Vous êtes sur la page 1sur 75

Polymer Flood

Process Manual
Revision 0I

By Computer Modelling Group Ltd.

This publication and the application described in it are furnished under license
exclusively to the licensee, for internal use only, and are subject to a confidentiality
agreement. They may be used only in accordance with the terms and conditions of
that agreement.
All rights reserved. No part of this publication may be reproduced or transmitted in any
form or by any means, electronic, mechanical, or otherwise, including photocopying,
recording, or by any information storage/retrieval system, to any party other than the
licensee, without the written permission of Computer Modelling Group.
The information in this publication is believed to be accurate in all respects. However,
Computer Modelling Group makes no warranty as to accuracy or suitability, and does
not assume responsibility for any consequences resulting from the use thereof. The
information contained herein is subject to change without notice.

Copyright 2012 Computer Modelling Group Ltd.


All rights reserved.

Builder, CMG, and Computer Modelling Group are registered trademarks of


Computer Modelling Group Ltd. All other trademarks are the property of their
respective owners.
Computer Modelling Group Ltd.
Office #150, 3553 - 31 Street N.W.
Calgary, Alberta Canada T2L 2K7

Tel: (403) 531-1300

Fax: (403) 289-8502

E-mail: cmgl@cmgl.ca

Contents
Introduction

Purpose ................................................................................................................................. 3
Organization ......................................................................................................................... 3
Polymer Flood Process ......................................................................................................... 3

Polymer Flood

Introduction .......................................................................................................................... 5
Theoretical Phenomena ........................................................................................................ 5
Polymer Adsorption......................................................................................................... 6
Permeability Reduction ................................................................................................... 7
Inaccessible Pore Volume (IPV) ................................................................................... 10
Relative Permeability/Wettability Alteration Effects .................................................... 11
Polymer Degradation ..................................................................................................... 14
Composition-Dependent Viscosity Effects ................................................................... 17
Shear-Dependent Viscosity Effects ............................................................................... 22
Power-Law Expression for Shear-Thinning or Pseudoplastic Fluids....................... 26
Power-Law Expression for Shear-Thickening or Dilatant Fluids ............................ 27
Tabular Input Option for Velocity/Shear-Rate-Dependent Viscosity ...................... 30
Velocity Dependent Skin Factor .............................................................................. 31
Convert Shear Rates to Velocities ............................................................................ 32
Salinity-Dependent Viscosity Effects............................................................................ 33
Lab and Field Information .................................................................................................. 34
Using the Process Wizard to Model a Polymer Flood ........................................................ 35
Viewing and Adjusting the Process Wizard Results .......................................................... 39
Components Generation ................................................................................................ 39
Polymer Consumption Reaction .................................................................................... 39
Polymer Adsorption....................................................................................................... 40
Polymer Adsorption Table ....................................................................................... 40
Langmuir Isotherm Option .......................................................................................40
Polymer Viscosity ......................................................................................................... 41
Other Considerations and Troubleshooting Information .................................................... 42
High Molecular Weight (or Low Mole Fractions) ........................................................ 42
Disproportionate Permeability Reduction Effect .......................................................... 43
Interpreting Polymer Flood Model Outputs........................................................................ 44

Appendix A Equations

45

Introduction ........................................................................................................................ 45
General Formulas................................................................................................................ 45
Revision 0I

Process Manual - Polymer Flood 1

Parts-per-million (ppm) ........................................................................................... 45


Weight Percentage ................................................................................................... 45
Mass Fraction (wt) ................................................................................................... 45
Mole Fraction (dim) ................................................................................................. 45
Polymer Flood Calculations ............................................................................................... 46
Polymer Mass Fraction ............................................................................................ 46
Polymer Mole Fraction ............................................................................................ 46
Volumetric Reaction Rate........................................................................................ 46
Stoichiometric Coefficients ..................................................................................... 48
Polymer Adsorption Table ....................................................................................... 49
Langmuir Isotherm Option ...................................................................................... 50
Permeability Reduction............................................................................................ 52
Polymer Viscosity .................................................................................................... 53

Appendix B Conversion of Eclipse Polymer Option to IMEX

64

Example of Eclipse Data .................................................................................................... 64


PLYVISC and PLYSHEAR Conversion to IMEX ............................................................ 65
PLYROCK Conversion to IMEX ...................................................................................... 66
PLYADS Conversion to IMEX.......................................................................................... 66

Appendix C Conversion of IMEX Polymer Option to STARS

68

Step 1: Builder Convert Simulator Type ............................................................................ 68


Step 2: Process Wizard....................................................................................................... 68
Input Polymer Data ....................................................................................................... 68
Input Adsorption Data................................................................................................... 68
Input Polymer Viscosity................................................................................................ 69
Step 3: Shear Dependent Viscosity .................................................................................... 69

Appendix D Example IMEX and STARS Simulations of Polymer

72

Comparisons of IMEX and STARS ................................................................................... 72


STARS Options For Real Polymer Phenomena ................................................................. 73

2 Process Manual - Polymer Flood

Revision 0I

Introduction
Purpose
The purpose of this manual is to provide users with the information they need to model
polymer flood processes using the CMG STARS simulator.

Organization
The following information is provided:
Theoretical concepts and how these concepts are represented in the model
Lab and field data required for the simulation
Procedure for inputting data using the Builder Process Wizard
Procedure for viewing and adjusting the input data
View the results in Results 3D
Appendices B, C, and D provide information for converting an Eclipse polymer
option to IMEX, and an IMEX polymer option to STARS.

Polymer Flood Process


Polymer flooding, illustrated below, is an EOR (enhanced oil recovery) technique in which
water-soluble polymers are added to the injection fluids to increase the viscosity of injected
water and/or formation water. This provides mobility control of the fluids, which improves
the volumetric sweep efficiency and reduces channeling and water breakthrough, thereby
increasing the oil recovery factor.

Revision 0I

Process Manual - Polymer Flood 3

Figure 1 shows a conceptual comparison of cumulative oil recovery for water, polymer,
alkaline/polymer, and ASP flooding techniques.
Cumulative
Oil

ASP Flood
Alkaline/Polymer Flood

Polymer Flood

Continued Water Flood

Pore Volume Injected

Figure 1 Comparison of Chemical Flooding Techniques

Figure 2 illustrates the potential improvement from chemical flood EOR processes:
Oil Rate

Water Flood

Chemical Flood
EOR Potential to
Extend and
Enhance
Production

Time

Figure 2 Improved Oil Recovery from Chemical Flood Processes 1

Adapted from Cairn India brochure Enhanced Oil Recovery.

4 Process Manual - Polymer Flood

Revision 0I

Polymer Flood
Introduction
Polymer flooding is a process where a thickening agent (polymer) is added to the injected
fluid (typically water) to produce a more favorable mobility ratio between the injected fluid
and the displaced oil. Polymers are macromolecules composed of repetitive units called
monomers. Some common polymers are hydrolyzed polyacrylamide (HPAM), co-polymers
of acrylamide (AMPS, NVP) and xanthan gum (biopolymer).
The polymer flood mechanisms that can be modeled in STARS include:
Viscosity and mobility variations of the injected fluid
Polymer Adsorption
Permeability Reduction
Inaccessible Pore Volume (IPV)
Relative Permeability/Wettability Alteration Effects
Polymer Degradation
Composition-Dependent Viscosity Effects
Shear-Dependent Viscosity Effects
Salinity-Dependent Viscosity Effects
Details about modeling these mechanisms are outlined in Theoretical Phenomena.
Modeling a polymer flood requires, as a minimum, the viscosity of the water-polymer
solution at different polymer concentrations. Other data, such as polymer adsorption,
degradation, and rheology, while not required, will yield a more accurate model. Refer to Lab
and Field Information on page 34. As well, refer to Other Considerations and
Troubleshooting Information on page 42 for information about specific polymer issues that
you may need to address or resolve.

Theoretical Phenomena
The following sections describe polymer flood process phenomena and, at a high level, how
they are modeled in STARS. For further information, refer to the STARS Users Guide.

Revision 0I

Process Manual - Polymer Flood 5

Polymer Adsorption
Adsorption is the adhesion of ions or molecules onto the surface of another phase 2. It is a
physical and/or chemical process by which a porous solid (at the microscopic level), for
example, is capable of retaining particles of a fluid on its surface after being in contact with
it. Polymer adsorption in an EOR process is related to the amount of polymer retained in the
smallest porous spaces or on the rock surface where the solution has passed. The adsorption
levels depend on fluid type and concentration, molecular weight, flow rate, temperature, brine
salinity, brine hardness and rock type (e.g. rock mineralogy and permeability)3.
Low polymer retention in the reservoir is essential for the success of a polymer EOR
operation. A substantial loss of polymer may be detrimental because the polymer
concentration reduction could impact its viscosity and cause a loss of mobility control or low
displacement efficiency. For this reason, adsorption is usually estimated from laboratory core
flood experiments conducted under conditions as close as possible to those prevailing in the
field.
STARS allows a description of these phenomena, through the input of a set of constant
temperature adsorption isotherms (adsorption level as a function of fluid composition). These
isotherms can be entered either in tabular form or using the Langmuir isotherm correlation:

Ad =

A ci
(1 + B ci )

(1)

where ci is the fluid component composition, and A and B are generally temperature
dependent. Note that the maximum adsorption level associated with the formula is A/B.
Coefficient B controls the curvature of the isotherm, and the ratio A/B, as mentioned,
determines the plateau value for adsorption. This is illustrated in Figure 3:
1.0

0.9

0.9

0.8

0.8

0.7

Adsorbed Concentration

Adsorbed Concentration

1.0

b increasing

0.6
0.5
0.4
0.3
0.2

0.0

0.7
0.6
0.5
0.4

a increasing

0.3
0.2

a/b constant

0.1

b constant

0.1
0

Concentration

10

0.0

6
Concentration

10

Figure 3 Typical Langmuir Isotherm Shapes3

Refer to Polymer Adsorption Table and Langmuir Isotherm Option in Appendix A


Equations of this manual for more information about the equations used by STARS to model
a polymer flood.
2

Ali, L. and Barrufet, M.A., Profile Modification due to Polymer Adsorption in Reservoir Rocks.
Energy & Fuels Vol. 8, No. 6, (1994), pp.1217-1222.
3
Lake, L.W., Enhanced Oil Recovery, Prentice-Hall (1989).

6 Process Manual - Polymer Flood

Revision 0I

Important Keywords for Modeling Polymer Adsorption:


Use keyword *ADSCOMP to specify components and fluid phase.
Use keywords *ADSLANG or *ADSTABLE to specify the input option used for adsorption,
as follows:
*ADSLANG

Denotes that composition dependence is specified by Langmuir


isotherm coefficients.

*ADSTABLE

Denotes that composition dependence is specified by a table of


adsorption versus composition values.

To define the dependence of the adsorption data on rock type (permeability) for the
component/phase specified by *ADSCOMP, use the following keywords:
*ADMAXT

Specifies the maximum adsorption capacity.

*ADRT

Specifies the residual adsorption level.

*ADSROCK
*ADSTYPE

Specifies the current rock type number.


Used to assign multiple adsorption rock type numbers to grid blocks.

Permeability Reduction
Many papers indicate a mobility reduction in the porous media after the polymer has flowed
through. This phenomenon occurs due to increased water viscosity and reduction of
permeability 4,5, caused in part by polymer adsorption, particularly if it is the chemical or
mechanical (entrapment)4 type. The variation in rock permeability due to this process is given
by:
=
K

where:

k
Rk

(2)

Absolute rock permeability prior to polymer flooding


Permeability reduction factor

, a function of polymer adsorption and the residual resistance factor RRF, is given by:
= 1 + ( 1)

(3)

Jennings, R. R., Rogers, J.H., and West, T. J., 1971. Factors Influencing Mobility Control By
Polymer Solutions. J. Pet. Technol., 23(3): 391-401. SPE 2867-PA.
5
Bondor, P.L., Hirasaki, G.J., Tham, M.J., 1972. Mathematical simulation of polymer flooding in
complex reservoirs. SPEJ (October), 369382. SPE 3524-PA.

Revision 0I

Process Manual - Polymer Flood 7

where:
AD

Cumulative adsorption of polymer per unit volume of reservoir rock

ADMAXT

Maximum adsorptive capacity of polymer per unit volume of reservoir


rock

Rk varies from 1.0 to a maximum of , as adsorption level increases.

The residual resistance factor can be obtained from core flooding experiments and can be
expressed as the water (or brine) mobility ratio before and after a polymer treatment4,6,7.
=

where:

, ( ),
=
( ),
,

Water mobility

kw

Effective water permeability

Water viscosity

(4)

As the water viscosity does not change before or after the treatment, equation (4) can be
reduced to:
=

,
,

(5)

If the core flooding experiment is linear, the effective permeability can be calculated using
the linear expression of Darcys equation:
=

( )
2

(6)

If the core flooding experiment is radial, then the effective permeability should be calculated
as follows:
(7)

If any of the above expressions are used to calculate the effective permeability before and
after the polymer flood in a core flooding experiment and if the injection rate is the same in

Chang, H. L., Polymer Flooding Technology Yesterday, Today, and Tomorrow, paper SPE 7043
presented at the 1978 SPE Symposium on Improved Methods for Oil Recovery, Tulsa, April 16-19.
7
Singleton, M. A., Sorbie, K. S., Shields, R. A., Further Development of the Pore Scale Mechanism
of Relative Permeability Modification by Partially Hydrolized Polyacrylamide, paper SPE 75184,
presented at the oil Recovery Symposium, Tulsa, Oklahoma, 2002.

8 Process Manual - Polymer Flood

Revision 0I

both tests, the residual resistance factor, RRF, can be expressed in terms of the pressure
drops, as follows 8:
=

,
,

(8)

It is typically assumed that only single-phase flow paths are altered by polymers; therefore,
the permeability reduction factor for each phase can be expressed as:
= 1 + ( 1)

(9)

which affects the effective permeability of phase , , as follows:


= =

where:

(10)

Water, oil or gas phase

Absolute permeability before and after the treatment, respectively.

