB.

100
Lot# SU0708
6
Mw = 26x10
Viscosity, cp

µ_msd
10 Lot# SU0508
Mw = 20x10
6 Carreau fit
µ_msd
Carreau fit
1
0.01 0.1 1 10 100 1000
Shear Rates (s‐1)

2+
Figure 8: Effect of molecular weight on Carreau model fit to 1500ppm lot# SU 0708 in SSFB w/20ppm Ca and
2+
lot# SU 0508 in SSFB w/20ppm Ca .

In this dynamic single-phase polymer flow study, our primary interest is the absolute permeability change during the flow
and interaction of the single-phase polymer solution with the rock media. Core properties Core#-3 and Core#-4 for varying
initial permeabilities and polymer lot# SU0508 (same concentration of 1500ppm) in different salinity and ionic conditions are
used to show the effect of the damage induced by the polymer. Figures 9 through 10 demonstrate the calculated magnitude of
absolute permeability alteration as a function of flow rate using our new model (equation 8). Figures 9 and 10 show how
single-phase polymer flood influences the rate of change of total permeability in a non-linear manner. Even though the same
degree of impairment is ultimately observed, the damage is accelerated and consequently greater in the less-permeable
medium. This is shown in Figure 11 for flow rates and Figure 12a and b for different ranges of pressure drops. According to
Figures 9 through 12s, different permeability change occurs for different flow rates/ pressure drops.
SPE 167525 9

4.0
Absolute Permeability (mD)

3.5
3.0
2.5
2.0
1.5
1.0
0.5
0.0
0.00 0.50 1.00 1.50 2.00 2.50
Flow Rates (ml/min)

Figure 9: Absolute permeability vs. flow rates calculated for single-phase polymer flood
(Lot# SU0508 in SFB-1500ppm, Core#-3).

4.0
3.5
Absolute permeability (mD)

3.0
2.5
2.0
1.5
1.0
0.5
0.0
0.00 0.50 1.00 1.50 2.00 2.50 3.00
Flow rates (ml/min)

Figure 10: Absolute permeability vs. flow rates calculated for single-phase polymer flood
2+
(Lot# SU0508 in SSFB w/20ppm Ca , 1500ppm, Core#-4).
10 SPE 167525

Absolute permeability (mD) 5.0

4.0

3.0 Cal kappBV‐3

2.0 Calc. kapp BV‐4

1.0

0.0
0.00 0.50 1.00 1.50 2.00 2.50
Flow rate (ml/min)

Figure 11: Comparison of absolute permeability vs. flow rates calculated for single-phase
polymer flood (Lot# SU0508, 1500ppm, Core#-3 and Core#-4).

4.0 4.0
Cal kappBV‐3 Calc. kapp BV‐4

Absolute permeability (mD)
3.5 3.5
Absolute permeability (mD)

Calc. kapp BV‐4 Cal kappBV‐3
3.0 3.0

2.5 2.5
2.0
2.0
1.5
1.5
1.0
1.0
0.5
0.5
0.0
0.0 0 200 400 600
0 20 40 60 Pressure drop (psi)
Pressure drop (psi)

(a) (b)
Figure 12: Absolute permeability vs.pressure drop calculated for single-phase polymer flood at different ranges of flow rates (Lot#
SU0508, 1500ppm, Core#-3 and Core#-4).

Though the relationship between absolute permeability and flow rate is non-linear, the plots of absolute permeability against
the inverse of flow rate and shear rate however show a linear relationship. Figures 13 and 14 show the linear relationship of
absolute permeability vs. inverse of flow rates and shear rates calculated for single-phase polymer flood (Lot# SU0508,
1500ppm, Core#-4). Since Darcy’s law was applied in both the flow experiments and the model formulation, the linear
relationship between absolute permeability and flow rate is in line with Darcy’s theory where the flow rate could be related to
the imposed pressure drop through a linear proportionality constant, k; where k is a permeability coefficient of the sand layer.
SPE 167525 11

4.0
y = 0.7704x + 0.5344
3.5
Absolute permeability (mD)

R² = 0.9628
3.0
2.5
2.0
1.5
1.0
0.5
0.0
0.00 0.50 1.00 1.50 2.00 2.50 3.00 3.50 4.00 4.50
Flow rate inverse(min/ml)

Figure 13: Linear relationship of absolute permeability vs. inverse of flow rate calculated for
single-phase polymer flood (Lot# SU0508, 1500ppm, Core#-4).

4.0
y = 568.29x + 0.5344
3.5
Absolute permeability (mD)

R² = 0.9628
3.0
2.5
2.0
1.5
1.0
0.5
0.0
0 0.001 0.002 0.003 0.004 0.005 0.006
Shear rate inverse (s)

Figure 14: Linear relationship of absolute permeability vs. inverse of shear rate calculated
for single-phase polymer flood (Lot# SU0508, 1500ppm, Core#-4).

