new insights into polymer retention in …baervan.nmt.edu/publications/media/pdf/thesis/zhang...

123
NEW INSIGHTS INTO POLYMER RETENTION IN POROUS MEDIA By GUOYIN ZHANG A DISSERTATION SUBMITTED TO THE DEPARTMENT OF PETROLEUM ENGINEERING IN PARTIAL FULFILLMENT OF THE REQUIREMENTS FOR THE DEGREE OF DOCTOR OF PHILOSOPHY NEW MEXICO INSTITUTE OF MINING AND TECHNOLOGY DECEMBER, 2013

Upload: vuanh

Post on 04-Jun-2018

216 views

Category:

Documents


0 download

TRANSCRIPT

NEW INSIGHTS INTO POLYMER RETENTION

IN POROUS MEDIA

By

GUOYIN ZHANG

A DISSERTATION

SUBMITTED TO THE DEPARTMENT OF PETROLEUM ENGINEERING

IN PARTIAL FULFILLMENT OF THE REQUIREMENTS

FOR THE DEGREE OF DOCTOR OF PHILOSOPHY

NEW MEXICO INSTITUTE OF MINING AND TECHNOLOGY

DECEMBER, 2013

DEDICATION

Dedicated to my Mom, Huatang Xu, and my Dad, Xingshan Zhang.

ABSTRACT

When water-soluble, high molecular weight polymers are used to reduce water mobility

and improve volumetric sweep efficiency in enhanced oil recovery technology, polymer

retention occurs when propagating through the reservoir. It is widely accepted that

polymer retention comprises the adsorption on the rock surface, mechanical entrapment

in small pores, and flow-induced hydrodynamic retention. Retention leads to polymer

loss in reservoir and delay oil bank during a polymer flooding. Therefore, proper

characterization of polymer retention in porous media is critical for a polymer flooding

project.

In this study, we first investigated the effect of concentration on HPAM retention

because the literature showed controversial results on this issue. To accomplish this, both

static adsorption on disaggregated sands and dynamic retention in sandpacks and

sandstone cores were measured. Different retention behaviors were observed in dilute,

semidilute and concentrated regions. In both dilute and concentrated regions, polymer

retention is basically concentration-independent. In contrast, in the semidilute region,

polymer retention shows concentration-dependent behavior. Little re-adsorption occurs at

high concentration after the adsorbent’s pre-contact with low concentration polymer. The

results also show macromolecule polymers have high adsorption tendency on rock

surface and the adsorption can be considered almost instantaneous and irreversible.

Based on experimental results, a concentration-related retention mechanism is

proposed which correlates the orientation of adsorbed polymer molecules on rock surface

with the interaction between molecular coils in solution.

Next, hydrodynamic retention caused by the increase of hydrodynamic force acting

upon polymer molecules was evaluated. As flow rate rises from 3.26 ft/day (base

reference) to 6.52 ft/day and 52.13 ft/day, retention of 500 ppm HPAM in 1.9 Darcy

sandstone core increases by 13.2% and 39.16%, respectively. Hydrodynamic retention

shows strong flow dependence. In low flow region, the retention increases abruptly with

increased flow rate. By comparison, in high flow region, the increase becomes much

more gradual. Our results also demonstrate that this flow-induced retention is totally

reversible (no incremental irreversible retention) and residual resistance factor is not

affected by this reversible retention.

To prove polymer rheology in porous is an intrinsic property, not caused by retention

related permeability reduction, xanthan polymer was also tested. The results indicate with

increase of flow rate, both HPAM and xanthan retention goes up in a 71 md sandstone

core. However, the resistance factor measurements show HPAM and xanthan show

distinct rheology in porous media. At low flow rate, even with retention increase, HPAM

shows Newtonian fluid behavior and no resistance factor increase is observed. In contrast,

xanthan polymer exhibits shear thinning behavior. Therefore, the hydrodynamic retention

has limited effect on polymer rheology in porous media.

Negative polymer inaccessible pore volume (IAPV) is observed with increase of flow

rate and decrease of permeability. This presence of negative IAPV is caused by reversible

polymer retention. Adsorptive retention on rock surface proves to be almost irreversible,

therefore, the mechanical entrapment and hydrodynamic retention should account for this

phenomenon.

Keywords: Adsorption, mechanical entrapment, hydrodynamic retention, reversibility,

inaccessible pore volume, core flooding

ii

ACKNOWLEDGEMENTS

There are many people who helped me during my study at New Mexico Institute of

Mining and Technology. This work would have never been done without them.

First, I would like to thank my advisor, Dr. Randall Seright for his persistent guidance,

encouragement, patience and tolerance. His extraordinary knowledge, wisdom in science

and engineering and attitude toward pursuit of truth always impress and inspire me.

Next, I would like to thank my committee members, Dr. Thomas Engler, Dr. Mike Kelly

and Dr. Reid Grigg for reading my dissertation and providing valuable comments and

suggestions.

I want to thank Kathryn Wavrik for her tremendous help in the lab. She taught me how to

use rheometer, TOC analyzer, prepared cores and setup equipment for me. I learned a lot

by working with her.

I would also thank Dr. Jianjia Yu and James Mclemore. They offered me generous help

whenever I got into trouble with my experiment. Thanks also go to Dr. Robert Lee, Dr.

Ning Liu, Elizabeth Bustamante, Xu Han, and other PRRC staff.

Finally, special thanks go to my wife, Ms. Dongling Xu and our daughter Yifan Zhang,

who have been creating a loving and supportive family environment for me.

iii

TABLE OF CONTENTS

DEDICATION

ABSTRACT

ACKNOWLEDGEMENTS ............................................................................................... ii

TABLE OF CONTENTS ................................................................................................. iii

LIST OF FIGURES ......................................................................................................... v

LIST OF TABLES ......................................................................................................... viii

CHAPTER 1. INTRODUCTION ....................................................................................... 1

1.1 Significance of Polymer Retention during Polymer Flooding ................................... 1

1.2 Statement of Problem ............................................................................................... 4

1.3 Approach ................................................................................................................... 7

1.3.1 Batch Adsorption or Static Retention Measurement. ......................................... 8

1.3.2 Flow Experiment or Dynamic Retention Measurement. .................................... 8

1.4 Outlines of Dissertation ........................................................................................... 10

CHAPTER 2. LITERATURE REVIEW .............................................................................. 11

2.1 Brief Introduction to Polymer Flooding .................................................................. 11

2.2 Overview of Polymer Retention Mechanisms in Porous Media ............................. 13

2.2.1Mechanisms of Polymer Retention in Porous Media ........................................ 14

2.2.2 Polymer Inaccessible Pore Volume (IAPV) ...................................................... 16

2.2.3 Factors Influencing Polymer Retention in Porous Media ................................ 19

2.3 Langmuir Adsorption Isotherm ............................................................................... 29

2.4 Concluding Remarks ................................................................................................ 31

CHAPTER 3. METHODS AND PROCEDURES ................................................................ 33

3.1 Introduction ............................................................................................................. 33

3.2 Equipment and Material ......................................................................................... 33

iv

3.3 Experimental Procedures ........................................................................................ 39

3.4 Polymer Injection at Different Concentrations. ...................................................... 45

3.5 Polymer Injection at Different Flow Rates. ............................................................. 46

CHAPTER 4. RESULTS AND DISCUSSIONS ................................................................... 47

4.1 Introduction ............................................................................................................. 47