Therefore, to account for permeability reduction in polymer flood simulations or in any other
EOR process in which the adsorption of components plays an important role, it is necessary
to input the residual resistance factor and the phase to which the resistance factor will be
applied.
Use the following keywords to model permeability reduction:
*ADMAXT

Specifies the maximum adsorption capacity, as outlined in Polymer


Adsorption.

*RRFT

Specifies the residual resistance factor for the adsorbing component


i specified via *ADSCOMP. The value of *RRFT must be greater
than or equal to 1. The default value is 1.

*ADSPHBLK

With sub-keyword phase_des, overrides the default phase to which


the resistance factor calculation is applied, as follows:
*W
Water (aqueous) phase
*O
Oil (oleic) phase
*G
Gas (gaseous) phase
*ALL All phases

Zaitoun, A. and Kohler, N., 1988, Two-phase Flow Through Porous Media: Effect of an Adsorbed
Polymer Layer, paper SPE 18085 presented at the 1988 SPE Annual Technical Conference and
Exhibition, Houston, TX, Oct. 2-5.

Revision 0I

Process Manual - Polymer Flood 9

DEFAULTS:
*RRFT

1 (no resistance effect)

*ADSPHBLK

If not specified, the resistance factor is applied to the phase that is the
source of the adsorbing component (specified by *ADSCOMP).

Note: Some polymers reduce water permeability more than they reduce hydrocarbon
permeability. The result of this phenomenon, referred to as disproportionate permeability
reduction (DPR)7,9,10 is that the residual resistance factor is neither the same nor constant for
all phases, so it should be calculated in core flooding experiments for each phase, as shown
below:
From water-oil experiments:
=
=

(11)

(12)

From water-gas experiments:


=
=

where:

(13)

(14)

Effective permeability to the phase , where = w, o or g

Sor

Residual oil saturation (water-oil experiment)

Sgr

Residual gas saturation (water-gas experiment)

Swirr

Irreducible water saturation (water-oil and water-gas experiments)

In the current version of STARS, it is only possible to assign a single value of the residual
resistance factor. For suggestions on how to simulate the DPR effect, refer to the workaround
outlined in Disproportionate Permeability Reduction Effect on page 43.

Inaccessible Pore Volume (IPV)


When the flow of polymer molecules through the porous media is restricted in small pore
throats, only the passage of water or brine is possible. When these pores cannot be contacted
by flowing polymer molecules, they are referred to as inaccessible pore volume (IPV). This

Botermans, C.W., van Batenburg, D.W., and Bruining, J., Relative Permeability Modifiers: Myth or
Reality? in SPE European Formation Damage Conference. 2001, The Hague, Netherlands.
10
Elmkies, Ph., et al., Polymer effect on gas/water flow in porous media, SPE/DOE IOR
Symposium, April 2002, SPE 75160.

10 Process Manual - Polymer Flood

Revision 0I

phenomenon was first reported by Dawson and Lautz (1972) 11, who showed that some pore
spaces may not be accessible to polymer molecules and that this allows polymer solutions to
advance and displace oil at a faster rate than predicted on the basis of total porosity. They
concluded that about 30% of the total pore volume may not be accessible to polymer
molecules, and this has been corroborated in recent experiments by Pancharoen, Thiele and
Kovscek (2010) 12. As a result, the effective porosity for a polymer solution is less than the
actual reservoir porosity. A reduced polymer porosity, , can be used to represent the
available pore volume to polymer solutions as follows:
= (1 )

(15)

where is the original porosity, adjusted for pressure and temperature of the block.

IPV can have beneficial effects on field performance. The rock surface in contact with the
polymer solution will be less than the total pore volume, thus decreasing polymer adsorption.
More importantly, if connate water is present in the smaller pores that are inaccessible to the
polymer, the bank of connate water and polymer-depleted injection water that precedes the
polymer bank is reduced by the amount of the inaccessible pore volume. One drawback,
however, is that movable oil located in the smaller pores will not be contacted by the polymer
and therefore may not be displaced.
In STARS, the difference (1 ) is requested directly, to update the porosity that will be
used for the adsorbing component i and the adsorption rock type. In the simulator, it is
denoted as the accessible pore volume or fraction of available pore volume to polymers or
any similar component, and it should be specified by keyword *PORFT. The default value of
*PORFT is 1, which means that there is no inaccessible pore volume.

Relative Permeability/Wettability Alteration Effects


The effect of the modification of relative permeability by polymer adsorption has been
intensively studied by many authors in the past and although they have found evidence of a
selective reduction of the relative permeability to water with respect to relative permeability
to oil 13,14,15, the conventional belief is that polymer flooding does not reduce residual oil
saturation on a micro-scale; rather, it allows the undisplaced oil to approach this low level

11

Dawson, R. and Lautz, R., Inaccessible Pore Volume in Polymer Flooding, SPE Journal, October
1972.
12
Pancharoen, M., Thiele, M.R., and Kovscek, A.R., Inaccessible Pore Volume of Associative
Polymer Floods, paper SPE 129910, SPE Improved Oil Recovery Symposium, Tulsa, Oklahoma,
2010.
13
Schneider, N. and Owens, W.W., 1982, Steady-state measurements of relative permeability for
polymer/oil systems, paper SPE 9408-PA, Society of Petroleum Engineers Journal, 79-86.
14
Barrufet, A., and Ali, L., Modification of Relative Permeability Curves by Polymer Adsorption,
paper SPE 27015 presented at the 1994 Latin American/Caribbean Petroleum Engineering Conference,
Buenos Aires, Argentina, April 27-29.
15
Zheng, C. G., Gall, B. L., Gao, H. W., Miller, A. E., and Bryant, R. S., Effects of Polymer
Adsorption and Flow Behavior on Two-Phase Flow in Porous Media, paper SPE 39632 presented at
the 1998 SPE/DOE Improved Oil Recovery Symposium, Tulsa, Oklahoma, U.S.A, 19-22 April 1998.

Revision 0I

Process Manual - Polymer Flood 11

more quickly, while producing less water in the process. In more recent studies 16,17,18,
however, it can be seen that all types of micro-scale residual oil were reduced after flooding
with viscous-elastic polymers, increasing micro-scale displacement efficiency in the cores.
Figure 4 and Figure 5 show typical relative permeability curves for water flooding and
polymer flooding obtained from previous studies (Wang et al):
100
Kro
Krw
Krop

Relative Permeability %

80

Krp
fw
fp

60

40

20

0
0

20

40

60

80

100

Water Saturation %

Figure 4 Comparison of kr-curves of Polymer/Oil and Water/Oil16


100

Relative Permeability %

500 mg/L
1500 mg/L

80

2000 mg/L

60

40

20

20

40

60

80

100

Water Saturation %

Figure 5 Influence of Different Polymer Concentration on kr-curves17


16

Wang, D., Cheng, J., Yang, Q., Gong, W., and Li, Q., Viscous-Elastic Polymer Can Increase
Microscale Displacement Efficiency in Cores, paper SPE 63227-MS presented at the 2000 SPE
Annual technical and Exhibition held in Dallas, Texas, U.S.A, 1-4 October 2000.
17
Wang, D., Wang, G., Wu, W., Xia, H., and Yin, H., The Influence of Viscoelasticity on
Displacement Efficiency From Micro- to Macroscale, paper SPE 109016-MS presented at the SPE
Annual Technical Conference and Exhibition, Anaheim, California, U.S.A, 11-14 November 2007.
18
Xia, H., Wang, D., Ma, W., and Liu, J., Mechanism of the Effect of Micro-Forces on Residual Oil
In Chemical Flooding, paper SPE 114335 presented at the 2008 SPE/DPE Improved Oil Symposium
Held in Tulsa, Oklahoma, U.S.A, 19-23 April 2008.

12 Process Manual - Polymer Flood

Revision 0I

Chun and Pope (2008)19 reported that while a tertiary polymer flood did not mobilize the
water flood residual oil saturation, a secondary polymer flood did cause a displacement of oil
saturation below the water flood residual oil saturation in the same core. Also, the authors
indicate that the water flood and secondary polymer flood results could not be matched in the
simulations conducted if the same residual oil saturation value is used for both floods. The
core flood data could be only matched when the residual oil saturation value for the polymer
flood was lower than that for the water flood 19.
Although it is known that polymer flooding will not be applied as an enhanced secondary
recovery, this is precisely what happens when the oil that was bypassed by a previous water
flood process and that is trapped in the low-permeable zone is mobilized by the polymer;
therefore, if the reduction in oil residual saturation in the simulations is neglected, the oil
recovery might be underestimated, resulting in large errors and improperly forecasted
results 20,21.
To simulate this modification, STARS can optionally interpolate basic relative permeability
and capillary pressure data as functions of concentration. With this option, the curvature and
endpoints of the curves can be modified based on laboratory experimental data, and used for
each grid block depending on the polymer adsorption or concentration level. Enabling this
option in STARS requires the following keywords:
*INTCOMP

With sub-keywords comp_name and phase, indicates respectively, the


name of the component upon whose composition the rock-fluid
interpolation will depend, and the phase from which the components
composition will be taken:
*WATER Water (aqueous) mole fraction
*OIL
Oil (oleic) mole fraction
*GAS
Gas mole fraction
*GLOBAL Global mole fraction
*MAX
Maximum of water, oil and gas mole fractions
*ADS
Adsorption phase, fraction of maximum

*KRINTRP

Indicates the interpolation set number, local to the current rock-fluid


rock type. Values start at 1 for each new rock type and increase by 1
for each additional interpolation set. For example, rock type #1 might
have local set numbers 1 and 2 while rock type #2 might have local set
numbers 1, 2 and 3.

19

Chun, H., and Pope, G. A., Residual Oil Saturation From Polymer Floods: Laboratory
Measurements and Theoretical Interpretation, paper SPE 113417 presented at the 2008 SPE/DPE
Improved Oil Symposium Held in Tulsa, Oklahoma, U.S.A, 19-23 April 2008.
20
Kamaraj, K., Zhang, G., and Seright, R., Effect of Residual Oil Saturation on Recovery Efficiency
during Polymer Flooding of Viscous Oils, paper OTC 22040 presented at the Arctic Technology
Conference Held in Houston, Texas, U.S.A, 7-9 February 2011.
21
Chen, G., Han, P., Shao, Z., Zhang, X., Ma, M., Lu, K., and Wei, C., History Matching Method for
High Concentration Viscoelasticity Polymer Flood Pilot in Daqing Oilfield, paper SPE 144538
presented at the SPE Enhanced Oil Recovery Conference held in Kuala Lumpur, Malaysia, 19-21 July
2011.

Revision 0I

Process Manual - Polymer Flood 13

*DTRAPW

Indicates the value of the wetting phase interpolation parameter (mole


fraction) for the current rock-fluid data set.

*DTRAPN

Indicates the value of the non-wetting phase interpolation parameter


(mole fraction) for the current rock-fluid data set.

*WCRV
*OCRV

Indicates the curvature change parameter for water relative


permeability.
Indicates the curvature change parameter for oil relative permeability.

*GCRV

Indicates the curvature change parameter for gas relative permeability.

*SCRV

Indicates the curvature change parameter for liquid relative


permeability.

DEFAULTS:
If *INTCOMP is absent, interpolation will not be enabled.
For a rock type, if *KRINTRP is absent then there is no rock-fluid interpolation.
At least one of *DTRAPW and *DTRAPN must be present to enable interpolation.
If only one is present, its value is applied to the absent keyword.
Each of *WCRV, *OCRV, *GCRV and *SCRV default to 1 if absent.
For more detail, refer to the STARS Users Guide.

Polymer Degradation
Polymer degradation refers to any process that breaks down the molecular structure of
polymer macromolecules. The main degradation mechanisms that may be of concern in an
EOR process are chemical, thermal, biological and mechanical22:
Chemical degradation of polymer, or polymer chemical stability, is mainly
controlled by oxidation-reduction reactions and hydrolysis, which are due to the
presence of divalent cations such as Ca2+, Mg2+, Fe2+ in the water, and oxygen,
which breaks down the polymer molecular chains 23,24.
Thermal degradation of polymers is commonly associated with chemical
degradation and is defined as molecular deterioration resulting from overheating.
At high temperatures, the components of the long-chain backbone of the polymer
can separate (molecular scission) and react with one another to change the
properties of the polymer rheology and phase behavior. Thermal degradation
generally involves changes to the molecular weight and it can occur at
temperatures which are much lower than those at which mechanical failure is
likely to occur.
Biological degradation is more prevalent for biopolymers than it is for synthetic
polymers, however it can occur in both. Biological degradation, the microbial
breakdown of polymer macromolecules in the presence of bacteria in the reservoir,
occurs more often at lower temperatures and salinities23,24.
22

Chang, H. L., Polymer Flooding Technology Yesterday, Today, and Tomorrow, paper SPE 7043
presented at the 1978 SPE Symposium on Improved Methods for Oil Recovery, Tulsa, April 16-19.
23
Littmann, W., Polymer Flooding, Developments in Petroleum Science, Vol. 24, Elsevier,
Amsterdam (1988), pp 32-34.
24
Sheng, J.J.: Modern Chemical Enhanced Oil Recovery: Theory and Practice, Elsevier, 2011.

14 Process Manual - Polymer Flood

Revision 0I

Mechanical degradation occurs when a polymer solution is exposed to high shear


conditions, which fragment/break the polymer molecular chains, resulting in loss
of viscosity and mobility control. This may happen when a polymer solution is
forced at high flow rates through a porous medium or is in the vicinity of an
injection well, where it is sheared by the high velocities4,25. Mechanical
degradation of polymer is more severe at higher flow rates and longer flow
distances, and with lower permeability porous media. The behavior of the polymer
as a non-Newtonian fluid in the presence of shear conditions and the effects of
shear-thinning and shear-thickening can be modeled by STARS as explained in
Shear-Dependent Viscosity Effects.