Conclusions
By the results presented in this study, we describe the first phase of much larger research efforts aimed at developing
formation damage models for oil field polymer viscosifiers. Hence, we offer the following conclusions based on the materials
presented:
 Using literature data for steady-shear viscosity measurements and core flood laboratory experiments involving
single-phase non-Newtonian polymer solution for different flow rates, shear rates, molecular weight, concentration,
degree of hydrolysis, ionic condition, permeabilities, a new model for understanding absolute permeability alteration
arising from the propagation of single-phase polymer flow in porous media has been developed and verified.
 The shear-thinning behaviour of the poymer solution considered in the model was characterised by the carreau
model. The Carreau equation was used because it covers both the first and second regions of the viscosity curve.
 Fundamentally, the results show that absolute permeability change as functions flow rate, shear rate and pressure
drop is greater for lower permeable medium systems.
 The absolute permeability damage due to the polymer flow was also correlated as a linear function of the inverse of
the flow rate and shear rates.
12 SPE 167525

 Polymer adsorption causes permeability impairment (formation damage) and is detrimental to oil recovery and
productivity.

Acknowledgments
The first author would like to acknowledge and express gratitude to the Federal Government of Nigeria through the
Petroleum Technology Development Fund, PTDF for sponsoring his PhD research study on which the work reported in this
paper is based.

References
Annie, A., Jean-Francois, A., Ladva, H., Way, P. and Hove, A., 1999. Role of Polymers on Formation Damage. SPE European Formation
Damage Conference, The Hague, Netherlands, 31 May-1 June.

Bird, R.B., Armstrong, R.C. and Hassager, O., 1987. Dynamics of Polymeric Liquids, Vol. 1, Fluid mechanics, John Wiley & Sons, New
York City, pp. 92-95.

Bird, R.B., Stewart, W.E., and Lightfoot, E.N., 2002. Transport Phenomena, John Wiley & Sons, New York City. 895 pp.

Cannella, W., Huh, C. and Seright, R., 1988. Prediction of Xanthan Rheology in Porous Media. SPE Annual Technical Conference and
Exhibition, 63rd Annual Technical Conference, Houston, Texas, 2-5 October.

Carr, M. and Yang, B., 1998. Evaluation for Polymer Damage Aids in Candidate Selection for Removal Treatment. SPE Permian Basin Oil
and Gas Recovery Conference, Midland, Texas, 23-26 March.

Carreau, P.J., 1972. Rheological Equations From Molecular Network Theories. Transactions of the Society of Rheology, 16(1), pp. 99-127.

Chauveteau, G., Denys, K. and Zaitoun, A., 2002. New Insight on Polymer Adsorption Under High Flow Rates. SPE/DOE Improved Oil
Recovery Symposium.

Clark, P.E., 2010. Analysis of fluid loss data II: Models for dynamic fluid loss. Journal of Petroleum Science and Engineering, 70(3), pp.
191-197.

Da Silva, I., Lucas, E. and De Franca, F., 2010. Study of Conditions for Polyacrylamide Use in Petroleum: Physical Flow Simulation in
Porous Media. Chemical Technology, Vol. 4, No. 1.

Dang, C., Chen, Z.J., Nguyen, N., Bae, W. and Phung, T.H., 2011. Development of Isotherm Polymer/Surfactant Adsorption Models in
Chemical Flooding. SPE Asia Pacific Oil and Gas Conference and Exhibition.

Darcy, H.P.G., 1856. The Public Fountains of the City of Dijon (appendix), Bookseller of the Imperial Corps of Bridges, Highways and
Mines, Quay of Augustins, 49. English translation: http//biosystems.okstate.edu/darcy/

Dehghanpour, H. and Kuru, E., 2011. Effect of viscoelasticity on the filtration loss characteristics of aqueous polymer solutions. Journal of
Petroleum Science and Engineering, 76(1), pp. 12-20

Delshad, M., Kim, D.H., Magbagbeola, O.A., Huh, C., Pope, G.A. and Tarahhom, F., 2008. Mechanistic Interpretation and Utilization of
Viscoelastic Behaviour of Polymer Solutions for Improved Polymer-Flood Efficiency. SPE/DOE Symposium on Improved Oil Recovery,
Tulsa, Oklahoma, USA: Society of Petroleum Engineers, 20-23 April.

Du, Y. and Guan, L., 2004. Field-Scale Polymer Flooding: Lessons Learnt and Experiences Gained During Past 40 Years. SPE
International Petroleum Conference in Mexico.