4.2 Dependence of Retention on HPAM Concentration ............................................... 48

4.2.1 Static Measurements ......................................................................................... 48

4.2.2 Dynamic Measurements ................................................................................... 53

4.2.3 Proposed Adsorption Model ............................................................................. 62

4.2.4 Overlap Concentration (C* and C**) Measurement ........................................ 66

4.3 Effect of Flow Rate on Polymer Retention .............................................................. 69

4.3.1 Method Established to Detect Hydrodynamic Retention ................................. 70

4.3.2 Hydrodynamic Retention in 1.9 Darcy Dundee Sandstone Core ..................... 73

4.3.3 Is HPAM Shear Thickening Behavior Caused by Hydrodynamic Retention? . 80

4.3.4 Hydrodynamic Retention in 71 mD Berea Sandstone Core ............................. 83

4.4 Polymer Inaccessible Pore Volume (IAPV) .............................................................. 88

4.5 Effect of Polymer Retention on Permeability Reduction ........................................ 93

4.6 Steady-State Flow in Porous Media ........................................................................ 95

CHAPTER 5. CONCLUSIONS ....................................................................................... 98

5.1 Conclusions .............................................................................................................. 98

5.2 Discussions and Future Work ................................................................................ 100

NOMENCLATURE .................................................................................................... 103

REFERENCES ........................................................................................................... 106

v

LIST OF FIGURES

Fig. 1. 1-Polymer bank delay factors associated with polymer retention. ......................... 3

Fig. 1. 2-Delay in oil recovery caused by retention............................................................. 4

Fig. 2. 1-Polymer retention mechanisms in porous media (Szabo and Corp, 1975). ....... 16

Fig. 2. 2-Effect of polymer retention and IAPV on polymer propagation. ........................ 18

Fig. 2. 3-Typical Langmuir adsorption isotherm. .............................................................. 30

Fig. 3. 1-Rheology of HPAM polymer in a viscometer. ..................................................... 35

Fig. 3. 2-Viscosity vs. concentration at shear rate of 7.3 s-1. ............................................ 35

Fig. 3. 3-Sand shaker (IKA KS 4000). .................................................................................. 36

Fig. 3. 4-Schematic diagram of polymer retention determination system. ..................... 39

Fig. 3. 5-Roller for static measurement. ........................................................................... 40

Fig. 3. 6-Total Organic Carbon (TOC) analyzer for concentration determination. ........... 41

Fig. 3. 7-Correlation between TOC and polymer concentration. ..................................... 42

Fig. 3. 8-Polymer retention and inaccessible pore volume (IAPV) determination. .......... 43

Fig. 3. 9- p vs. Cp when HPAM flowing through a 10 m filter. ...................................... 45

Fig. 4. 1-Kinetics of polymer adsorption on sand. ............................................................ 49

Fig. 4. 2-Desorption tests for 100-, 500-, and 1,000-ppm HPAM. .................................... 50

Fig. 4. 3-Adsorption isotherm of HPAM using static method. .......................................... 51

Fig. 4. 4-Comparsion of retention on fresh sands and used sands. ................................. 52

vi

Fig. 4. 5-Adsorption isotherm using dynamic method (fresh sandpacks used for each

case). ................................................................................................................................. 54

Fig. 4. 6-Retention determination for 50 ppm HPAM. ..................................................... 55

Fig. 4. 7-Retention determination for 500 ppm HPAM. ................................................... 55

Fig. 4. 8-Retention of 1,000 ppm in pre-treated sandpack with 500 ppm. ...................... 56

Fig. 4. 9-Effect of concentration on polymer retention, 347 mD core. ............................ 58

Fig. 4. 10-Retention determination for 80 ppm HPAM, 347 mD core. ............................. 58

Fig. 4. 11-Retention isotherm of HPAM in 71 mD core. ................................................... 59

Fig. 4. 12-Retention determination for 20 ppm HPAM. ................................................... 60

Fig. 4. 13-Retention determination for 40 ppm HPAM. ................................................... 60

Fig. 4. 14-Polymer molecule interaction at different concentrations. ............................. 64

Fig. 4. 15-Proposed polymer adsorption mechanism on the rock surface. ...................... 64

Fig. 4. 16-Overlap concentration (C*) determination by linearity deviation. .................. 68

Fig. 4. 17-Overlap concentration (C*) determination by intrinsic viscosity. .................... 68

Fig. 4. 18-Polymer retention in the near wellbore region, radial flow. ............................ 70

Fig. 4. 19-Mehod to determine hydrodynamic retention. ............................................... 72

Fig. 4. 20-Effect of flow rate on KI (tracer) retention. ...................................................... 73

Fig. 4. 21-Polymer retention at flow rates from 3.26 ft/day to 52.16 ft/day. .................. 74

Fig. 4. 22-Incremental retention of HPAM vs. flux. .......................................................... 75

Fig. 4. 23-Determination of irreversible retention at 13.04 ft/day. ................................. 77

Fig. 4. 24-Determination of irreversible retention at 52.16 ft/day. ................................. 77

Fig. 4. 25-Resistance factor of HPAM at different flow rates. .......................................... 79

Fig. 4. 26-Rheology of HPAM and xanthan polymers in a viscometer. ............................ 81

vii

Fig. 4. 27-Hydrodynamic retention of 150 ppm xanthan in 1.9 Darcy core. .................... 82

Fig. 4. 28-Resistance factor of xanthan at different flow rates. ....................................... 83

Fig. 4. 29-Hydrodynamic retention and resistance factor for HPAM. .............................. 86

Fig. 4. 30-Hydrodynamic retention and resistance factor for xanthan. ........................... 87

Fig. 4. 31-IAPV determination for 100 ppm HPAM, 60 ml/hr, 347 mD core. ................... 88

Fig. 4. 32-IAPV determination for 20 ppm at flow rate of 1,000 ml/hr. ........................... 90

Fig. 4. 33-IAPV determination for 20 ppm HPAM at flow rate of 60 ml/hr. ..................... 91

Fig. 4. 34-2nd injection of 160 ppm HPAM and tracer at 60 ml/hr, 71 mD core. ............ 92

Fig. 4. 35-2nd injection of 1,000 ppm HPAM and tracer at 240 ml/hr, 71 mD core. ....... 92

Fig. 4. 36-Effect of polymer concentration on residual resistance factor, 347 mD core. 94

Fig. 4. 37-Effect of polymer concentration on residual resistance factor, 71 mD core. .. 94

Fig. 4. 38-Pressure drop during polymer injection in 347 mD core. ................................. 96

Fig. 4. 39-Pressure drop during polymer injection in 71 mD core. ................................... 96

viii

LIST OF TABLES

Table 2.1-Summary of Polymer Retention Using Dynamic Measurement....................... 22

Table 3.1-Core Properties. ................................................................................................ 37

Table 4.1-Dynamic Retention in Sandpacks ..................................................................... 54

Table 4.2-Retention on Sandstone Cores ......................................................................... 61

Table 4.3-Summary of Adsorption vs. Polymer Concentration ........................................ 65

Table 4.4-Properties of the Dundee Sandstone Core. ...................................................... 74

Table 4.5-Retention Summary. ......................................................................................... 74

1

CHAPTER 1. INTRODUCTION

1.1 Significance of Polymer Retention during Polymer Flooding

When water-soluble, high molecular weight polymers are used for enhanced oil recovery

(EOR), polymer retention delays polymer propagation into the formation. The presence

of the polymer is needed to provide high viscosity and low mobility levels—which in

turn are needed to improve oil displacement and sweep efficiency. Consequently, high

polymer retention can substantially delay oil displacement and recovery. To illustrate this

point, consider the range of polymer retention levels reported in the literature—9 to 700

g/g (Green and Willhite 1998)—and the range of polymer concentrations used in

polymer floods—500 to 3,000 ppm. Given the rock density ( rock, 2.65 g/cm3 for quartz),

porosity ( e.g., 0.3), polymer retention in g/g (Rpret), and polymer concentration in

ppm (Cpoly), Eq. 1.1 can be used to calculate the delay (PVret, pore volume delay per pore

volume injected).