All of the above degradation mechanisms, if applicable, should be considered in an EOR


process. To be effective, polymer solutions must remain stable for a long time at reservoir
conditions. These degradations are often neglected at the simulation level because the
processes are quite complex. A simple approximation for predicting these phenomena is to
use a first-order reaction where the polymer is degraded to water. In this case, the polymer
concentration is reduced in each grid cell over time, which results in a reduced viscosity that
can be calculated with the same nonlinear mixing rule used to account for the behavior of the
aqueous phase viscosity with polymer concentration.
For the simple case of polymer consumption, the first-order reaction can be expressed as
follows:

Sto1 Polymer( w ) Sto 2 Water( w )

(16)

where:

Sto1

Stoichiometric coefficient of reacting component

Sto 2

Stoichiometric coefficient of produced component

Typical values of polymer viscosity stability over time are shown in Table 1 and Figure 6:
Table 1 Polymer Viscosity Specifications

Time (Days)
0
5
10
15
30
45
60

Polymer 1
6.17
2.84
1.91
1.47
0.92
0.91
0.84

Viscosity (cps)
Polymer 2
7.15
3.65
2.97
2.42
1.65
1.47
1.25

Polymer 3
8.02
5.00
3.50
3.00
2.30
2.00
1.80

25

Seright, R.S., The Effects of Mechanical Degradation and Viscoelastic Behavior on Injectivity of
Polyacrylamide Solutions, paper SPE 9297 presented at the 55th Annual Fall Technical Conference,
Society Petroleum Engineers, Dallas, Sept. 1980.

Revision 0I

Process Manual - Polymer Flood 15

Figure 6 Viscosity Stability of Sample Polymers

The STARS reaction modeling capability is quite robust and can be used to model more
complex reactions if desired. The STARS simulator allows the user to model kinetic reactions
in the formation and/or the breaking of any component as a function of fluid flow velocity,
temperature, concentration and type of process.
The mandatory keywords for the chemical reaction data are:
*STOREAC

Used to specify the stoichiometric coefficient of reacting component

*STOPROD

Used to specify the stoichiometric coefficient of produced component

*FREQFAC

Used to specify the reaction frequency factor

If the reaction is different from the first-order reaction and the components are reacting in a
different phase than they were originally part of, *RPHASE and *RORDER must also be
specified:
*RPHASE

Used to specify a flag defining the phase of the reacting component.


The allowed values of this flag are as follows:
0 = non-reacting components,
1 = water phase (fluid components only)
2 = oil phase (fluid components only)
3 = gas phase (fluid components only)
4 = solid phase (solid components only)
Note: An adsorbing component may not react in the adsorbed phase.

*RORDER

Used to specify the order of the reaction with respect to each reacting
components concentration factor. It must be non-negative.
Enter zero for non-reacting components. Normally, you would use a
value of one (1); however, if the value is zero (0), the reaction rate
will be independent of that components concentration.

16 Process Manual - Polymer Flood

Revision 0I

DEFAULTS:
If *RPHASE is absent, the assumption is:
iphas = 0 for non-reacting components
iphas = 1 for aqueous components 1 to numw
iphas = 2 for oleic components numw+1 to numx
iphas = 3 for noncondensable components numx+1 to numy
iphas = 4 for solid components numy+1 to ncomp
If *RORDER is absent, the assumption is:
enrr = 0 for non-reacting components
enrr = 1 for reacting components
If the process you are simulating is thermal (i.e., the polymer degradation rate depends on the
temperature and activation energy [Ea]), the first-order reaction may be rewritten to depend
on the absolute temperature in the grid cells according to the Arrhenius equation. In this case,
new keywords *RENTH and *EACT are required:
*RENTH

Used to specify the reaction enthalpy (J/gmol | Btu/lbmol). It is


positive for exothermic reactions and negative for endothermic
reactions. The default is 0.

*EACT

Used to specify the single activation energy (J/gmol | Btu/lbmol).


Defines the dependence of reaction rate on grid block temperature. If
absent, the reaction is independent of temperature (equivalent to
*EACT with Ea = 0).

For information about the formulas used in determining the volumetric reaction rate, refer to
Volumetric Reaction Rate on page 46. For information about the formulas used to calculate
the stoichiometric coefficients, refer to Stoichiometric Coefficients on page 48. For more
details refer to the STARS Users Guide.

Composition-Dependent Viscosity Effects


The main purpose of using a polymer is to achieve a favorable mobility ratio. The efficiency
of a waterflood is heavily dependent on the mobility ratio of the displacing and displaced
fluids. The mobility ratio is defined as follows:
=

/
=
=

Revision 0I

(17)

Process Manual - Polymer Flood 17

where:
kw

Effective permeability to water

w
ko

Water viscosity
Effective permeability to oil

Oil viscosity

With a high mobility ratio, the displacing fluid moves much faster than the displaced fluid.
As a result, a phenomenon called viscous fingering occurs in which the displacing fluid
bypasses the displaced fluid and channels towards the producer. Figure 7 illustrates this
effect. The fingering of the water leaves behind large areas of oil that are unswept by the
water. The water also channels itself towards the producer (upper left). Once this
communication has been established, water will go straight from the injector to the producer,
bypassing the remaining oil in the reservoir.

Figure 7 Water Fingers through Oil due to Adverse Viscosity Effects 26

If polymer is added to the injected water, the mobility ratio is lowered (more favorable). The
polymer increases the viscosity of the injected water, thus reducing its mobility. The flood
front is more uniform, and there is very little evidence of viscous fingering. The displacement
is more piston-like and leaves very little trapped oil behind. This is illustrated in Figure 8.

Figure 8 Improved Sweep Efficiency and Oil Recovery through Use of Polymer26

Component viscosity can be entered in two ways. The first way is to enter a viscosity versus
temperature table using the keyword *VISCTABLE, which contains a temperature column
26

Adapted from Cairn India brochure Enhanced Oil Recovery.

18 Process Manual - Polymer Flood

Revision 0I

followed by columns corresponding to the number and order of the non-solid components
listed under the *MODEL keyword. For example, consider a fluid model containing water,
oil, gas and polymer. The *VISCTABLE will appear similar to that shown below:
VISCTABLE
T1

w,T1

o,T1

g,T1

P,T1

T2

w,T2

o,T2

g,T2

P,T2

T3

w,T3

o,T3

g,T3

P,T3

w,Tn

o,T4

g,Tn

P,Tn

Tn

Different VISCTABLEs can be defined at different pressures, allowing temperature and


pressure dependency to be modeled. The *ATPRES keyword can be used to specify a
*VISCTABLE at different pressures, as follows:
ATPRES pres_1
VISCTABLE
ATPRES pres_2
VISCTABLE
ATPRES pres_n
VISCTABLE
The second way to define component viscosity is to enter the coefficients and use a
correlation to calculate the viscosity at different temperatures. The correlation uses the
following equation:
= exp( /Tabs ) (18)

where:

avisc1, bvisc1

Coefficients of the correlation for the temperature-dependence of


a component viscosity in the liquid phase

Tabs

Absolute temperature

Within STARS, the viscosity of a liquid phase is modeled using either a linear mixing rule or
a nonlinear mixing rule. The default method in STARS is the linear mixing rule, which uses
the following equation to calculate the viscosity of a multi-component mixture:

( ) = (19)
=1

Revision 0I

Process Manual - Polymer Flood 19

where:

nc

Viscosity of mixture
Weighting factor of component i in the aqueous ( = w) or oleic ( = o)
phase viscosity in the linear mixing rule.
Viscosity of component i in the aqueous ( = w) or oleic ( = o) phase.
Number of components in the oleic or aqueous phase.

Factors = (oil mole fractions) and = (water mole fractions) are used for linear
mixing. To specify nonlinear mixing (for example, solution gas in the oleic phase or polymer
in the aqueous phase), use keywords *VSMIXCOMP, *VSMIXENDP and *VSMIXFUNC
(one instance for each key component), where factors and are different from the mole
(mass) fractions and respectively, for the key component.
To accomplish nonlinear mixing with alternate weighting factors, is replaced with the
nonlinear mixing function for each component i S and with for each
component i S, where S denotes the set of key components. The nonlinear mixing rule for
the liquid viscosity is calculated as:

=1

=1

( ) = +

where:

(20)

Viscosity of mixture aqueous ( = w) or oleic ( = o).


Viscosity of component i in the aqueous ( = w) or oleic ( = o) phase.
Weighting factor of the non-key component i in the aqueous ( = w) or
oleic ( = o) phase viscosity in the nonlinear mixing rule.
Weighting factor of the key component i in the aqueous ( = w) or oleic
( = o) phase viscosity in the nonlinear mixing rule.
Number of key components in the liquid phase.
Number of components in the liquid phase that are not key components.

In the determination of the weighting factors for the nonlinear mixing rule, each key
component acts independent of other key components, which is reflected in the fact that the
nonlinear mixing function depends only on the mass or mole fraction ( or ). This
implies that the function data entries must be generated assuming the absence of other key
components.

20 Process Manual - Polymer Flood

Revision 0I

The format for specifying the nonlinear mixing rule for liquid viscosities is as follows:
VSMIXCOMP comp_name

Specifies the component that is exhibiting nonlinear


viscosity mixing

VSMIXENDP xlow xhigh

Specifies the minimum and maximum compositions


of the specified comp_name

VSMIXFUNC f1 . . .f11

Nonlinear mixing function corresponding to the 11


intervals between xlow and xhigh

DEFAULTS:
If *VSMIXCOMP is absent, linear mixing is assumed for all components.
If *VSMIXENDP is absent, xlow = 0 and xhigh = 1 are assumed.
If *VSMIXFUNC is absent, entries = ( 1)/10, for i = 1 to 11, corresponding to
linear spacing from 0 to 1.

CONDITIONS:

The phase to which this data will be assigned depends on which of *LIQPHASE,
*WATPHASE and *OILPHASE is in force.
A nonlinear function may be specified for more than one component in each of the water and
oil phases. At least one component in each liquid phase must not be a key component, since
the algorithm involves adjusting the weighting factors of the non-key components.
Keywords *VSMIXENDP and *VSMIXFUNC are applied to the last key component defined
via *VSMIXCOMP. A key component may not be specified more than once in each liquid
phase.
An example of how these keywords should be entered in the simulation dataset to model the
nonlinear mixing rule of the water viscosity with the presence of a polymer in the aqueous
phase is shown below:
VSMIXCOMP 'Polymer'
VSMIXENDP 0 0.001
VSMIXFUNC 0 0.0 0.0759 0.1598 0.2514 0.3498 0.4534 0.5608 0.6704 0.7808 0.891 1.0
Internally, STARS will divide the composition interval (xhigh - xlow) into 11 equal subintervals,
corresponding to the f1...f11 values. Inside STARS, the table will appear as shown in the
following example:

Revision 0I

Process Manual - Polymer Flood 21

Composition
(wp)
0.0000
0.0001
0.0002
0.0003
0.0004
0.0005
0.0006
0.0007
0.0008
0.0009
0.0010

Mixing Function
f(wp)
0.0000
0.0759
0.1598
0.2514
0.3498
0.4534
0.5608
0.6704
0.7808
0.8910
1.0000

At any composition wp, a corresponding mixing function f(wp) will be determined either
directly from the table or through interpolation. This function will be used to calculate the
viscosity of the solution. As the composition changes due to injection, decomposition or
adsorption, the mixing function will change accordingly, resulting in a change of water
viscosity.
For further information and a calculation example, refer to the Polymer Viscosity section of
Polymer Flood Calculations in Appendix A Equations.

Shear-Dependent Viscosity Effects


Flow in a porous medium is affected by the morphology of the medium and the rheology of
the fluid. Many applications of flow in porous media involve Newtonian fluids, in which the
viscosity is independent of shear rate. Any fluid that does not obey Newtons law of viscosity
is a non-Newtonian fluid.
Generally, a fluid can be classified as Newtonian or non-Newtonian depending on its flow
behavior; that is, how its viscosity changes in the presence of shear stress and the rate of
shear applied. If the shear stress () is plotted against shear rate ( ) at constant temperature
and pressure, the so-called flow curve or rheogram (the response of a Newtonian fluid) is
a straight line with slope , passing through the origin. The constant of proportionality, ,
referred to as the Newtonian or dynamic viscosity is, by definition, independent of shear rate
and shear stress and depends only on the material, its temperature, and its pressure.
On the other hand, for a non-Newtonian fluid, the curve does not pass through the origin
and/or does not result in a linear relationship between the shear stress () and the shear rate
( ). The coefficient of viscosity is not constant and is a function of and .
The relationship between shear stress and shear rate is as follows:
=

(21)

22 Process Manual - Polymer Flood

Revision 0I

where:

Shear stress
Shear rate
Apparent viscosity of the fluid

A typical shear stress versus shear rate plot for a non-Newtonian fluid is shown in Figure 9:

Shear Stress ()

Pseudoplastic Fluid
(Shear Thinning)

Newtonian
Fluid

Dilatant Fluid
(Shear Thickening)

Shear Rate (s-1)


Figure 9 Typical Plot of Shear Stress vs. Shear Rate for Newtonian and non-Newtonian Fluids 27

Often the relationship between shear stress () and shear rate ( ) for these fluids is plotted on
log-log coordinates, and the relationship can be approximated as a straight line over a limited
range of shear rate (or stress), that is:
=

(22)

where:
Shear stress

Shear rate

Fluid consistency coefficient or index.


Flow behavior index or power-law exponent.

When viscosity decreases with increasing shear rate, the fluid is called shear-thinning. In the
opposite case, where viscosity increases as the fluid is subjected to a higher shear rate, the
fluid is called shear-thickening. Shear-thinning behavior is more common than shearthickening. Shear-thinning fluids are also referred to as pseudoplastic fluids, while shearthickening fluids are referred to as dilatant fluids. The behavior of these fluids is illustrated in
the following figures.
27

Derived from figure Rheology_of_time_independent_fluids.png, Wikimedia Commons.