Escudier, M. et al., 2001. On the Reproducibility of the Rheology of Shear-Thinning Liquids. Journal of Non-Newtonian Fluid Mechanics,
97(2), pp. 99-124.

Fletcher, A.J.P., Flew, S.R.G., Lamb, S.P., Lund, T., Bjornestad, E., Stavland, A. and Gjovikli, N.B., 1991. Measurements of
polysaccharide polymer properties in porous media. SPE International Symposium on Oilfield Chemistry, Anaheim, California, USA, 20-
22 February.

Green, D.W. and Willhite, G.P., 1998. Enhanced Oil Recovery. Richardson, Tex.: Henry L. Doherty Memorial Fund of AIME, Society of
Petroleum Engineers.
SPE 167525 13

Hirasaki, G. and Pope, G., 1974. Analysis of Factors Influencing Mobility and Adsorption in the Flow of Polymer Solution Through Porous
Media. Old Society of Petroleum Engineering Journal, 14(4), pp. 337-346, August.

Kim, D.H., Lee, S., Ahn, C.H., Huh, C. and Pope, G., 2010. Development of a Viscoelastic Property Database for EOR Polymers. SPE
Improved Oil Recovery Symposium, Tulsa, Oklahoma, USA, 24-28 April.

Lake, L.W., 1989. Enhanced Oil Recovery. Prentice-Hall, Inc., Englewood Cliff, NJ, 600 pp.

Lee, S., Kim, D., Huh, C. and Pope, G., 2009. Development Of A Comprehensive Rheological Property Database For EOR Polymers. SPE
Annual Technical Conference and Exhibition, New Orleans, Louisiana, 4-7 October.

Liang, Y. et al., 2011. Practice and Understanding of Separate-Layer Polymer Injection in Daqing Oil Field. SPE Production & Operations,
26(3), pp. 224-228.

Littman, W., 1988. Polymer Flooding, Elsevier, 212 pp.

Magbagbeola, O.A., 2008. Quantification of the Viscoelastic Behaviour of High Molecular Weight Polymers used for Chemical Enhanced
Oil Recovery, the University of Texas at Austin.

Masuda, Y. et al., 1992. 1D Simulation of Polymer Flooding Including the Viscoelastic Effect of Polymer Solution. SPE Reservoir
Engineering, 7(2), pp. 247-252.

Navarrete, R., Himes, R. and Seheult, J., 2001. Applications of Xanthan Gum in Fluid-Loss Control and Related Formation Damage. SPE
Permian Basin Oil and Gas Recovery Conference, Midland, Texas, 21-23 March.

Navarrete, R. et al., 2000. Experiments in Fluid Loss and Formation Damage with Xanthan-Based Fluids While Drilling. IADC/SPE Asia
Pacific Drilling Technology, Kuala Lumpur, Malaysia, 11-13 September.

Seright, R. et al., 2011. New Insights into Polymer Rheology in Porous Media. SPE Journal, 16(1), pp. 35-42

Seright, R.S., Seheult, J.M., and Talashek, T., 2009. Injectivity Characteristics of EOR Polymers. SPE Annual Technical Conference and
Exhibition. SPEReservoir Evaluation and Engineering 12 (5): 782-792.

Sorbie, K.S., 1991. Polymer-Improved Oil Recovery. Blackie and Son Ltd. Glasgow and London.

Sorbie, K. and Huang, Y., 1991. Rheological and Transport Effects In The Flow Of Low-Concentration Xanthan Solution Through Porous
Media. Journal of colloid and interface science, 145(1), pp. 74-89

Stavland, A., Jonsbraten, H.C., Lohne, A., Moen, A. and Giske, N.H., 2010. Polymer Flooding-Flow Properties in Porous Media versus
Rheological Parameters. SPE EUROPEC/EAGE Annual Conference and Exhibition, Barcelona, Spain, 14-17 June.

Sun, W., Li, F., Qu, Y., Yang, Y. and Li, K., 2012. Novel Method for Characterizing Single-phase Polymer Flooding. SPE EOR
Conference for Oil ang Gas West Asia held in muskat, Oman, 16-18 April.

Sydansk, D.R. and Seright, R.S., 2007. When and Where Relative Permeability Modification Water-Shutoff Treatments Can Be
Successfully Applied, SPE Production and Operations, May.

Wang, D., Seright, R.S., Shao, Z. and Wang, J., 2008. Key Aspects Of Project Design For Polymer Flooding At The Daqing Field, SPE
Reservoir Evaluation and Engineering, December, 1117.

Wreath, D., Pope, G.A. and Sepehrnoori, K., 1990. Dependence of Polymer Apparent Viscosity on the Permeable Media and Flow
Conditions, In Situ, 14(3), 263-284.

You might also like