(1 ) / /ret rock pret polyPV R C …………………………………………………...(1.1)

Using this equation and the parameters mentioned above, Fig. 1.1 shows delay factors.

With a very low retention level of 10 g/g and a polymer concentration of 2,000 ppm, the

2

delay factor is only about 3% of one pore volume (PV). In contrast, for a high retention

of 500 g/g and a polymer concentration of 500 ppm, the delay factor is over 6 PV. For

more typical values of 150 g/g for retention and a polymer concentration of 1,500 ppm,

the delay factor is about 0.6. For this latter combination, a 20% difference in retention

would mean an extra 12% PV polymer bank needed (if the retention is higher) or not

needed (if the retention is lower) to accomplish a given objective. In one 40-acre 5-spot

pattern with a height of 20 ft and a porosity of 0.3, 0.12 PV of 1,500-ppm HPAM

(costing $1.5/lb) would represent a polymer cost of about $176,000.

From another viewpoint, the mass of rock in the above 40-acre pattern is

40*43560*(12*2.54)3*2.65*(1-0.3)/0.3=6.10 x10

12 grams. Given the retention levels of

10, 50, 120, 150, 180, and 500 g/g, and an HPAM cost of $1.5/lb, the polymer costs

required to satisfy the retention requirements of the rock would be $201,777, $1,008,883,

$2,421,320, $3,026,650, $3,631,980, and $10,088,835, respectively.

Of course, the delay in polymer propagation also delays oil recovery. Fig. 1.2

illustrates this point using fractional flow calculations (from

http://baervan.nmt.edu/randy/). For these calculations, we assumed oil viscosity of 1,000

cp, water viscosity of 1 cp, and the reservoir was initially at connate water saturation

(Swr=0.3). The reservoir was then flooded with one PV of water (before continuous

polymer flooding with 100-cp polymer), one homogeneous layer was present, flow was

linear, and the following relative permeability curves (Eq. 1.2 and 1.3) were used.

20.1 [( 0.3) / (1 0.3 0.3)]rw wk S ………………………………………….………(1.2)

3

21 [(1 0.3 ) / (1 0.3 0.3)]ro wk S ………………………………………….……...(1.3)

In Fig. 1.2, the term, IAPV, refers to inaccessible pore volume, which is defined as the

fraction of the pore space being inaccessible to the large polymer molecules but

accessible to the small solvent and salt molecules and ions. IAPV accelerates polymer

propagation, whereas polymer retention (PVret) retards it. Three different levels were

considered in Fig. 1.2—where retention plus IAPV (1) were perfectly balanced to cause

no delay in polymer propagation (i.e., PVret + IAPV=0), (2) caused a one PV delay (i.e.,

PVret + IAPV=-1), and (3) caused a 2.5 PV delay (i.e., PVret + IAPV=-2.5). Figure. 1.2

illustrates that the delay in the arrival of the oil bank is directly proportional to the delay

in polymer propagation. Consequently, high polymer retention is economically

detrimental because of increased cost for polymer and delayed oil recovery.

Fig. 1. 1-Polymer bank delay factors associated with polymer retention.

0.01

0.1

1

10

1 10 100 1000

Po

re v

olu

me

de

lay

fact

or

Polymer retention, µg/g

500 ppm

1000 ppm

1500 ppm

2000 ppm

Porosity=0.3

rock=2.65 g/cm3

IAPV=0

4

Fig. 1. 2-Delay in oil recovery caused by retention.

In this research, experimental studies are performed to yield some new insights into

polymer retention in porous media. The effects of concentration, injection rate, and core

permeability on polymer retention are investigated. The permeability reduction caused by

polymer retention as well as retention reversibility is also analyzed based on experimental

results.

1.2 Statement of Problem

When enhanced oil recovery (EOR) polymers propagate through reservoir matrix, they

tend to adsorb on the rock surface due to the affinity of polymer molecules for many

reservoir rocks. In addition to the adsorption, polymer molecules tend to be trapped and

accumulate in the small pores. The latter retention is commonly known as mechanical

entrapment. Another retention, which is called hydrodynamic retention, may occur when

flow rate suddenly increases. Retentions caused by different mechanisms show different

40

50

60

70

80

90

100

1 1.5 2 2.5 3 3.5 4 4.5 5

% o

f m

ob

ile

oil

rec

ove

red

PV injected

PVret + IAPV = 0 PVret + IAPV

= -1

PVret + IAPV = -2.5

krw = 0.1 [(Sw-0.3)/(1-0.3-0.3)]2

kro = 1 [(1-0.3-Sw)/(1-0.3-0.3)]2

1,000 cp oil, 1 cp water, 100 cp polymer

5

reversibility and permeability reduction behaviors. Many factors prove to influence

polymer retention in porous media. Briefly, they can be divided into three categories.

(1) Formation properties. These include rock permeability, mineralogy, clay content,

salinity and pH of formation brine, and rock wettability, as well as reservoir

temperature. Generally, polymer retention increases dramatically with decreasing

permeability for less permeable rock (below several hundred mD). If more clay is

present, retention tends to increase because of increased surface area. Compared

with permeability and clay content, other parameters may also affect polymer

retention, but generally, they show minor impact (Smith 1970; Hirasaki and Pope

1974; Vela et al 1976; Espinasse and Siffert; 1979; Shah et al. 1985; Huang and

Sobie 1993; Broseta et al. 1995).

(2) Polymer and solvent types. Due to their molecule orientation in the solution or the

distinct functional groups in the molecules, different types of polymers such as

xanthan polymers, polyacrylamides (PAM), and partially hydrolyzed

polyacrylamides (HPAM) show different retention behaviors under the same

conditions (Chiappa et al 1999; Sydansk and Romero-Zeron 2011). Adsorption of

polyacrylamide on silica sand decreases with increasing hydrolysis (Martin and

Sherwood 1975). A study conducted by Meiste et al (1980) confirms that the

increase of hydrolysis of polyacrylamide reduces retention on the negatively

charged surface. Research from He et al (1990) indicates that smaller polymer

molecules, instead of larger molecules tend to be preferentially adsorbed if a

polymer mixture was injected simultaneously, resulting in a higher weight-

averaged Mw for the early effluent than for the injected polymer. Studies from

6

both Koral et al (1957) and Stromberg et al (1959) show solvents also play an

important role in polymer adsorption. About two to four times as much polymer

was absorbed from poor solvent (high solubilization ability) as from the good

solvent (low solubilization ability).

(3) Besides these factors mentioned above, polymer flow rate should also be taken

into account when investigating polymer retention in porous media. The increase

of flow rate is accompanied with additional polymer retention in porous media.