Revision 0I

Process Manual - Polymer Flood 23

Apparent Viscosity

Apparent Viscosity

Shear-Thinning Fluid

Shear-Thickening Fluid

Shear Rate

Shear Rate

Figure 10 Apparent Viscosity vs. Shear Rate for Shear-Thinning and Shear-Thickening Fluids 28

Many shear-thinning and shear-thickening fluids exhibit Newtonian behavior at extreme


shear rates, both low and high. For such fluids, when the apparent viscosity is plotted against
log shear rate, the curve appears as follows:

log Apparent Viscosity

Newtonian Region

Power-Law Region

Shear-Thinning
Fluid

Newtonian Region

log Shear Rate

Figure 11 Apparent Viscosity vs. Shear Rate for a Shear-Thinning Fluid28

28

Subramanian, R. Shankar, Non-Newtonian Flows, Department of Chemical and Biomolecular


Engineering, Clarkson University.

24 Process Manual - Polymer Flood

Revision 0I

log Apparent Viscosity

Newtonian Region

Power-Law Region

Shear-Thickening
Fluid

Newtonian Region

log Shear Rate

Figure 12 Apparent Viscosity vs. Shear Rate for a Shear-Thickening Fluid

The regions where the apparent viscosity is approximately constant are known as Newtonian
regions. The behavior between these regions can usually be approximated by a straight line
on these axes. In this region, which is known as the power-law region, the behavior can be
approximated by the following expression:
= + ( )

(23)

This can be rewritten as:


= ()

(24)

Equation (24) can be replaced by one more commonly used in the literature, which comes
from combining equations (21) and (22):
= 1

(25)

K represents the fluid consistency index and is equal to () in equation (24).


represents the power-law exponent, and the expression ( 1) is equal to b in equation
(24). Note the following:
When n < 1, the fluid exhibits shear-thinning properties.
When n = 1, the fluid shows Newtonian behavior.
When n > 1, the fluid shows shear-thickening behavior.
The above shows that rheology in porous media has an important impact on enhanced oil
recovery projects where polymer solutions are used, because if the effective viscosity of the
polymer solution is high at the high velocities experienced near an injection well, the polymer
injection rate and/or the oil production rate (at the producer well) may be decreased. On the
other hand, if the effective viscosity of the polymer solution is too low at the low velocities
experienced away from the injection well or deep into the reservoir, oil displacement may be
inefficient.

Revision 0I

Process Manual - Polymer Flood 25

To simulate the behavior of the pseudoplastic or dilatant fluids used in enhanced oil recovery
(foam, high molecular weight liquids which include solutions of polymers, as well as liquids
in which fine particles are suspended [suspensions]), STARS uses the power-law or Ostwald
de Waele model, as described below.
Power-Law Expression for Shear-Thinning or Pseudoplastic Fluids
The *SHEARTHIN keyword is used to represent the pseudoplastic behavior or the shearthinning effect as follows:
*SHEARTHIN ,

DEFINITIONS:
Power-law index or exponent in the viscosity shear thinning equation

(dimensionless). The allowed range is from 0.1 to 0.99, inclusive. Values


below 0.3 can result in unacceptable numerical performance and so are not
recommended. Values close to 1 approximate Newtonian behavior.
,
The meaning of , depends on *SHEAREFFEC (*SHV | *SHR).
*SHV: Reference Darcy velocity (m/day | ft/day | cm/min) in viscosity
shear thinning equation. The allowed range is 10-10 to 1010 m/day
(3.2810-10 ft/day to 3.281010 ft/day | 6.9410-12 to 6.94108 cm/min).
*SHR: Reference shear rate (1/day | 1/day | 1/min) in viscosity shear
thinning equation. The allowed range is 10-10 to 1010 1/day (6.9410-14 to
6.94106 1/min).
The bounded power-law relation between the apparent fluid viscosity and the Darcy
fluid velocity is:

,
for ,

= ,

for , < < ,


,

,0
for ,

(26)

The upper velocity boundary of the shear thinning regime, , , is defined by the point on
the power-law curve where the apparent viscosity, , equals the phase fluid viscosity in
the absence of polymer (,0 ). The lower velocity boundary of the shear thinning regime,
, , is defined by the point on the power-law curve where the apparent viscosity, ,
equals the fluid phase viscosity in the absence of thinning (, ). For further discussion on the
calculation of phase viscosities for Newtonian flow, refer to the STARS Users Guide for
information on *AVISC and *BVISC.
The bounded power law relation of apparent viscosity versus velocity for shear thinning is
depicted in the log/log plot of Figure 13. The shear thinning regime is represented by a linear
relation of slope ( 1).

26 Process Manual - Polymer Flood

Revision 0I

app = l,p

log (Apparent
Viscosity)

app = l,0
ul = ul,upper

ul = ul,lower

log (Darcy Velocity)

Figure 13 Shear Thinning Power Law - Apparent Viscosity vs. Darcy Velocity

Power-Law Expression for Shear-Thickening or Dilatant Fluids


The *SHEARTHICK keyword is used to represent the dilatants behavior or the shearthickening effect as follows:
*SHEARTHICK , ,

DEFINITIONS:
Power-law index or exponent in the viscosity shear thickening equation

(dimensionless). The allowed range is from 1.01 to 5, inclusive. Values


above 2.5 can result in unacceptable numerical performance and so are not
recommended. Values close to 1 approximate Newtonian behavior.
,

The meaning of , depends on *SHEAREFFEC (*SHV | *SHR).


*SHV: Reference Darcy velocity (m/day | ft/day | cm/min) in viscosity
shear thickening equation. The allowed range is 10-10 to 1010 m/day
(3.2810-10 ft/day to 3.281010 ft/day | 6.9410-12 to 6.94108 cm/min).
*SHR: Reference shear rate (1/day | 1/day | 1/min) in viscosity shear
thickening equation. The allowed range is 10-10 to 1010 1/day (6.9410-14 to
6.94106 1/min).

Maximum viscosity (cp) in viscosity shear thickening equation. The


allowed range is 10-5 to 106 cp.

Revision 0I

Process Manual - Polymer Flood 27

The power-law relation between apparent fluid viscosity and Darcy fluid velocity is:

,
for ,

= ,

for , < < ,


,

,
for ,

(27)

The lower velocity boundary of the shear thickening regime, , , is defined by the point
on the power law curve when the apparent viscosity, , equals the phase fluid viscosity in
the absence of thickening (, ). For further discussion on the calculation of phase viscosities
for Newtonian flow, refer to the STARS Users Guide for information about *AVISC and
*BVISC.
The upper velocity boundary of the shear thickening regime, , , is defined by the point on
the power-law curve where the apparent viscosity, , equals the user-defined maximum
viscosity (, ).

The bounded power-law relation of apparent viscosity versus velocity for shear thickening is
depicted in the log/log plot of Figure 14. The shear thickening regime is represented by a
linear relation of slope ( 1).
app = l,max

log (Apparent
Viscosity)

app = l,p

ul = ul,lower

ul = ul,max

log (Darcy Velocity)

Figure 14 Shear Thickening Power Law - Apparent Viscosity vs. Darcy Velocity

In cases where fluids exhibit both behaviors (shear-thinning and shear thickening), the
keywords can be used together. The apparent viscosity of the combined effect is the sum of
the shear thinning and thickening apparent viscosities defined in the above sections:
= , + ,

28 Process Manual - Polymer Flood

(28)

Revision 0I

The summed power-law relation between apparent fluid viscosity, , and Darcy fluid
velocity, , is:

,
for ,
= , + , for , < < ,
,
for ,

(29)

The lower velocity boundary of the shear thinning and thickening regime, , , is defined
by the point on the thinning power-law curve where the apparent viscosity, , equals the
fluid phase viscosity in the absence of thinning (, ). The upper velocity boundary of the
shear thinning and thickening regime, , , is defined by the point on the thickening powerlaw curve where the apparent viscosity, , equals the user-defined maximum viscosity
(, ).
The summed power-law relation of apparent viscosity versus velocity is depicted in the
log/log plot of Figure 15.

app = l,max

log (Apparent
Viscosity)

app = l,p

ul = ul,lower

ul = ul,max

log (Darcy Velocity)

Figure 15 Shear Thinning and Thickening Power Laws - Apparent Viscosity vs. Darcy Velocity

The above explanations are for velocity-dependent viscosity, which is the default option
implemented in STARS. For the shear-rate-dependent viscosity option, the same logic
applies, with the term shear rate replacing velocity throughout.
If you need to replace the default option (velocity-dependent viscosity) with the shear-ratedependent viscosity option, the *SHEAREFFEC keyword must be used. The format of this
keyword is as follows:
*SHEAREFFEC (*SHV | *SHR)

Revision 0I

Process Manual - Polymer Flood 29

DEFINITIONS:
*SHV

Viscosity shear depends on Darcy velocity.

*SHR

Viscosity shear depends on shear rate.

DEFAULTS:
If *SHEAREFFEC is absent, *SHEAREFFEC *SHV is assumed.
Tabular Input Option for Velocity/Shear-Rate-Dependent Viscosity
In addition to the keywords described above, STARS has a tabular input option for velocitydependent viscosity or shear-rate-dependent viscosity, which is useful when the viscosityversus-velocity relation or viscosity-versus-shear-rate relation is specified by laboratory data,
or when a simple power-law relation is not sufficient. The format is as follows:
FORMAT:
*SHEARTAB
{ velocity viscosity }
or
*SHEARTAB
{ shear-rate viscosity }
DEFINITIONS:
*SHEARTAB A viscosity-versus-velocity table follows. The maximum allowed number
of table rows is 40. The first column is either velocity or shear rate,
depending on *SHEAREFFEC.
Velocity

*SHEAREFFEC 0: The first column is phase velocity (m/day | ft/day |


cm/min). The allowed range is 10-10 to 1010 m/day (3.2810-10 to 3.281010
ft/day | 6.9410-12 to 6.94108 cm/min).

shear-rate

*SHEAREFFEC 1: The first column is phase shear-rate (1/day | 1/day |


1/min). The allowed range is 10-10 to 1010 1/day (6.9410-14 to 6.94106
1/min).

viscosity

Viscosity (cp) of the component at corresponding velocity. The allowed


range is 10-5 to 106 cp.

The following conditions apply to these keywords:


a) *SHEARTAB is applied to the component and phase specified by the immediately
preceding *VSMIXCOMP, so *VSMIXCOMP must be present before
*SHEARTAB.
b) *SHEARTAB may not be used together with *SHEARTHICK and
*SHEARTHIN.
c) The first column is either velocity or shear rate, depending on *SHEAREFFEC.
d) For phase velocity/shear-rate outside the velocity/shear-rate table range, the nearest
viscosity table entry is used.

30 Process Manual - Polymer Flood

Revision 0I

Velocity Dependent Skin Factor


In many cases, much of the shear thinning or thickening occurs within the injection well grid
block. Although non-Newtonian flow between grid blocks is accounted for, modeling of nonNewtonian flow within the well grid block requires a user-input skin factor. The following is
a discussion of the standard radial inflow well model based on Newtonian flow and a
derivation of the recommended skin factor that can be used to correct for actual nonNewtonian flow within the well grid block.
For model input protocol and a detailed discussion of well skin factors, refer to Well
Element Geometry (Conditional) of section Well and Recurrent Data and Radial Inflow
Well Model of section Appendix A: Well Model Details in the STARS Users Guide.
The Radial Inflow Well Model is based on a steady-state Newtonian flow which incorporates
a dimensionless pressure-drop skin factor. This equation uses an equivalent well block radius,
, defined as the radius at which the steady-state flowing pressure of the actual well is equal
to the numerically calculated pressure for the well block.

The pressure drop between the well bottom-hole pressure and the well block pressure, , can
be represented by:

= + (30)

A similar equation can be written for a velocity-dependent viscosity and non-Newtonian


flow29:
=

1

[1 1 ] (31)
2
1

where , and are the power index, reference viscosity and reference velocity.

The user must determine whether shear thinning or shear thickening is dominating in the well
grid block and use the appropriate power index and reference parameters from the power law
relations. For example, if thinning is the dominant (or only) velocity-dependent viscosity
effect, then = , = , , and = , .
A skin factor to account for the pressure drop difference between Newtonian and nonNewtonian flow can be determined by equating the above equations:
1
1

=
1 + (32)
1

29

Odeh, A.S., and Yang, H.T., Flow of Non-Newtonian Power-Law Fluids Through Porous Media,
SPEJ, June 1979, pp. 155-163.

Revision 0I

Process Manual - Polymer Flood 31

where:
1

(33)

and:

(34)

This skin factor should provide the necessary correction to the Newtonian Radial Inflow Well
Model to account for non-Newtonian flow within the well grid block. The equation can be
solved by assuming that the equivalent well block radius for Newtonian and non-Newtonian
flow is similar. However, technically these values will be different and a more accurate
determination of the well block radius for non-Newtonian flow can be done using the
velocity-dependent power law for viscosity and a derivation similar to the one outlined in
Peaceman 30.
Convert Shear Rates to Velocities
The information for converting shear rate data from laboratory measurements to Darcy
velocities comes from a study by Cannella et al.31 The equation relating the effective porous
media shear rate and the fluid Darcy velocity is as follows:
=

| |

(35)

where and are the absolute permeability and porosity, and , , and are the phase
Darcy velocity, relative permeability, and saturation respectively.
The shear rate factor is given by:

3 + 1 1
=

(36)

where is the shear thinning power exponent and is a constant value, usually equal to 6
and related to the tortuosity of the porous medium. The default shear rate factor of 4.8
corresponds to C = 6 and n = 0.5; however, the user can adjust this conversion via the
*SHEAR_FAC keyword as follows:
FORMAT:
*SHEAR_FAC factor
DEFINITIONS:
factor
30

Factor in the shear rate equation described above.

Peaceman, D.W., Interpretation of Well-Block Pressures in Numerical Reservoir Simulation with


Non-Square Grid Blocks and Anisotropic Permeability, SPEJ, June 1983, pp. 531.
31
Cannella, W., Huh, C., and Seright, R.S., Prediction of Xanthan Rheology in Porous Media, paper
SPE 18089, presented at the 63rd Annual Tech. Conference SPE, Houston, Texas, October 1988.

32 Process Manual - Polymer Flood

Revision 0I

DEFAULTS:
If *SHEAR_FAC is absent then = 4.8 is assumed.

If velocity-dependent viscosity is used, the reference velocity applies to a reference


permeability. Thus, the relevant parts of the reservoir should be adequately represented by an
average permeability. If the reservoir is highly non-uniform in permeability, the use of a
shear-rate dependent viscosity model is recommended.
An example of Darcy velocity calculation using the *SHEARTAB keyword is shown in
Shear Effect on Polymer Viscosity in Appendix A Equations on page 58.

Salinity-Dependent Viscosity Effects


Consideration for brine salinity is important for polymer selection. For example, hydrolyzed
polyacrylamide (HPAM) is very sensitive to salinity and hardness, so the viscosity
improvement property will be significantly reduced when it dissolves in water with high
salinity or hardness. This characteristic represents the disadvantage of using this polymer. On
the contrary, the biopolymer Xanthan is more tolerant to salinity or hardness than HPAM,
which is its main advantage. If the formation water is of high salinity, the polymer should be
stable to salt concentration or else a preflush of fresh water must be injected to precondition
the reservoir and prevent the polymer from losing most of its viscosity.
To simulate this effect, the STARS simulator has a power-law equation that uses the viscosity
of the key component in the non-linear mixing rule (*AVISC, *BVISC or *VISCTABLE).
The ratio of salt concentration is as follows:

=
where:

for

0 for
>

(37)

Salinity component mole/mass fraction below which the nonlinear mixing


component viscosity is considered independent of salinity. In other words,
is the enforced minimum salinity of the phase. The allowed range is
10-5 to 10-1.
Salinity component mole/mass fraction.

Polymer viscosity as pure component (for information about the *AVISC or


*VISCTABLE keywords, refer to the STARS Users Guide).

Resultant polymer component viscosity in the saline solution. This new


polymer component viscosity is then used with the nonlinear mixing
function to determine the phase viscosities.

Slope on a log-log plot of polymer component viscosity versus ratio of

salinity, . The allowed range is -100 to 100.

Revision 0I

Process Manual - Polymer Flood 33

For further information, refer to the *VSSALTCMP keyword in the STARS Users Guide. For
an example of the calculation, refer to the Polymer Viscosity subsection of the Polymer Flood
Calculations section in Appendix A Equations of this manual.

Lab and Field Information


Before an EOR method can be applied to a particular well or reservoir, laboratory and field
measurements must be carried out to determine the optimal method and formulation. For a
polymer flood process, for example, the laboratory tests will consist of polymer injectivity,
retention, stability, fluid compatibility, and effective viscosity measurements, while field
testing will include injectivity, biological stability, and in-situ viscosity measurements 32.
The main objective in a polymer flood process is the selection of the polymer most suitable
for the reservoir conditions, based on temperature stability, salinity tolerance, quality of the
mixing water, and economic conditions. To correctly choose the polymer, a detailed
screening program should be performed to evaluate the basic fluid rheological properties and
quality, such as screen factor tests, filtration ratio, and viscosity measurement.
Once the polymer is selected, core-flooding experiments should be performed as the second
stage of the experimental program. The measured displacements are used to describe the
rock-polymer interactions, based on polymer adsorption, resistance factor and residual
resistance factor. With this information, polymer concentration in the effluents can be
determined, as well as the water and/or oil volumes produced.
Finally, a thermal aging study may be needed to verify that the selected polymer does not
degrade significantly at reservoir temperature over time 33.
In summary, the parameters needed to model a polymer flood simulation at the laboratory
and/or field scale are as follows:

Polymer Screening
- Viscosity (dependent on shear rate and concentration)
- Molecular weight
- Polymer stability (thermal, biological, mechanical, and/or chemical)
Polymer screening is used to determine the polymer half-life, which
in turn, is used to generate the degradation reactions explained in
Polymer Degradation on page 14.
Core flooding
- SCAL for understanding rock and fluid properties:
Porosity
Permeability
Relative permeability curves (initial and residual saturations)

32

Castagno, R.E., Shupe, R. D., Gregory, M. D. and Lescarboura, J. A., A Method for Laboratory and
Field Evaluation of a Proposed Polymer Flood, SPE 13124, presented at the 59th Annual technical
Conference and Exhibition, Houston, Texas, September 16-19, 1984.
33
Saavedra, N.F., Gaviria, W, and Davitt, H.J., Laboratory Testing of Polymer Flood Candidates: San
Francisco Field, prepared for the SPE/DOE Improved Oil Recovery Symposium, Tulsa, Oklahoma,
April 13-14, 2002 SPE 75182.

34 Process Manual - Polymer Flood

Revision 0I

Oil and water properties (densities, viscosities, and so on)


Pressure taps on core
Polymer concentration and slug size
Flow rates
Polymer adsorption (static or dynamic)
Polymer concentration of effluent
Viscosity of effluent over the range of injection rates (shear rates)
To be compared and matched with those measured in the rheometer.
Residual resistance factor and permeability reduction
Oil and water produced volumes

Using the Process Wizard to Model a Polymer Flood


To use the Process Wizard to build a model for a polymer flood:
1. Open the dataset in Builder.
2. Click Components in the menu bar then select Process Wizard. The Process
Wizard Step 1 - Choose Process dialog box is displayed:

Revision 0I

Process Manual - Polymer Flood 35

3. Select Alkaline, surfactant, foam and/or polymer model:

4. Click Next. The Step 2 - Input Specific Data For A.S.P. Models dialog box is
displayed:

36 Process Manual - Polymer Flood

Revision 0I

5. Select the Polymer flood (add 1 component) model:

Polymer adsorption
Fraction of the pore
volume accessible to the
polymer molecules
Type and/or density of
rock for the adsorption
conversion

Residual resistance
factor from lab
Polymer consumption
and/or degradation with
time
Enter polymer half life

6. Configure the model options as directed above then click Next. The Step 3 Component Selection dialog box is displayed:

Step 3 of the Process Wizard allows you to either add a new polymer component if
one has not been defined, as shown below:

or to select or update the polymer component that has already been defined, as
shown below:

If you select the check box, a new component will be added. If you leave the check
box blank, you can use the drop-down menu below it to select the component
whose properties you want to change or update.

Revision 0I

Process Manual - Polymer Flood 37

7. If necessary, add components then click Next. The Step 4 - Set Adsorption
Values dialog box is displayed:
Porosity used in the
adsorption lab
experiment
Polymer
concentration in
weight %

Polymer adsorption
reported by lab
(mg / 100g of rock)

8. As necessary, enter the polymer adsorption values as directed above then click
Next. The Step 5 - Set Polymer Values dialog box is displayed:

Polymer
concentration in
weight %

Polymer solution
viscosity (cps)

9. Enter the polymer viscosity values as shown above then click Finish. If, in
Step 3 Component Selection, you changed or updated the polymer component
that was already defined and the polymer molecular weight is now greater than
8 kg/gmmole (or 8000 lb/lbmole), the following message will be displayed:

10. As stated in the message, click Yes to use the recommended polymer weight of
8 kg/gmmole (8000 lb/lbmole), or click No to continue with the current value. The
Process Wizard calculations are performed, the dataset is updated, and the main
Builder window becomes active.

38 Process Manual - Polymer Flood

Revision 0I

Viewing and Adjusting the Process Wizard Results


Components Generation
To check (and if necessary, edit) the polymer components that have been generated by the
Process Wizard, click Component | Add/Edit Components in the menu bar. The
Components and Phase Properties dialog box is displayed:

Polymer
definition

Polymer Consumption Reaction


To view details of the polymer consumption reaction, click Component | Reactions in the
menu bar. The Reactions dialog box is displayed, as shown in the following example:

Kinetic
reaction rate

Stoichiometric
coefficients
Reaction orders
Phase of the
reaction
Summary of the
reaction

For information about the formulas used in determining the volumetric reaction rate, refer to
Volumetric Reaction Rate on page 46. For information about the formulas used to calculate
the stoichiometric coefficients, refer to Stoichiometric Coefficients on page 48.

Revision 0I

Process Manual - Polymer Flood 39

Polymer Adsorption
Polymer Adsorption Table
To view the details of the polymer adsorption table, click Rock-Fluid | Components in the
menu bar. The Component Adsorption dialog box is displayed:
Modify adsorption
settings

Component i

Phase where
component i is
encountered

Component ranges

Adsorption ranges
Phase to apply the
permeability reduction

Adsorption rock
types
Maximum
adsorption

Residual adsorption

1. Totally reversible
(ADRT = 0)
2. Partially reversible
(0 < ADRT < ADMAXT)
3. Totally irreversible
(ADRT = ADMAXT)

Fraction of pore
volume accessible
Residual resistance
factor

For information about the calculations used to produce the polymer adsorption table, refer to
Polymer Adsorption Table on page 49.
Langmuir Isotherm Option
Click the Change Adsorption Options button to change the adsorption dependence to
Langmuir isotherm coefficients. The Component Adsorption dialog box will change as
shown below, with a table for entering the Langmuir isotherm coefficients:
Component i

Component ranges

Modify adsorption
settings
Phase where
component i is
encountered
tad1

Phase to apply the


permeability reduction
Residual adsorption

1. Totally reversible
(ADRT = 0)
2. Partially reversible
(0 < ADRT < ADMAXT)
3. Totally irreversible
(ADRT = ADMAXT)

tad2
tad3
Adsorption rock types
Maximum adsorption
Fraction of pore
volume accessible
Residual resistance
factor

For information about the Langmuir isotherm calculations, refer to Langmuir Isotherm
Option on page 50.
Note: When one value of adsorption is used, Builder calculates the corresponding adsorption
values for intermediate concentrations using a linear interpolation, which will not follow a
Langmuir shape.

40 Process Manual - Polymer Flood

Revision 0I

For information about the adsorption equations used in the Permeability Reduction
calculations, refer to Permeability Reduction on page 52.

Polymer Viscosity
To view the polymer viscosity results, click Component | Add/Edit Components in the
menu bar then select the Liquid phase viscosities tab:

Polymer viscosity
(as pure component)
calculated by the Wizard
Non-linear mixing rule
for the viscosity

Note: A value of 0 for BVISC indicates that the components viscosity does not change with
temperature.
To specify the nonlinear viscosity mixing rule, click Specify. The Nonlinear Mixing dialog
box is displayed:

Maximum polymer
concentration is
extrapolated to consider
future optimizations

Viscosity function
calculated by the Wizard

For information about the calculations used to model polymer viscosity, refer to Polymer
Viscosity on page 53.

Revision 0I

Process Manual - Polymer Flood 41

Other Considerations and Troubleshooting Information


This section includes considerations that you may need to make, and information to help you
troubleshoot problems with polymer flood models.

High Molecular Weight (or Low Mole Fractions)


STARS uses numerical derivatives with a default concentration shift of 10e-7 instead of
algebraic derivatives to form the flow equations. As a result, the use of polymers of high
molecular weight (for example 10e+6) might reduce the numerical performance of the
simulator. This is because, in a mole-fraction-based model, the polymer concentration values
are expected to be very low (for example 10e-9). Refer to Polymer Flood Calculations on
page 46 for information about the mole fraction calculation.
If numerical shifts are larger than the injected mole fractions, inexact derivatives may occur,
which will interfere with convergence. As well, numerical performance may degrade as the
chase water pushes the polymer out of the reservoir at the end of the simulation and polymer
concentrations are reduced further, resulting in poorer derivative estimates.
There are four ways to address this problem. The first way is to change (decrease) the default
convergence tolerance for *ZAQ, which ties the numerical shift level to the convergence
tolerance, so that decreasing one decreases the other. By making this modification in the
numerical section, along with reducing the maximum time step size *DTMAX and limiting
the number of wasted Newton iterations by setting *NEWTONCYC < default value, it is
possible to improve the run. A good approximation can be obtained by using the following
numerical parameters:
*NUMERICAL
*NORM *ZAQ 0.075
*CONVERGE *ZAQ 5.0e-4
*DTMAX 15.0
*NEWTONCYC 8
The second way is to define a pseudo-polymer of reduced molecular weight but with the
same viscosity and adsorption effects as for a higher mole fraction. This pseudo-polymer
mole fraction approach will preserve the mass fraction used by the true polymer mole fraction
model. Refer to Using the Process Wizard to Model a Polymer Flood starting on page 35 for
information about defining the recommended pseudo-polymer molecular weight.
The third way to address this problem is to run the model with mass fraction concentrations
by using the *MASSBASIS option. If the laboratory simulations are run this way, we
recommend that the field model also be run this way, particularly if no gas phase is generated
and the field GOR stays at its initial value, in which case there is little motivation for mole
fraction representation. Note that using *MASSBASIS will require converting the solution
gas oil mole fraction composition to mass fraction, and also conversion of the solution gas kvalues.
The fourth way is to use the *SHIFT MWT keyword (refer to the latest STARS Users Guide
for information about this keyword).

42 Process Manual - Polymer Flood

Revision 0I

Disproportionate Permeability Reduction Effect


As mentioned in previous sections, some polymers have the ability to reduce the permeability
to water more than that to hydrocarbons. This phenomenon is known as the disproportionate
permeability reduction or DPR effect.
STARS does not explicitly handle this effect by using residual resistance factors for each
phase. This section provides options and tools for simulating the permeability reduction of
the phases by using different blockage alternatives. The four options are described below.
The first option is to obtain pseudo residual resistance factors for the each of the phases by
dividing the values obtained from coreflooding experiments by one of them, i.e., a reference
residual resistance factor. Normally, the reference residual resistance factor would be that of
the oil because the oleic phase is not altered as much as the aqueous phase is, in which case:
=
=

> 1 because normally >

=1

(38)

(39)

no resistance effect, and it is not necessary to assign it to the simulator

The second alternative is to reduce the absolute permeability of all phases by first using the
STARS variable permeability option and then further modifying the water mobility with the
residual resistance factor. Here, the absolute permeability is a function of porosity. As the
porosity is reduced by the adsorbed polymer, the permeability should be reduced as follows:
Carmen-Kozeny type formula using keyword *PERMCK

A permeability multiplier obtained from table look-up through the keyword


*PERMTAB
A user-defined multiplier factor specified by keywords *PERMULI, *PERMULJ
or *PERMULK

The third option is to use the relative permeability interpolation method to modify only the
relative permeability curve to oil, considering the effect in this phase. The residual resistance
factor to water is then used as the default for modifying the mobility to water.
The fourth option involves the creation of a first-order reaction where the polymer is divided
into two new components or pseudo-polymers. One of these components will reduce the
mobility to water (by using the water residual resistance factor) and the other will reduce the
mobility to oil (by using the oil residual resistance factor).
+

(40)

The stoichiometric coefficients for balancing the reaction can be obtained from the residual
resistance factor information using the following equations:
= +
Revision 0I

(41)

Process Manual - Polymer Flood 43

=
=

<1

<1

(42)
(43)

The adsorption must be adjusted and care should be taken in the water viscosity calculation
with polymer concentration.