Extensive work has been done to describe the retention behaviors of the EOR

polymers. Nevertheless, there are still some areas needed to be clarified. Among them,

the effect of polymer concentration on retention is an outstanding one. Dawson and Lantz

(1972) proposed that polymer retention in porous media follows the Langmuir isotherm

without justification. Most of these researchers who claimed that polymer retention in

porous media either fits the Langmuir isotherm or is strongly concentration-dependent

arrived at their conclusions based on static adsorption measurements (Mungan 1969;

Szabo and Corp 1975; Deng et al 2006). Zheng et al (1998) suggest their experimental

data obtained from dynamic method using the same core fits the Langmuir isotherm.

However, a careful examination of their data shows the highest retention at 1,500 ppm is

less than 1.5 times higher than the lowest value at 250 ppm and no retention data was

provided from concentrations below 250 ppm. Few researchers except Vela et al (1976),

Shah et al (1978), and Szabo and Corp (1975) have tried to measure retention through

dynamic measurement. Again, the data is very limited. Currently, the Langmuir

adsorption model which is well-known for describing reversible adsorption is used in

most chemical flooding simulators to describe polymer retention in porous media which

7

shows little reversibility (Satter et al. 1980, Vossoughi et al. 1984, Camilleri et al. 1987,

Yuan et al. 2010, Dang et al. 2011).

Next, we will focus on how the flow rate influences polymer retention in porous rock.

Previous studies (Maerker 1973; Dominguez and Willhite 1976; Aubert and Tirrell 1980;

Zaitoun and Kohler 1987; Huh et al 1990) reported that more polymer molecules would

be retained with injection rate increase. However, no specific amount of retention is

measured at increased flow rates. Chauveteau et al (1974) suggested that shear

thickening behavior that was widely reported for HPAM solutions in porous media was

caused by “bridging adsorption”. In our study, we will address this question: whether

polymer rheology in porous media is an intrinsic property, or strongly affected by this

flow-induced retention.

We will also investigate the reversibility of polymer retention under various

conditions and how polymer retention in porous media alters the rock permeability.

Basically, the following issues will be addressed in this study:

1) Does polymer retention in porous media depend on polymer concentration? Or,

does it follow the Langmuir isotherm?

2) How can quantify hydrodynamic retention be quantified for different rates?

3) Under what circumstances does polymer retention becomes more reversible?

4) Does hydrodynamic retention dominate polymer rheology in porous media?

5) How does polymer retention affect rock permeability?

1.3 Approach

Two approaches were devised to measure polymer retention in porous rocks. One is

called batch adsorption or static retention measurement and the other is called flow

8

experiment or dynamic retention measurement. They are chosen based on different

scenarios.

1.3.1 Batch Adsorption or Static Retention Measurement.

Batch adsorption is used to estimate polymer retention on disaggregated sand grains or in

unconsolidated rocks with relatively high permeability where the adsorptive retention

dominates. To determine the adsorption, sand grains with a particular size distribution are

prepared by grinding sandstone cores. Next, a polymer solution with known

concentration is contacted sufficiently with known mass of dry and fine sand grains. The

system containing both sand grains and polymer solution will be thoroughly mixed. After

the retention reaches equilibrium, the upper liquid phase is separated from the solids and

sands are removed by centrifuging. Polymer concentration is determined by total organic

carbon (TOC) analyzer. The amount of polymer adsorbed on the sand surface is

calculated by mass balance.

1.3.2 Flow Experiment or Dynamic Retention Measurement.

If consolidated porous media is used to determine polymer retention, the batch adsorption

method is no longer applicable. This is because extra surface area will be generated

during the fine grain preparation and the application of batch adsorption may introduce

significant error. As a result, dynamic retention measurement that involves the injection

of polymer solution through a porous media is used. Both polymer adsorption retention

and mechanic entrapment retention can be measured this way. To date, it is still a

challenging task to distinguish between these two types of retentions in porous media.

9

Several methods have been proposed to measure dynamic polymer retention in

porous media (API RP63 1990, Dawson and Lantz 1972, Szabo 1975, 1979, Dominguez

and Willhite 1977, Gupta and Trushenski 1978, Castagno et al. 1987, Huh et al. 1990,

Mezzomo et al. 2002). Several of them advocate injection of a slug of polymer solution,

followed by brine, and performance of a mass balance on the polymer (i.e., retention =

polymer injected minus polymer produced). Key problems with this type of method are:

(1) recovery of the polymer may require an extended period of brine injection because of

the unfavorable displacement and (2) cumulative errors associated with measurements of

low polymer concentrations in the produced fluid can introduce considerable uncertainty

to the mass balance.

We prefer the method used by Lotsch et al. (1985), Hughes et al. (1990), and

Osterloh and Law (1998). In this method, two banks of polymer solution are injected

which are separated by a brine slug. Polymer retention can be determined by the plot of

the two effluent polymer concentration profiles versus pore volume injected.

Hydrodynamic retention can also be measured this way by varying polymer injection rate.

Another important parameter affecting polymer propagation in porous media is

polymer inaccessible pore volume (IAPV). It is defined as the faction of pore space not

contacted by polymer molecules due to a smaller inlet diameter compared to the size of

polymer molecules. To estimate the inaccessible pore volume, KI tracer with

concentration of 40 ppm is injected together with polymer solution and its concentration

in the effluent is measured by an absorbance detector. The area between the second

polymer and tracer breakout curves is used to estimate polymer inaccessible pore volume.

10

During dynamic retention measurement, pressure drops across the core will be

recorded which are used to calculated resistance factor and residual resistance factor.

Again, polymer rheology in porous and its dependence on hydrodynamic retention will

also be addressed in our study.

1.4 Outlines of Dissertation

Chapter 2 is the literature review. It will briefly introduce the concept of polymer

flooding, polymer retention and the retention mechanisms in the porous media proposed

by the researchers. Most importantly, in Chapter 2, the effect of polymer concentration,

and flow rate on retention will be reviewed in detail.

Chapter 3 describes the experimental equipment setup and testing procedures. As

introduced earlier, polymer retention will be estimated using both static and dynamic

methods. The measurement of polymer and tracer concentration in the effluent is a key

part of this test. In Chapter 3, a new and convenient method to determine effluent

polymer concentration is established.

Chapter 4 deals with the experiment results and discussions. The effect of

concentration, flow rate on HPAM retention, polymer reversibility and permeability

reduction caused by retention will be included in this chapter. Besides these observations,

a concentration-related retention mechanism is proposed that considers the orientation of

the adsorbed polymer molecules and the interaction between molecular coils in solution.

Chapter 5 summarizes this work and recommends areas needs further study.

11

CHAPTER 2. LITERATURE REVIEW

In this chapter, a brief introduction to polymer flooding and the main role of polymer in

improving sweep efficiency is provided, followed by the concept and mechanisms of

polymer retention in porous media. Most importantly, the previous findings associated

with the effect of concentration and flow rate on polymer retention will be reviewed.

2.1 Brief Introduction to Polymer Flooding

Waterflooding is usually performed after the primary recovery during the development of

a typical oil reservoir. However, due to the low viscosity of water or brine, viscous

fingers form during water injection, resulting in an early breakthrough and poor sweep

efficiency. To mitigate this unfavorable situation, a water-soluble, high molecular weight

polymer is usually added to the water phase to increase its viscosity and thus reduce its

mobility. Both biopolymers (e.g., xanthan) and synthetic polymers such as partially

hydrolyzed polyacrylamide polymer (HPAM) have been tried. Currently, HPAM

polymers are most widely used in polymer flooding due to their low cost, vast

commercial availability, excellent viscosity-enhancing performance and resistance to

microbial degradation.

12

Mobility ratio, M, the ratio of the displacing phase to displaced phase mobility, is the

most important parameter for polymer flooding operation.

( ) / ( )rw roD

d w o

k kM ……………………………………………………………… (2.1)

where, D is mobility of the displacing phase (water) and d is mobility of the displaced

phase (crude oil). krw and kro are the relative permeability to water and oil, respectively.

w and o refer to the water viscosity and oil viscosity.

Based on the value of mobility ratio (M) relative to unity, the displacing process is

considered to be either favorable, where M ≤ 1, or unfavorable, where M > 1. To attain a

favorable mobility ratio (M) and to improve the sweep efficiency, increasing viscosity of

the water phase is the most common way used.

The most important mechanism of polymer flooding is its capability of improving

volumetric sweep efficiency and conformance control, which can be attributed to

viscosity-enhancing property of polymers. Polymer flooding is not expected to reduce

residual oil saturation lower than waterflooding because the addition of polymer into the

aqueous solution does not significantly change the interfacial tension between aqueous

phase and oil phase. However, some researchers (Mohammad et al. 1992, Wang et al.

2001, Mojdeh, et al. 2008, Zhang et al. 2010, Urbissinova et al. 2010, Wang 2010)

proposed that the viscoelasticity of polymer solution could improve the microscopic

sweep efficiency after extensive pore volumes of water injection. So far, this is still a

controversial subject.

13

Polymer flooding has been applied in the field on a substantial scale. For instance, in

Daqing oilfield, China, it presently contributes to about one quarter of the annual oil

production. There are 37 polymer flooding operations and about 9,000 wells involved as

of 2005 and over 10 percent original oil in place (OOIP) has been recovered by

conducting this technique (Liu et al. 2009). Early screening criteria indicated that

polymer flooding should be applied in reservoirs with oil viscosity between 10 and 150

cp (Taber et al. 1977a, 1977b). However, with extensive use of horizontal wells and

fracturing technology, polymer flooding also shows great potential for heavy oil recovery

(Seright 2010, Delamaide et al. 2013).

2.2 Overview of Polymer Retention Mechanisms in Porous Media

Among the factors influencing the performance of a polymer flooding, polymer retention

is recognized as very important. Suppose severe retention occurs in the reservoir but it is

not properly considered during the project design, it may cause polymer flooding to fail

technologically and economically.

In this section, three mechanisms on how polymer molecules tend to be retained in

the porous media are first reviewed. When dealing with polymer retention, another

parameter called inaccessible pore volume (IAPV) should not be avoided. IAPV describes

the fraction of pore space that cannot be contacted by the injected polymers due to the

small pore diameter relative to the polymer molecule size. The combination effect of

IAPV and retention on polymer propagation in reservoir will be demonstrated. Finally,

the review of effects of polymer concentration, flow rate, rock mineralogy on retention,

14

retention reversibility as well as the role of retained polymer molecules in altering

permeability will be conducted.

2.2.1Mechanisms of Polymer Retention in Porous Media

Polymer retention primarily comprises three mechanisms. The first mechanism is caused

by physical adsorption onto the pore surface. It is the result of the high affinity of

polymers for many reservoir rocks, for example, due to van der Waal’s and hydrogen

bonding forces (Stutzmann and Sffert 1977, Pefferkorn et al. 1985, Shah et al. 1985,

Sorbie 1991). This retention is believed to be basically irreversible and the amount of

polymer adsorption is proportional to the surface area accessible to polymer molecules.

The second retention is called mechanical entrapment, which happens when polymer

molecules enter pores with smaller outlet diameter relative to the size of polymer

molecule. Small molecules such as water and salt can travel through, but large polymer

molecules will be trapped and accumulate in these small pores (Gogarty 1967; Szabo and

Corp, 1975).

The third retention is called hydrodynamic retention, which is associated with the

local velocity of the polymer. After the retention reaches equilibrium, sudden increase of

flow rate will cause extra polymer loss in the porous media. This flow-related

hydrodynamic retention is believed to be reversible, i.e., when the flow rate is reduced or

flow is completely stopped, the newly-retained polymer molecules will be released and

migrate to the main flow channels (Maerker 1973, Dominguez and Willhite 1977, Huh et

al. 1990). Maerker (1973) suggested that a significant pressure gradient causes polymer

molecules to deform and become trapped within the core, particularly in relatively small

15

pores. Zitha et al (1998) and Chauveteau et al (2002) proposed a mechanism called

“bridging adsorption” to explain hydrodynamic retention. In concept, polymer molecules

may be stretched sufficiently in the elongational flow field during flow through a porous

medium so that the molecules can span the distance over a pore constriction. If the ends

of the molecules attach to the rock, a plugging or increased resistance to flow might

develop. Chauveteau et al. (1974, 2002) suggested that the shear-thickening behavior that

was widely reported for HPAM solutions in porous media was caused by “bridging

adsorption”.

Figure. 2.1 shows the polymer retention model proposed by Szabo and Corp (1975).

In medium-permeability (several hundred millidarcies), high surface-area Berea cores,

physical adsorption is more dominant than mechanical entrapment. By comparison, in

low permeability rocks (several tens of millidarcies), mechanical entrapment is expected

to increase. As shown by Fig. 2.1, adsorption dominates the retention in the main flow

channel, while, mechanical entrapment occurs in the small pores with a pore throat inlet

large enough for polymer molecules to enter but an outlet small enough to trap polymer

molecules. In these pores, though restricted, a slow flow of brine is allowed. It also

demonstrates the concept of polymer inaccessible pore volume (IAPV), i.e., pores with a

small inlet that prevents the polymer penetration will be unreachable for polymer

molecules. Mungan (1969) suggests that if only connate water-not oil-occupies these

small and narrow channels, oil recovery by polymer flooding could be significantly

improved.

16

Fig. 2. 1-Polymer retention mechanisms in porous media (Szabo and Corp, 1975).

2.2.2 Polymer Inaccessible Pore Volume (IAPV)

As mentioned earlier, inaccessible pore volume (IAPV) plays a significant role in

influencing polymer propagation in porous media. Results show that IAPV exists for

EOR polymers (Shah et al. 1978, Vela et at. 1976, Liauh et al. 1978, Lotsch et al. 1985).

The presence of IAPV theoretically accelerates the polymer propagation to be more than

expected from the pore volume injected. On the other hand, polymer retention will retard

polymer transportation in porous media.

The combined effect of IAPV and retention of polymer propagation in the reservoir is

illustrated in Fig. 2.2 (Dawson and Lantz 1972), assuming piston-like displacement.

Case A: no retention and no IAPV. Polymer breaks through at precisely 1 PV;

Case B: no retention, 0.25 PV IAPV. The polymer slug breaks through at 0.75 PV;

Case C: 0.25 PV retention and no IAPV present, polymer bank breaks through at 1.25

PV;

17

Case D: 0.2 PV retention, 0.25 PV IAPV. For this case, the polymer bank emerges at

0.95 PV.

Undoubtedly, if both the retention and IAPV are 0.25 PV, the breakthrough of

polymer solution still happens at 1 PV.