Interpreting Polymer Flood Model Outputs


To interpret the chemical flood results in Results 3D, first add the following grid outputs
through IO Control | Simulation Results Output:

ADSORP: Component adsorbed


ADSPCMP: Composition of key component used in the calculation of adsorbing
component
KRINTER: Relative perm interpolation value
SHEARO: Shear rate of oil phase
SHEARW: Shear rate of water phase
SHEARG: Shear rate of gas phase
SHEARSTRG: Shear stress (viscosity times shear rate) of gas phase
SHEARSTRW: Shear stress (viscosity times shear rate) of water phase
SHEARSTRO: Shear stress (viscosity times shear rate) of oil phase
VISCVELW: Magnitude of Darcy velocity of water phase
VISCVELO: Magnitude of Darcy velocity of oil phase
VISCVELG: Magnitude of Darcy velocity of gas phase
VISWCOM: Component composition of the key component in the nonlinear
mixing of water viscosity given by *VSMIXCOMP

After you have output the above grid variables, the following plots are required, on a gridblock basis, wherever there is chemical concentration and are relatively significant for
analyzing the chemical flood results:

ADSORP vs. time: The mole fraction of component that has been adsorbed in
rock as a function of time.
KRINTERP vs. time: Used to analyze how the relative permeability interpolation
parameter is changing with time, and the set of relative permeability to which
STARS is referring.
Output Mole Fraction vs. time: To show what is happening inside the reservoir.
VISCVELW vs. SHEARW: To analyze shearing in the water phase at respective
Darcy velocities.

44 Process Manual - Polymer Flood

Revision 0I

Appendix A Equations
Introduction
This chapter provides additional information about the equations used by the Process Wizard,
including descriptions of the variables and constants used.
Note: To avoid confusion, all windows and calculations in this manual have been shown or
made in *Field Units. For further information about the specific units used, refer to the
STARS Users Guide.

General Formulas
The following general formulas are used by the Process Wizard:
Parts-per-million (ppm)

1 ppm = 1 mg/l
Weight Percentage

wt% =

ppm
1 10 4

Mass Fraction (wt)

wt =

ppm
wt%
=
6
100
1 10

Mole Fraction (dim)

(wt /Mw )
i
i
x =
i n
c
wti /Mwi
i =1

Revision 0I

(44)

Process Manual - Polymer Flood 45

where:
wti

Mass fraction of component i

Mwi
nc

Molecular Weight of component i


Number of components in the solution

Polymer Flood Calculations


Assuming a polymer solution (water + polymer) with a polymer concentration of 0.1 wt%
and a molecular weight of 8000 lb/lbmol, the mass and mole fraction of the polymer are
calculated as follows:
Polymer Mass Fraction

wt =

0.1
= 0.001
100

Polymer Mole Fraction

xi =

(0.001/8000 ) = 2.25412 10 6
0.001 (1 0.001)
+
8000

18.015

To simulate the polymer degradation and/or consumption in STARS, use the reaction module
and input the properties required to express the volumetric reaction rate.
Volumetric Reaction Rate
A kinetic reaction provides information about the speed with which the reaction proceeds.
The expression for the volumetric reaction rate in STARS is as follows:
nc
Ea
rrf Ciai exp

R T
i =1

(45)

The concentration factor for component i in the liquid phase is usually based on density, as
shown in the following formulas:

C i = f S xi

k =1

f = v 1

C sk
sk (P, T )

(46)
(47)

In the above equations, void porosity, v, is corrected for pore pressure and temperature, and
the effective porosity to fluids, f, is corrected for the volume of the solid phase in the pore
space. Each component k in the solid phase has concentration and density (, ).
For further information, refer to keyword *SOLID_DEN in the STARS Users Guide.

46 Process Manual - Polymer Flood

Revision 0I

The concentration factor for component i in the solid phase, if one exists, is:

C i = v C si

(48)

The following constants and variables are used in the above formulas:
rrf

Constant factor in the expression, known as the kinetic constant

Ea

Activation energy, gives the dependence of reaction rate on temperature

Temperature

Universal gas constant

Ci

Concentration factor contributed by reactant component i

ai

Order of reaction with respect to component i

Effective porosity to fluids

Saturation of the phase where component i is reacting

Molar density of the phase where component i is reacting

xi

Mole fraction of the component i from the phase where it is reacting

Void porosity

Csk

Concentration of component k in the solid phase

sk (P, T )

Density of component k in the solid phase

Csi

Concentration of component i in the solid phase

Ci

Actual concentration of component i in the porous space

For the simple case of polymer consumption, the reaction is:

Sto1 Polymer( w ) Sto 2 Water( w )

(49)

where:

Sto1
Sto 2

Stoichiometric coefficient of reacting component


Stoichiometric coefficient of produced component

The kinetic rate for this first-order reaction is typically expressed as:

rrf =

Revision 0I

Ln(2 ) 0.69315
=
t1/2
t1/2

(50)

Process Manual - Polymer Flood 47

where:

rrf

Reaction frequency factor (day-1)

t1 / 2

Half-life (days)

If the half-life of the polymer uses the Process Wizard default value of 180 days, the reaction
frequency factor will have the following value:

rrf =

0.69315
= 0.00385 day 1
180

(51)

Stoichiometric Coefficients
Typically, the stoichiometric coefficients are based on 1 mole of one of the reactive
components. The user is responsible for ensuring that the stoichiometric coefficients entered
as data represent a mass-conserving set. A set of mass-conserving coefficients will satisfy:
nc

nc

Mw Sto1 = Mw Sto2
i =1

i =1

(52)

where:

Mwi

Molecular weight of component i

Sto1i

Stoichiometric coefficient of component i as reactant

Sto2i

Stoichiometric coefficient of component i as product

Assuming the following reaction:

aA + bB cC

(53)

The stoichiometric coefficients can be calculated as follows:

a
ma
Mwb
=

b Mwa mb

(54)

a
ma Mwc
=

c Mwa mc

(55)

where:

a, b, c

Moles of components A, B, and C respectively

ma , mb, mc

Mass of components A, B, and C respectively

The stoichiometric coefficients for the first-order reaction:

Sto1 Polymer( w ) Sto2 Water( w ) (56)

48 Process Manual - Polymer Flood

Revision 0I

are as follows:

MwW
mP
Sto1
=

Sto 2 Mw p
mW

(57)

Reorganizing:

MwP
Sto 2 = Sto1
MwW

mW

mP

(58)

The mass fraction is used more often than the mass of the components for this type of
simulation, so the expression is reorganized using:

mW

mP

wtW
=
wt P

(59)

Substituting into equation (58):

MwP
Sto 2 = Sto1
MwW

wtW

wt P

(60)

Usually the stoichiometric coefficient for the main reacting component is chosen to be equal
to 1, as follows:

8000 1
Sto2 = 1
= 444.074
18.015 1

(61)

The final expression for this example is:

Polymer( w ) 444.074 Water( w )

(62)

Polymer Adsorption Table


The form of the adsorption term in flow equations is generally expressed as:

[Ad i ] (63)
t
For this reason, the adsorption levels in the simulator are described as moles (or mass) of
component i adsorbed per unit volume; however, other measures of adsorption level can be
used and these need to be converted to meet the simulator input requirements:

Ad iStars = Ad ilab

Revision 0I

r (1 )

(64)

Process Manual - Polymer Flood 49

where:

Ad iStars

Adsorption of component i, to be used in STARS (gmol/m3, lbmol/ft3, or


gmol/cm3)

Ad ilab

Adsorption obtained in laboratory (mg polymer/100 gr-rock)

Rock density (gr/cm)

Porosity

Using the data assigned to the Process Wizard, the appropriate conversion is as follows:

Ad iStars = 50

mg Polymer
100grrock

2.65

g
(1 0.2494 )
cm 3
Cf
0.2494

(65)

where:

Cf

Unit conversion factor

In the following example, Cf is used to convert to lbmol/ft3:

(30.48 )3 cm 3

1 lb

Cf =
453592.37
mg
1 ft 3

3
1 lbmol
6 lbmol cm

= 7.8 10

mg ft 3
8000 lb

Substituting:

Ad iStars = 3.11184 10 5 lbmol/ft 3


Langmuir Isotherm Option
The Langmuir adsorption isotherm gives the adsorbed moles of component i per unit pore
volume, as follows:

Ad iStars =

(tad1 + tad2 xnacl ) ci


(1 + tad3 ci )

50 Process Manual - Polymer Flood

(66)

Revision 0I

where:
tad1

1st parameter in the Langmuir expression (gmol/m3, lbmol/ft3, or


gmol/cm3)

tad2

2nd parameter in the expression associated with salt effects (gmol/m3,


lbmol/ft3, or gmol/cm3)

tad3

3rd parameter in the Langmuir expression (dim)

xnacl

Salinity of the brine

ci

Mole fraction of component i

The Langmuir isotherm equation can be rewritten as:

Ad iStars =

(tad1 + tad2 xnacl )


tad 3

tad 3 ci
(1 + tad3 ci )

(67)

At high concentrations (large ci), the maximum adsorption, obtained using the above
conversion, is:

(tad1 + tad2 xnacl )


tad3

(68)

In addition to the maximum adsorption, the rate of increase of adsorption with fluid
composition should be known in order to fit the Langmuir parameters tad1 and tad3. If this is
not reported, as is often the case, you can use the fluid composition at the adsorption
maximum to indirectly determine this second factor.
Assuming tad2 is negligible, the equation can be rewritten as:

tad3 ci
tad1
Ad iStars =

tad3 (1 + tad3 ci )

(69)

The first factor in the above expression is the maximum adsorption and the second factor is of
order one (1) when tad3 ci 10 . In the above example, the maximum adsorption level
occurs at 0.1 wt% of polymer which is equivalent to a mole fraction concentration of
ci = 2.2541210-6. In this case, tad3 can be calculated as follows:

tad3 =

10
ci

Substituting for ci:

tad3 =

Revision 0I

10
= 4436321.05
2.25412 10 -6

Process Manual - Polymer Flood 51

With the value of tad3 established, the adsorption maximum level can be used to determine
tad1:

tad1 = Ad iStars tad3 = 3.11184 10 5

lbmol
lbmol
4436321.05 = 138.05
3
ft
ft 3

If Ad ilStars is the maximum adsorption that can be obtained, then it should be used as
*ADMAXT. Maximum adsorption levels *ADMAXT and residual adsorption levels *ADRT
can be made region-dependent, so that these properties can be varied from grid block to grid
block. Specification of residual adsorption levels provides flexibility for modeling both
reversible (chemical, adsorption ADRT = 0.0) and irreversible (mechanical, ADRT =
ADMAXT) processes, as well as partially reversible processes (0 < ADRT < ADMAXT).
Permeability Reduction
Permeability alteration often accompanies adsorption, especially if adsorption is of the
mechanical blockage type. The simulator accounts for this through the use of resistance
factors, RRF, and region dependence, which allow for the correlation of local permeability
with local adsorption levels (Adcell). It is assumed that only single-phase flow paths are
altered. When this is the case, for each phase, the permeability reduction (equilibrium
blockage) is calculated as follows:

Rk = 1 + (RRF 1)

Ad cell
ADMAXT

(70)

where:

Rk

Permeability reduction factor of phase

RRF

Residual resistance factor of phase

Rka varies between 1.0 and a maximum of RRF, as adsorption level increases. The mobility
of the each phase is divided by Rk , so accounting for blockage:

k ef =

k r k abs
Rk

(71)

where:

k ef

Effective permeability of phase

k abs

Absolute permeability of the rock

k r

Relative permeability of phase

52 Process Manual - Polymer Flood

Revision 0I

Polymer Viscosity
The viscosities of the liquid phases are obtained by applying the following linear mixing rule:
nc

( )

Ln( ) = f i Ln i

(72)

i =1

where:

Viscosity of aqueous ( = w) or oleic ( = o) phase.

Weighting factor of component i in the aqueous ( = w) or oleic ( = o)


phase viscosity in the linear mixing rule.

Viscosity of component i in the aqueous ( = w) or oleic ( = o) phase,


obtained from the correlation of viscosity or from the table assigned to
simulator. For further information, refer to the keywords *AVISC, *BVISC or
*VISCTABLE in the STARS Users Guide.

nc

Number of components in the oleic or aqueous phase.

The weighting factors, , in the linear mixing rule expression are normally expressed as
mole (mass) fractions of the components in the respective phases, i.e., wi in the aqueous
phase and xi in the oleic phase.
In the case where the viscosity of the phase is nonlinear (for example, in the presence of
solution gas or polymers), the simulator uses a modified mixing rule to represent these
changes in viscosity. This nonlinear mixing option partitions all the components into two
groups: those that are key components specified by *VSMIXCOMP (call it set S), and those
are that are not. The mole (mass) fractions ( or ) of the two groups sum to 1:

=1

=1

+ = 1

where:

=1

=1

+ = 1

(73)

Number of key components in the liquid phase.


Number of components in the liquid phase that are not key components.