A. No retention, no inaccessible pore volume.

B. No retention, inaccessible pore volume=0.25.

PV

0.25 0.5 0.75 1 1.25 1.5

0.5

0

1

Cp/C0

1.75 2 2.25

PV

0.25 0.5 0.75 1 1.25 1.5

0.5

0

1

Cp/C0

1.75 2 2.25

18

C. 0.25 PV retention, no inaccessible pore volume.

D. 0.2 PV retention, PV inaccessible pore volume=0.25.

Fig. 2. 2-Effect of polymer retention and IAPV on polymer propagation. (Modified from Dawson and Lantz, 1972).

One might expect IAPV to increase with decreasing permeability. However,

confusing results were reported in the literature. For instance, using Pusher 700 HPAM,

Dawson and Lantz (1972) observed almost the same IAPV in 470 mD Berea core (22%)

as in 2090 mD Bartlesville sandstone core (24%). Using Pusher 500 HPAM, Dabbous

(1977) noted an IAPV value of 19% in 761 mD Berea with no residual oil. In contrast, for

the same polymer in Berea with a 28—35% residual oil, the permeability to water ranged

from 49 to 61 mD, and IAPV ranged from 17% to 37%. Osterloh and Law (1998)

reported IAPV values up to 48% in sand packs with permeabilities up to 11 darcies.

However, they acknowledged the experimental difficulties of accurately determining

IAPV values.

PV

0.25 0.5 0.75 1 1.25 1.5

0.5

0

1

Cp/C0

1.75 2 2.25

PV

0.25 0.5 0.75 1 1.25 1.5

0.5

0

1

Cp/C0

1.75 2 2.25

19

2.2.3 Factors Influencing Polymer Retention in Porous Media

Polymer retention in porous media has proven to be a very complicated process and many

factors need to be taken into consideration when dealing with this problem. A broad

range of research has been completed by earlier researchers to unveil polymer retention

mechanisms. In this chapter, special efforts will be made to review the effects of polymer

concentration, polymer injection rate, rock permeability on retention and retention

reversibility because these are also the areas the author attempts to investigate.

2.2.3.1 Effect of Polymer Concentration on Retention.

Numerous studies have been done to investigate the effect of polymer concentration on

retention by researchers. However, most of them were completed using the static method.

No systematic studies on this issue have been carried out using dynamic method.

Mungan (1969) measured the co-polymer (containing approximately 25%

polyacrylate and 75% polyacrylamide) retention on unconsolidated rocks using static

approach. His results revealed that adsorption is higher on rocks with high specific

surface area and it increases with polymer concentration. For instance, retention on

Ottawa sand with BET equal to 0.50 m2/g is 340 g/g sand at polymer concentration of

1,000 ppm (in distilled water). For the same polymer solution, the retention increases to

680 g/g sand when Silica power with BET of 1.65 m2/g is used. In contrast to the

retention at 1,000 ppm, retention decreases to 310 g/g sand at polymer concentration of

250 ppm for the same Silica sand. Dynamic measurement was performed for the

consolidated Berea sand-packed core; the results indicate that the retention is only about

20

55 g/g rock at 500 ppm polymer concentration, which is much lower than the value

obtained from the static method (610 g/g).

Szabo and Corp (1975) studied retention of hydrolyzed polyacrylamide polymers on

sand grains using the static method. They found that polymer retention showed almost

linear dependency on polymer concentration, i.e., the polymer retention increases with

increased concentration. But if sufficient brine or water (over 40 PV dilution) is used for

sand soaking to remove the reversible adsorption, residual polymer retention depends

only slightly on the initial polymer concentration. They ascribed this phenomenon to the

partially reversible adsorption on the surface. Deng et al (2006) measured retention of

three polyacrylamides (PAMs), cationic, nonionic and anionic polymers on clay

minerals—smectite, kaolinite and illite. Their results show that the adsorption isotherms

of anionic PAM and nonionic PAM are L-type and could be fitted with Langmuir

equations. Again, the amount of retention is obtained through static adsorption

measurement.

Retention isotherms from Chiappa et al (1998) show that polymer concentration plays

different roles in polymer retention on pure quartzite. For cationic polyacrylamide (CAT),

retention at high concentration could be several fold higher than that at low concentration.

In contrast to CAT, the retention of 1% hydrolysis polyacrylamide (PAM) shows only

slight concentration-dependence.

Only a few papers mentioned the polymer retention results in porous media via

dynamic measurement; see Table 2.1. Shah et al (1978) measured HPAM polymer

retention in Berea core. The result indicates that retention increases from about 25 g/g

21

sand at concentration of 51 ppm, to 27 g/g sand at 500 ppm and 31.5 g/g sand at 1069

ppm. Therefore, polymer retention only increases a little bit with increased concentration.

Vela et al (1976) measured the retention of 300 ppm and 600 ppm HPAM (5.5 million

Mw, 20% hydrolysis) in porous media and the amounts of retention are almost

independent of the concentration. As mentioned earlier, Zheng et al (1998) claimed

Langmuir isotherm applied to their results; however, they only measured retention for

250, 750 and 1,500 ppm polymer solutions. In addition, the highest retention at 1,500

ppm is less than 1.5 times that from the lowest concentration (250 ppm). Szabo and Corp

(1975) determined retention by injecting polymer solution into 1,200 mD unconsolidated

sand. Polymer retention increased from 3.50 g/g rock at 300 ppm to 6 g/g rock at 600

ppm. This concentration-dependent retention could be ascribed to the similar scenario of

static measurement, since unconsolidated sand was used. Mungan (1969) only measured

retention for 500 ppm copolymer flowing through porous media. Huang and Sobie (1993)

found that retention of Scleroglucan in Bollotini packed columns increased from 8.21

g/g to 11.71 g/g, increasing by 43% as concentration rose from 50 ppm to 200 ppm.

Dawson and Lantz (1972) provided a curve to propose that Langmuir isotherm could be

used to describe polymer retention in porous media, but no actual data measurement was

performed.

22

Table 2.1-Summary of Polymer Retention Using Dynamic Measurement.

References Porous Media Polymer Type Polymer

conc, ppm

Polymer Retention,

g/g rock

Remarks

Shah et al 1978

Berea core HPAM with 5 million Daltons Mw

51 25.0 Same core used

500 27.0

1069 31.5

Zheng et al 1998

623.8 mD Berea core

HPAM with 22 million Daltons Mw

250 4.0 Same core used

750 5.4

1500 5.9

Vela et al 1976

120 mD HPAM with 5.5 million Daltons Mw and 20% hydrolysis

300 12.2 Assuming

=20%,

g=2.65 g/cm3 600 13.9

Szabo and Corp 1975

1,200 mD sand (unconsolidated)

HPAM with 18-20% hydrolysis

300 7.34 Sor=0

600 12.93

Mungan 1969

Berea core

25% polyacrylate+ 75% polyacrylamide, 3-10 million Mw

500 55

Huang and Sobie 1993

Bollotini packed columns

Scleroglucan

50 8.21

100 8.73

200 11.71

Dawson

and Lantz

1972

Propose Langmuir isotherm

Results reported in different literature studies are comparable only if measurements

were conducted under similar conditions. This is especially true when comparing results

from static measurements with those from dynamic measurements. In this study, a series

of tests were designed to clarify literature discrepancies concerning how polymer

concentration affects retention in porous media. Several types of experiments were

performed, including static measurements of polymer retention on fresh sand for each

concentration case, and dynamic measurements of polymer retention in new sandpacks

with similar permeability and porosity for different HPAM concentrations. We also

examined polymer retention measurements where a single sand, sandpack, or sandstone

core was exposed to successive solutions with increasing polymer concentration. HPAM

23

polymer solutions with a broad concentration range (from 10 to several thousand parts

per million) were utilized.

2.2.3.2 Effect of Flow Rate

As mentioned previously, the increase of flow rate will cause additional polymer loss

in porous media by mechanical entrapment. These phenomena have been observed by

many researchers (Maerker 1973, Dominguez and Willhite 1976, Aubert and Tirrell 1980,

Zaitoun and Kohler 1987, Huh et al 1990). By monitoring polymer concentration profile

in the effluent, they found when the retention equilibrium was reached at a low flow rate,

subsequent increase in flow rate would render the effluent concentration lower than the

concentration injected. Conversely, a decrease of flow rate will cause the effluent

concentration to be higher than the injected concentration, which demonstrates that this

flow-induced retention is reversible to some extent.

An example is shown here. Using an 86 md core prepared by compressing Teflon

powder, Dominguez and Willhite (1976) observed that the increase of interstitial velocity

from 3.22 ft/day to 6.32 ft/day caused the effluent concentration to decrease from 391

ppm to 367 ppm. Because polymer adsorption on the Teflon powder was almost

negligible, they believed the extra polymer loss in porous media due to flow rate

variation should be attributed to mechanical entrapment. The results also revealed that a

quantity of this flow-induced retention was reversible (reduction of the flow rate from

10.2 ft/day to 0.38 ft/day caused the normalized concentration to increase from 0.92 to

1.07).

24

By monitoring the mobility reduction (pressure drop during polymer injection divided

by that during brine injection) or residual resistance factor instead of direct measurement

of polymer retention, Zitha et al (2001) proposed a concept of pore-bridging adsorption.

It could cause severe polymer loss at high flow rate if other conditions such as high

polymer adsorption and low permeability are satisfied simultaneously, which leads to an

unsteady-state flow (continuous build-up of injection pressure). Chauveteau et al (2002),

Ogunberu and Asghari (2004) arrive at conclusions that if the shear rate (injection rate) is

greater than the critical value ( > c), the increased hydrodynamic force will push

additional macromolecules into the already adsorbed polymer layer and increase both the

density and thickness of the adsorbed layer.

There is no doubt polymer retention is affected by flow rate, but very limited

retention data is published to quantify this hydrodynamic retention for different flow rates.

A correct description of the flow rate-retention relationship may be of great significance

to the understanding of polymer propagation through reservoir because the polymer flow

rate in reservoirs varies considerably with distance to the wellbore. Other questions

related to this hydrodynamic retention are whether all or part of this retention is

reversible and how does it affect polymer rheology and rock permeability? Our

experiments conducted a thorough investigation on these issues.

2.2.3.3 Degree of Retention Reversibility

Polymer reversibility is a controversial subject. For example, Szabo and Corp (1975)

measured the residual polymer adsorption on the surface of sand grains and found it

decreased with desorption time before it eventually reached a constant level. Therefore,

25

they concluded that retention adsorption on the rock surface is partially reversible. At the

same time, they proposed that little or no desorption occurred in the area of mechanical

entrapment in small pores. Ogunberu and Asghari (2004) found that the residual

resistance factor determined after the polymer retention depended on the brine injection

rate. After the injection of polymer solution at shear rate greater than 110 s-1

, the residual

resistance factor would decrease if the subsequent injection rate of brine was greater than

5.0 ml/min. However, at low flow rates, such as 0.8-2.0 ml/min were applied, an

increased residual resistance factor was encountered. Therefore, they arrived at

conclusion that the polymer adsorbed on the rock surface could be weakened by brine

injection.

On the contrary, experimental results from Maerker (1973), Dominguez and Willhite

(1976) suggested that the polymer retained in the form of mechanical entrapment proved

to be reversible. Zaitoun and Kohler (1987) also showed that at higher clay content or

lower permeability core, reversible polymer retention (mechanical entrapment in finest

pore throats) occurred. Deng et al (2006) observed that after four consecutive washings

with water, an accumulative of less than 3% of the adsorbed polymer (regardless of the

charge type they have) was removed; therefore, the degree of reversibility of polymer

adsorption on the rock surface was negligible.

In our study, we investigated the reversibility of polymer adsorption on rock surface

by determining residual adsorption after washing sands with adsorbed polymer molecules.

For the dynamic case, if present, a negative IAPV could be a good indicator of retention

reversibility. The variation of retention reversibility with the flow rate and rock

26

permeability was also studied in order to provide some ideas on the issues of which kind

of retention, physical adsorption or mechanical entrapment, is more likely to be reversible.

2.2.3.4 Effect of Polymer Retention on Rock Permeability

When the retained polymer molecules form an adsorption layer on the rock surface, the

effective pore size is reduced, resulting in a decrease of rock permeability or increase of

residual resistance factor. This phenomenon becomes more severe when the rock

permeability decreases (Smith 1970, Vela et al. 1976, Seright 1992, Seright and Martin

1993).

Hirasake and Pope (1974) proposed a model that correlated polymer adsorption on

the pore surface with polymer molecular weight, water salinity, rock permeability,

porosity, and flow rate. In their model, the adsorption of polymer is assumed to form a

monolayer of polymer molecular coils with thickness approximately equal to the

diameter of the molecular coil in that particular solvent based on previous findings

(Rowland 1963, Rowland and Eirich 1966). This layer may be laterally compressed,

resulting in an increase in segment density. The increase of segment density results in

increased polymer loss due to adsorption, but will not affect the adsorbed layer thickness.

Therefore, the permeability will not be further reduced.

When polymer molecules adsorb on pore surface and form a thin layer, the effective

pore size will be reduced, resulting in a reduced permeability. However, in the literature,

few people have reported measuring how the resistance factor and the residual resistance

factor vary with hydrodynamic retention. In our study, by measuring residual resistance

factors after different concentration injection with same injection rate and also the same

27

polymer solution with different injection rates, we addressed the question whether the

rock permeability varies dramatically when remarkable hydrodynamic retention occurs.

2.2.3.5 Other Factors Influencing Polymer Retention

Other factors influencing polymer retention in porous media have also been studied.

Smith (1970) found that the adsorption of HPAM varies from one type of mineral to

another. For instance, the retention on calcium carbonate is five times higher than that on

silica, which shows calcium carbonate appears to have a much greater affinity for

polymer than silica. This increased retention is ascribed to the high content of calcium

ions on the surface, which may provide calcium bridges to enhance polymer retention.

They also found polymer adsorption increases with salt concentration. The amount of

polymer retained increases from about 11 g/m2 at 1% NaCl to 60 g/m

2 at 10% NaCl.

Broseta et al. 1995 found that in oil-wet porous media, polymer retention (PAM) in

the presence of residual oil saturation (Sor=0.1-0.19) will decrease considerably by factors

ranging from 2 to 5 compared to the retention when the core is 100% water-saturated. But

in water-wet porous media, the influence of residual oil saturation is less noticeable. For

example, the retention with Sor=0.2 is 7.5 g/g sand compared with the retention of 10

g/g sand at Sor=0. They suggest the variation in polymer retention is due to the change

of interfaces accessible to the polymer under these conditions.