In the modified mixing rule the weighting factors of the linear expression, are replaced
with alternate weighting factors: a nonlinear mixing function for each component i=S
and with for each component iS, where N is a normalizing factor that comes from the
requirement that the alternate weighting factors still sum to 1 and are derived as follows:

=1

=1

+ = 1

Revision 0I

=1

=1

+ = 1 (74)

Process Manual - Polymer Flood 53

=1

=1

(75)

= 1

Therefore the nonlinear mixing rule for the liquid viscosity has the following expression:

=1

=1

( ) = + (76)

where:

Viscosity of mixture aqueous ( = w) or oleic ( = o).


Viscosity of component i in the aqueous ( = w) or oleic ( = o) phase.
Weighting factor of the non-key component i in the aqueous ( = w) or oleic ( =
o) phase viscosity in the nonlinear mixing rule.
Weighting factor of the key component i in the aqueous ( = w) or oleic ( = o)
phase viscosity in the nonlinear mixing rule.
Number of key components in the liquid phase.
Number of components in the liquid phase that are not key components.
Normalizing factor.

Generating Data for the Polymer Case in the Aqueous Phase

Let polymer (p) be the key component in an aqueous mixture whose function data ( ) is
to be generated. That component has mole (or mass) fraction wp and pure-component
viscosity p. It is the only key component in the mixture of current interest, so:

=1

=1

+ = 1

where:

( ) =
=1

=1

=1

( ) + = 1 (77)

= 1

(78)

=1

Therefore, the normalizing factor N is:


=

1
1

(79)

54 Process Manual - Polymer Flood

Revision 0I

Substitute the expression for N, into the mixing rule equation and solve for ( ):

1
= +
( )
1

Now solving for ( ):


=

where:

( )

(80)

=1

(81)

1
=
( ) (82)
1
=1

This calculation is done for each of the 11 (xlow, xhigh and 9 intermediate) values specified
by *VSMIXENDP, and the result is entered for the 11 values f1 ... f11 required by
*VSMIXFUNC. This entire process must be done for each different key component
comp_name specified by *VSMIXCOMP.
Keywords *VSMIXENDP and *VSMIXFUNC cause function ( ) to be continuous and
piecewise linear in the domain [0,1] and range [0,1]. Beyond these constraints, any
function can be matched at the 11 points. If = , the nonlinear mixing option
reduces to the linear mixing rule.
Example:

Consider the following table in which the aqueous phase viscosity (aq) is a function of
polymer concentration (Cp), and the linear mixing rule is less appropriate for modeling the
changes of the liquid viscosity. Therefore, the key component in the mixing rule of the
viscosity will be the polymer and this needs to be specified by the *VSMIXCOMP keyword.
Cp (%)

aq (cp)

0.00

1.000

0.03

1.417

0.05

1.875

0.10

4.000

In the aqueous phase, only the water and polymer components will be present. It can be
inferred from the above table that the water component viscosity (w) is equal to 1 cp and the
polymer viscosity, as pure component, is 4 cp.

Revision 0I

Process Manual - Polymer Flood 55

The nonlinear viscosity function will depend upon mole (mass) fraction of polymer
component. The next step involves converting the weight percent of polymer to mass fraction
= /100 then determining the nonlinear viscosity, as shown in the following table:
wp (mass frac)

aq (cp)

0.0000

1.000

0.0003

1.417

0.0005

1.875

0.0010

4.000

Assuming the *MASSBASIS keyword is activated (which means the component property
data is based on mass instead of mole), this data can be plotted directly to obtain the equation
that represent the aqueous viscosity as a polymer concentration function:

The aqueous viscosity equation for this example will be:


= 1 + 1 103 + 1 106 2 + 1 109 3

If the polymer injection concentration is assumed to be 1000 ppm (0.1 %wt) then the values
of xlow and xhigh will be 0 and 0.1 respectively. The nine intermediate values will be calculated
internally in STARS using the following expression:
+1 = + 10

where 1 = 11 = .

( = 2 10) (83)

The corresponding viscosity values for the nine intermediate points of mass fraction are
obtained from the above aqueous viscosity equation:
j

wp (mass frac)

aq (cp)

0.0000

1.000

0.0001

1.111

56 Process Manual - Polymer Flood

Revision 0I

wp (mass frac)

aq (cp)

0.0002

1.248

0.0003

1.417

0.0004

1.624

0.0005

1.875

0.0006

2.176

0.0007

2.533

0.0008

2.952

10

0.0009

3.439

11

0.0010

4.000

The last step is to determine the values of the nonlinear function for j = 2, as shown below:
=

1
=
( ) (84)
1
=1

where water is the common component and the polymer is the key component. Further:
1
=
(0.9999) (1) = 0
1 0.001

(1.111)
(85)
(4)

Proceeding in the same way, the remaining cells in the table can be calculated:
wp (mass frac)

aq (cp)

f(wp)

0.0000

1.000

0.0000

0.0001

1.111

0.0759

0.0002

1.248

0.1598

0.0003

1.417

0.2514

0.0004

1.624

0.3498

0.0005

1.875

0.4534

0.0006

2.176

0.5608

0.0007

2.533

0.6704

0.0008

2.952

0.7808

0.0009

3.439

0.8910

0.0010

4.000

1.0000

The keywords for representing the nonlinear mixing rule should appear as follows:
*VSMIXCOMP 'Polymer'
*VSMIXENDP 0 0.001
*VSMIXFUNC 0.0 0.0759 0.1598 0.2514 0.3498 0.4534 0.5608 0.6704 0.7808 0.891 1.0

Revision 0I

Process Manual - Polymer Flood 57

Shear Effect on Polymer Viscosity

The following figure shows an example of polymer viscosity at different shear rates. The data
for generating this figure will be used to illustrate the Darcy velocity calculation using the
study of Cannella et al.31, mentioned earlier:

Figure 16 Polymer Viscosity at Different Shear Rates

For this example, assume the following coreflood experimental and rheological data:
kabs
krw
Sw

= 1 Darcy
= 0.15
= 0.78
= 0.35

Table 2 Polymer Viscosity vs. Shear Rate

SR (1/sec)

Aqueous Phase Viscosity (cps)


0.05 wt% Polymer

0.1 wt% Polymer

0.01

1.875

4.000

0.1

1.836

3.865

1.507

2.737

10

1.253

1.868

100

1.126

1.432

1000

1.063

1.215

The first step is to use the expression in Cannella et al.31 to convert shear rate to Darcy
velocity:
=

0.01 1 0.15 0.35 0.78


=
= 4.18 108 /

4.8 10066

58 Process Manual - Polymer Flood

Revision 0I

where 4.8 is the default shear rate factor and 10066 is a conversion factor to obtain cm/sec as
the units of velocity.
Proceeding in the same way, the following table can be generated:
SR (1/sec)

Darcy velocity (cm/sec)

0.01

4.188E-08

0.1

4.188E-07

4.188E-06

10

4.188E-05

100

4.188E-04

1000

4.188E-03

The shearing applies to the component and not to the phase, so the second step is to calculate
the pure component viscosity by using the non-linear mixing rule expression. For this
example the data used in Generating Data for the Polymer Case in the Aqueous Phase, will be
used. Assuming a polymer injection concentration of 500 ppm ( = 0.0005), the function
will be equal to 0.4534 and the aqueous phase viscosity, 1.875 cps, based on which
the polymer viscosity as pure component will be:
=

1
( )
1
(86)

1 0.4534
0.9995 (1)
(1.875)
1 0.0005
=
= 4
0.4534

Proceeding in the same way, all the polymer viscosity values as pure component can be
calculated and the table of *SHEARTAB will appear as follows:
*SHEARTAB
**Darcy Velocity (cm/sec)
4.188E-08
4.188E-07
4.188E-06
4.188E-05
4.188E-04
4.188E-03

Polymer Viscosity (cps)


4.0000
3.8172
2.4695
1.6455
1.2998
1.1441

If = 1 at the injection concentration, then the pure polymer viscosity will equal the
aqueous phase viscosity. For polymer concentrations at the grid blocks in the simulation
which are lower than the injection concentration, SHEARTAB is scaled down on the basis of
log (viscosity) versus log (velocity).

Revision 0I

Process Manual - Polymer Flood 59

In brief, STARS performs the following calculations:


1. STARS calculates the Darcy velocity for each grid cell. From the example, take
the value 4.188E-06 cm/sec.
2. STARS looks up the shear table with the calculated velocity to obtain the pure
component viscosity (polymer); in our example, 2.4695 cps.
3. STARS substitutes the new viscosity in the *AVISC/*BVISC keywords table.
4. STARS calculates the phase viscosity using the polymer mole fraction in each grid
cell, the polymer viscosity from the previous step, and the nonlinear mixing rule.
Assuming a polymer mole fraction of 0.0005, the viscosity of the aqueous phase or
polymer solution will be:
= +

1
( )
1

(87)

1 0.4534
= 0.4534 (2.4695) +
(1) (0.9995)
1 0.0005
= 0.40988
= 1.507

This last value means that the equations reproduce the viscosity values of the
polymer solution reported from the laboratory at different velocities or shear rates.
Salinity Effect on Polymer Viscosity

An example of the effect of salt concentration on polymer solution viscosity is shown in


Figure 17. The values used to build this plot are shown in Table 3.

Figure 17 Salinity Effect on Polymer Solution Viscosity

60 Process Manual - Polymer Flood

Revision 0I

Table 3 Salinity Effect on Polymer Solution Viscosity

Polymer Solution Viscosity, cps @

Salt Concentration
wt%

0.05 wt%

0.1 wt%

0.15 wt%

1.875

8.125

1.125

2.3

5.6875

10

0.75

1.6

4.0625

The keyword format requested by STARS to account for the salinity effect on the polymer
viscosity is as follows:
*VSSALTCMP _

where _ is the component (e.g. salt or alkaline) affecting the viscosity of the
nonlinear mixing component (e.g. polymer). This must be one of the component names
specified by *COMPNAME. is the minimum salinity of the phase and is the slope of
the log-log plot of polymer component viscosity versus ratio of salinity over minimum
salinity ( ).

To build the log-log plot, first calculate the polymer viscosity as a pure-component using the
non-linear mixing rule for phase viscosity, as outlined in Generating Data for the Polymer
Case in the Aqueous Phase:
*VSMIXCOMP 'Polymer'
*VSMIXENDP 0 0.001

*VSMIXFUNC 0.0 0.0759 0.1598 0.2514 0.3498 0.4534 0.5608 0.6704 0.7808 0.891 1.0
With a water component viscosity of 1 cps and a polymer component viscosity of 4 cps at
1 wt% Polymer, *AVISC = 4, and *BVISC = 0. In the above non-linear mixing rule, the rule
of mixture was made based on a maximum concentration of 1 wt% polymer with f = 1
at this maximum concentration. At this point, the polymer viscosity as pure component is
equal to that of the phase. For this example, however, we will assume the injection
concentration is 500 ppm (0.05 wt%). At this concentration, = 0.4534.
To illustrate, calculate the pure component viscosity at 0.05 wt% polymer concentration
using the following formula and the values of polymer solution or aqueous viscosity from
Table 3:

1
( )
1
(88)

The viscosity of the water component is 1 cp, so the above equation can be reduced to:
=

Revision 0I


(1.125)
=
= 1.297 (89)
0.4534

Process Manual - Polymer Flood 61

Repeat the calculation for the remaining polymer viscosity values as pure component, as
shown in Table 4.
Table 4 Polymer Solution and Component Viscosities vs. Salt Concentration

Salt Conc. (wt%)

Viscosity @ 0.05 wt% (cps)


Solution

Component

1.875

4.000

1.125

1.297

10

0.75

0.530

The second step is to calculate the ratio of the salinities ( ). For this, assume that
the brine salinity (or minimum salinity) is 30,000 ppm (3 wt%). The results are shown below:
(xsalt /xmin)

Component Viscosity (cps),


@ 0.05 wt%

4.000

1.297

3.33

0.530

The third step consists of plotting the viscosity of the component versus the salinity ratio in a
log-log graphic and calculating the slope of this curve, as shown in Figure 18:

Figure 18 Component Viscosity vs. Salinity Ratio

From Figure 18, = 1.6754 and = 0.03 in mass fraction. The keywords can be
added to STARS as follows:
*VSSALTCMP 'Salt' 0.03 -1.6754

62 Process Manual - Polymer Flood

Revision 0I

Using this information and the STARS power-law equation, the most updated viscosity (final
or resultant) of polymer (as pure component) in the saline solution will be obtained and used
with the nonlinear mixing function to determine the phase viscosities. To verify this, assume
the salinity in the system increases to 6 wt%. Since *AVISC = 4, 0 = 4 . Therefore:
= 0


0.06 1.6754
= 4

= 1.2523 (90)

0.03

and the aqueous phase viscosity is:

= = 0.4534 (1.2523) = 1.107 (91)

This value of 1.107 cps of the phase viscosity at 6 wt% of salt is comparable with that in
Figure 18, 1.125 cps obtained for example from laboratory. Therefore, the equations and the
calculations made can reproduce the effect observed in laboratory.
If both effects (shear rate and salinity) are considered in the simulation, then the polymer
viscosity as pure component is updated first by the velocity (or shear rate), then used in the
equation of salinity as 0 , to be updated by the salt concentration, and finally is used in the
non-linear mixing rule to calculate the phase viscosity.

Revision 0I

Process Manual - Polymer Flood 63

Appendix B Conversion of Eclipse


Polymer Option to IMEX
Example of Eclipse Data
--polymer solution viscosity function
--column 1: polymer concentration in solution kg/m = ppm/1000
--column 2: corresponding factor by which water viscosity in PVTW has to be multiplied
PLYVISC
0.0000 1.00000
0.5000 3.3333
1.0000 6.0
1.125 6.666
1.250 8 /
PLYSHEAR
--shear thickening
0.0 1.000
0.05 1.001
0.5 1.300
0.7 1.380
0.9 1.400
1.5 1.420
4.0 1.500 /
PLYROCK
--column 1: inaccessible PV, should change with rock type (more IPV for bad rock)
--column 2: RRF = decrease in krw when max amount of poly has been adsorbed
--column 3: mass density of rock [kg/m]
--column 4: adsorption index 2= no desorption
--column 5: max polymer adsorption [kg/kg] for max. RRF
0.07 1.5 1.9 2 5.0E-5 /
--polymer adsorption function, units in Eclipse: column 1: g/cm column 2: g/g
PLYADS
--column 1: concentration of polymer in solution around the rock [kg/m]
--column 2: corresponding saturated concentration of polymer adsorbed by the rock [kg/kg]
0.000 0.0000000
0.500 0.0000250
1.000 0.0000500
1.500 0.0000750 /

64 Process Manual - Polymer Flood

Revision 0I

PLYVISC and PLYSHEAR Conversion to IMEX


Eclipse PLYVISC converts to IMEX *PMIX keyword plus the use of *PREFCONC. If the
Eclipse keyword PLYSHEAR is present, then the subkeyword for *PMIX should be
*VELTABLE, otherwise if PLYSHEAR is not present the subkeyword for *PMIX should be
*TABLE.
Polymer Concentration Values
The IMEX keyword *PREFCONC should be set to the maximum value of the polymer
concentration given in the Eclipse keyword PLYVISC, and all concentration values in the
table for PLYVISC should be divided by this maximum value.
Viscosity Multipliers
The viscosity multipliers given in the second column in the table for PLYVISC can be used
without any changes for the second column of values for *PMIX in IMEX. If PLYSHEAR is
present, then a table must be generated for each velocity under this keyword. The table will
consist of the same values of polymer concentration and viscosity multipliers that were
converted from PLYVISC, except the second column of values will be multiplied by the
values corresponding to the velocity in the table.
The converted keywords PLYVISC and PLYSHEAR for the example on the first page are as
follows:
PREFCONC 1.25 **kg/m3
PMIX VELTABLE
VWT 0.0 **velocity=0.0 m/day
0.0000 1.00000
0.4000 3.3333
0.8000 6.0000
0.9000 6.6660
1.0000 8.0000
VWT 0.05 **velocity=0.05 m/day.
0.00 1.0000
0.40 3.3366
0.80 6.0060
0.90 6.6727
1.00 8.0080
VWT 0.5 **velocity=0.5 m/day
0.00 1.0000
0.40 4.3333
0.80 7.8000
0.90 8.6658
1.00 10.4000
VWT 0.7 **velocity=0.7 m/day
0.00 1.0000
0.40 4.6000
0.80 8.2800

Revision 0I

Process Manual - Polymer Flood 65

0.90 9.1991
1.00 11.0400
VWT 0.9 **velocity=0.9 m/day.
0.00 1.0000
0.40 4.6666
0.80 8.4000
0.90 9.3324
1.00 11.2000
VWT 1.5 **velocity=1.5 m/day.
0.00 1.0000
0.40 4.7333
0.80 8.5200
0.90 9.4657
1.00 11.3600
VWT 4.0 **velocity=4.0 m/day.
0.00 1.0000
0.40 5.0000
0.80 9.0000
0.90 9.9990
1.00 12.0000

PLYROCK Conversion to IMEX

Eclipse keyword PLYROCK converts to the IMEX keyword *PPERM. The first entry on
PLYROCK converts to the fourth table column on *PPERM via the formula p_pore = 1.0
IPV. The 2nd entry on PLYROCK is used directly for entry as the 5th column on *PPERM
(RRF). The 3rd entry on PLYROCK is used for calculation of IMEX adsorption. The 4th entry
on PLYROCK is used to indicate whether the 3rd column on *PPERM (res_ad) is set equal to
the 2nd column (i.e. for no desorption, set res_ad = max_ad, for desorption set res_ad=0). The
5th entry on PLYROCK is used directly for entry in the 2nd column of *PPERM (max_ad).

PLYADS Conversion to IMEX

Eclipse keyword PLYADS converts to the IMEX keyword *PADSORP. The first column on
PLYADS is used directly for the 1st column in *PADSORP. The 2nd column on PLYADS
converts to the 2nd column in *PADSORP via the formula: PADSORP(2nd col) =
PLYADS(2nd col) * rock_density * (1.0 porosity)/porosity. Rock_density is input as the 3rd
parameter in PLYROCK.

66 Process Manual - Polymer Flood

Revision 0I

The converted keywords PLYROCK and PLYADS for the example shown on page 64 are as
follows:
PPERM
** perm max_ad res_ad p_pore rrf
10.0 0.530 0.530 0.93 1.5
10000.0 0.530 0.530 0.93 1.5
PADSORP
**col 1: Polymer concentration kg/m3
**col 2: adsorption level (kg/m3) = kg_polymer/kg_rock*rock_density*(1-porosity)/porosity
**
use: rock_density=sandstone=2650.0 kg/m3 and porosity = 0.2. Factor=10600.
** p_con adsorp_level
0.000 0.000
0.500 0.265 ** =0.0000250 * 10600
1.000 0.530 ** =0.0000500 * 10600
1.500 0.795 ** =0.0000750 * 10600

Revision 0I

Process Manual - Polymer Flood 67

Appendix C Conversion of IMEX


Polymer Option to STARS
Step 1: Builder Convert Simulator Type
Open the IMEX data set in Builder and select the menu item File | Convert Simulator Type
For Data Set | To STARS. Enter a new data set name and select the option for fluid model
conversion Convert from Black oil model. Enter reservoir temperature and bubble point as
required to convert the PVT data into a STARS fluid model. This PVT conversion is
explained in more detail in the user manual for Builder (press F1 in Builder) under the
heading Using the STARS Import Black Oil PVT Wizard.

Step 2: Process Wizard


Open the STARS data set created in Step 1 in Builder and select the menu item
Components | Process Wizard. From the drop-down box, select Alkaline, surfactant, foam
and/or polymer model and then click Next. Choose model Polymer flood (add 1
component).

Input Polymer Data


Obtain the polymer resistance factor from column 5 of the IMEX keyword *PPERM and
enter this number in the wizard. Similarly, obtain the accessible pore volume from column 4
of *PPERM and enter this number in the wizard. Cancel Polymer quantity decreases with
time and enter the rock density appropriate for the reservoir to be modeled (if this file was
converted from Eclipse, then enter the rock density input from the 3rd parameter on Eclipse
keyword PLYROCK). Now click Next and Next again on the Component Selection dialog
box.

Input Adsorption Data


For adsorption, enter the porosity of the laboratory sample used for adsorption measurements
or if this file was converted from Eclipse, enter the porosity used for conversion of adsorption
from Eclipse (PLYADS) to IMEX (PADSORP). Locate the adsorption table in the IMEX
data file (*PADSORP) and determine the number of rows used for the table. In the process
wizard, right click on the adsorption table and insert rows until the number of rows in the
table equals the number of rows in *PADSORP. Calculate the 1st column for the wizard
(poly_wt%) from the 1st column of *PADSORP (p_con) with:
poly_wt% = p_con * 100 / water_density

68 Process Manual - Polymer Flood

Revision 0I

Calculate the 2nd column for the wizard (poly_ads, mg/100 gm rock) from the 2nd column of
*PADSORP (ads_level) with the equation below, then click Next.
poly_ads = ads_level / rock_density * porosity/(1porosity) * 1.0e5

Input Polymer Viscosity


Locate the polymer viscosity data in the IMEX data file under the keyword *PMIX.
Determine the number of rows of data used on this keyword (first table at velocity=0 if PMIX
VELTABLE is used). Make the number of rows on the process wizard equal to this number
of rows by right clicking on the table of viscosity vs. weight % polymer. Calculate the 1st
column of data for the wizard (poly_weight%) from the 1st column of PMIX (poly_conc,
kg/m3) and the value input in IMEX on keyword *PREFCONC using:
Poly_weight% = poly_conc * prefconc * 100 / water_density
Calculate the 2nd column of data for the wizard (poly_visc) from the 2nd column of PMIX
(visc_mult) using the equation below (note that the water viscosity with no polymer is shown
in the wizard on row 1 of the viscosity table). Click Finish when done with this table and
save the data set.
Poly_visc = visc_mult * water_viscosity(no polymer)

Step 3: Shear Dependent Viscosity


If the IMEX keyword *PMIX *VELTABLE is used, then the shear dependent viscosity
option must be used in STARS. With STARS versions 2012.10 or earlier, *VISCTABLE
cannot be used with SHEARTAB, so with these versions of STARS the *VISCTABLE data
must be converted into the *AVISC and *BVISC format. This can be done easily for
isothermal simulations by locating the *VISCTABLE entries at reservoir temperature and
entering these values on *AVISC, while setting all *BVISC values to zero.
To enable the input of velocity vs. viscosity (instead of shear rate vs. viscosity), open the
STARS data file in a text editor and add the keyword *SHEAREFFEC *SHV in the I/O
section of the data set (i.e. at the top).
Plot the values of polymer viscosity vs. velocity at different concentrations of polymer from
the IMEX data on PMIX. The Eclipse example data given at the beginning of this section is
plotted as shown below.

Revision 0I

Process Manual - Polymer Flood 69

Viscosity vs. Velocity vs. Polymer Concentration


6.0

Viscosity, cp

5.0
4.0

Viscosity at 0 wt% Polymer


Viscosity at 0.05 wt% Polymer
Viscosity at 0.1 wt% Polymer
Viscosity at 0.1125 wt% Polymer
Viscosity at 0.125 wt% Polymer

3.0
2.0
1.0
0.0
0

0.5

1.5

2.5

3.5

Velocity, m/day

From this plot, select the polymer concentration line that most represents the polymer versus
velocity behavior at the polymer concentrations and velocities expected to be dominant in the
simulation. This selection needs to be done because STARS will calculate the velocity
dependent viscosity at different polymer concentrations using the viscosity mixing rule (see
the discussion in the Shear-Dependent Viscosity Effects section of this manual). For the
example plot of viscosity versus velocity above, the STARS calculations of velocity
dependent viscosities will be as shown in the following figure. Some small differences can be
noted.
STARS Calculation of Viscosity vs. Velocity vs. Polymer Concentration
6.0

Viscosity, cp

5.0
4.0

Viscosity at 0.000 wt% Polymer


Viscosity at 0.050 wt% Polymer
Viscosity at 0.100 wt% Polymer
Viscosity at 0.1125 wt% Polymer
Viscosity at 0.1250 wt% Polymer

3.0
2.0
1.0
0.0
0

0.5

1.5

2.5

3.5

Velocity, m/day

70 Process Manual - Polymer Flood

Revision 0I

Using a text editor with the STARS data file, insert the *SHEARTAB keyword after the
*VSMIXFUNC keyword corresponding to Polymer. Then, with the table of velocity vs.
viscosity selected above, enter the velocity values directly into the first column in the table
following *SHEARTAB. Locate the value corresponding to the component Polymer on the
STARS keyword *AVISC (avisc_poly). Use this value together with the viscosity values
(visc_selected) from the table of velocity versus viscosity selected above to calculate the
values for the 2nd column of data following *SHEARTAB using the following formula:
SHEARTAB(2nd col) = visc_selected * avisc_poly / visc_selected(first_row)
The equivalent STARS keywords that were converted from the example Eclipse data are as
follows:
MODEL 4 4 4 2
COMPNAME 'Water' 'Polymer' 'Dead_Oil' 'Soln_Gas'
CMM
0
8
0.281591 0.017346
.
.
AVISC
0.4837 4.64398
52.2
1.26719
BVISC
0.0
0.0
0.0
0.0
VSMIXCOMP 'Polymer'
VSMIXENDP 0 3.10137e-006
VSMIXFUNC 0 0.1465 0.2930 0.43951 0.55849 0.62995 0.7014 0.77286 0.82949 0.91137 1
SHEARTAB
0.05
4.64398
0.5
6.031142857
0.7
6.40229011
0.9
6.495076923
1.5
6.587863736
4
6.959010989

Revision 0I

Process Manual - Polymer Flood 71

Appendix D Example IMEX and


STARS Simulations of Polymer
Comparisons of IMEX and STARS

To make sure that IMEX and STARS are predicting the same for both a normal waterflood
and a polymer flood with no velocity dependence, simulation runs should be done to compare
the results. A sample comparison, shown below, compares oil and water rates and well block
pressure.
10.0
Oil Rate SC (m3/day)

IMEX_BaseWF.irf
IMEX_Polymer_NoVelocityDep.irf
Stars_BaseWF.irf
Stars_Polymer_NoVelocity.irf

1.0

0.1

0.0

2012

2014

2016
Time (Date)

2018

2020

2022

Water Rate SC (m3/day)

2.00
IMEX_BaseWF.irf
IMEX_Polymer_NoVelocityDep.irf
Stars_BaseWF.irf
Stars_Polymer_NoVelocity.irf

1.50
1.00
0.50
0.00

2012

2014

2016
Time (Date)

2018

2020

2022

Well Block Pressure (kPa)

13,500
13,000
12,500
12,000
IMEX_BaseWF.irf
IMEX_Polymer_NoVelocityDep.irf
Stars_BaseWF.irf
Stars_Polymer_NoVelocity.irf

11,500
11,000
10,500

2012

2014

72 Process Manual - Polymer Flood

2016
Time (Date)

2018

2020

2022

Revision 0I

STARS Options For Real Polymer Phenomena


Simulations using STARS polymer injection options to determine the effects of various real
phenonena are given in the following plot. As shown, the effects of using polymer
degradation [see section Polymer Degradation] results in a lower amount of incremental oil.
Adding wettability effects (see section Relative Permeability/Wettability Alteration Effects)
that cause the rock to shift towards water wet results in additional incremental oil.

Cumulative Oil SC (m3)

800

600

400

Stars_BaseWF.irf
Stars_Polymer.irf
Stars_Polymer_Degradation.irf
Stars_Polymer_Degradation_Wettability.irf

200

Revision 0I

2012

2014

2016
Time (Date)

2018

2020

2022

Process Manual - Polymer Flood 73

Vous aimerez peut-être aussi