Chiappa et al (1998) investigated the role of electrostatic interactions in polymer

adsorption. They tested polymers with different charges (cationic, anionic and weakly

28

anionic) on quartzite, which is negatively charged. Effects of clay content with a high

specific surface area, ion strength and composition were also studied. Their results

showed that polymer adsorption is dominated by electrostatic interactions between the

charged groups that present at the polymer/brine and rock/brine interfaces. A correct

match between the polymer and the surface charges can greatly increase adsorption. For

example, adsorption on negatively-charged quartzite increased from 270 to 340 and 610

g/g sand when polymer of HPAM (anionic), PAM (weakly anionic) and CAT (cationic)

were used, respectively. Because of the high surface area and predominately negative

charge of the clay mineral, a small amount of clay can cause a significant increase in

polymer retention. The retention of cationic polymer increased from 610 to 1.45*104,

1.8*105 g/g sand when the porous media was switched from pure quartzite to 8% clay

quartzite with 8% clay and 100% clay. Again, the presence of divalent cation (as Ca2+

)

can greatly enhance the adsorption of negatively changed polymers (HPAM and PAM)

onto quartzite. For instance, HPAM retention of approximately 80, 340 and 800 g/g

sand was determined on the pure quartzite in the system containing 0, 2% and 8% CaCl2.

Chiappa et al also suggest that the present of divalent calcium ions can enhance the

adsorption of negatively charged polymers by forming an ion bridge.

Efforts were also made to distinguish the adsorptive retention from the mechanical

entrapment retention by Cohen and Christ (1986). In their study, they used HPAM with

an estimated molecular weight of 5.5 million and degree of hydrolysis of 25%. Two kinds

of packed silica sand beds were applied as the porous media. One was an adsorbing

material and the other was a non-adsorbing material generated by the chemical

modification of a siliceous surface. Their results showed that adsorption accounted for

29

about 35.2% of the total polymer retention and the remaining 64.8% was attributed to

mechanical entrapment.

2.3 Langmuir Adsorption Isotherm

Some researchers proposed that retention of EOR polymers on reservoir rocks depends

on polymer concentration, or the Langmuir adsorption model applies. Therefore, it is

necessary to make a brief introduction to this adsorption model. Equation 2.1 describes

the Langmuir adsorption isotherm which shows the solute adsorption on the substrate

surface is a function of solute concentration.

1 1

11

a b C

b C……………………………………………………………………………..2.1

where, is solute adsorption. C is solute concentration in solution and a1, b1 are

constants.

30

Fig. 2. 3-Typical Langmuir adsorption isotherm.

For adsorption fits to the Langmuir isotherm model, constants of a1 and b1 can be

determined graphically. Plotting 1/ versus 1/C on a linear scale ends up with a straight

line. The slope of this line is 1/a1b1 and it intercepts with y-axis at 1/a1.

Fig. 2.3 shows a typical Langmuir adsorption isotherm where a1 and b1 are assumed

to be 25 and 0.02, respectively. For the Langmuir adsorption model, the adsorption

depends strongly on the concentration. Especially in the low concentration range, the

adsorption increases linearly with the concentration. When concentration approaches zero,

the adsorption is also decreasing to zero.

0

5

10

15

20

25

0 500 1000 1500

Ad

sorp

tio

n,

g/g

rock

Concentration, ppm

a1=25 b1=0.02

31

2.4 Concluding Remarks

The literature review shows that retention is a very complex process. Though several

mechanisms are proposed to elucidate this phenomenon, many disagreements still exist.

For instance, what role does polymer concentration plays in polymer retention? Besides

the claim made by Dawson and Lantz (1972) without actual retention measurement, some

researchers (Mungan 1969, Szabo and Corp 1975, Deng et al 2006) found that polymer

retention is a function of polymer concentration; or, the retention follows the Langmuir

isotherm based on the static measurement. However, a careful analysis of limited

experimental results from dynamic measurement (Szabo and Corp 1975, Vela et al 1976,

Shah et al 1978, Zheng et al 1998) shows that retention is almost independent of polymer

concentration. A systemic study on the effect of concentration on retention using

dynamic method is highly recommended.

Researchers (Maerker 1973, Dominguez and Willhite 1976, Aubert and Tirrell 1980,

Zaitoun and Kohler 1987, Huh et al 1990) observed flow-induced, or hydrodynamic,

retention by monitoring polymer concentration in the effluent. Nevertheless, few studies

found in the literature quantified retention for different rates. This may be of great

importance if retention is highly velocity-dependent because the flow velocity in the

reservoir varies considerably as the invasion radius changes.

Regarding to the degree of retention reversibility, results from Szabo and Corp (1975)

and Ogunberu and Asghari (2004) suggest polymer adsorption on the rock surface shows

partially reversible behavior. On the contrary, Deng et al (2006) conclude that the

reversibility of adsorption is almost negligible. Results from Dominguez and Willhite

(1976) Zaitoun and Kohler (1987) indicate polymer molecules retained in the form of

32

mechanical entrapment proves to be reversible. In our study, we will investigate the

reversibility of polymer retention on rock surface and in porous media. This may provide

some ideas on the issue of which kind of retention, physical adsorption or mechanical

entrapment, is more likely to be reversible.

Polymer retention on the rock surface forms an adsorption layer that reduces effective

pore size and causes permeability reduction. Some researchers propose that the increase

of flow rate may either induce pore-bridge adsorption ,which results in an unsteady-state

flow (continuous injection build-up) (Zitha et al 1998 and 2001) or increases both the

density and thickness of the adsorbed layer (Chauveteau et al 2002, Ogunberu and

Asghari 2004). If it is true for either case mentioned above, the polymer resistance

factor/residual resistance will be dramatically increased. In our research, we probed if

severe pore-bridge adsorption occurred in our tests by recording the polymer injection

pressure under various conditions. We also focused on the relationship of hydrodynamic

retention and permeability reduction.

In summary, to better understand the retention behaviors of HPAM polymers in

porous media, as mentioned previously, the following issues were addressed in our study:

1) Does polymer retention in porous media depend on polymer concentration? Or,

does it follow the Langmuir isotherm?

2) How should quantify hydrodynamic retention be quantified for different rates?

3) Under what circumstances, does polymer retention becomes reversible?

4) Does hydrodynamic retention dominate polymer rheology in porous media?

5) How does polymer retention affect rock permeability?

33

CHAPTER 3. METHODS AND PROCEDURES

3.1 Introduction

In this section, the materials, experimental equipment, and procedures will be introduced.

Again, both static and dynamic measurements will be used to determine polymer

retention in porous media.

3.2 Equipment and Material

Polymer and Brine. A partially hydrolyzed polyacrylamide (HPAM) (SNF Flopaam

3230S) and a xanthan polymer (Kelco Oil Field Group) were used in our tests. Both were

provided by the manufacturer as white granular powders. HPAM is estimated to have a

molecular weight of 6—8 million daltons and degree of hydrolysis of approximately 30%.

HPAM solution was prepared using the magnetic stirrer vortex method. Xanthan solution

was prepared using a blender. After the polymer solutions were preparation, they were

filtered through a 10 m filter to remove any possible microgels and other debris present

in the solution. The purpose of this filtration is to minimize the face plugging effect

caused by these impurities. Studies show that during polymer injection, presence of this

34

debris and microgel may plug the core face by forming external filter cake (Seright et al.

2009).

Two kinds of brine were used. One was 2% NaCl for the static measurements,

dynamic retention in sandpacks, and polymer hydrodynamic retention measurements in

consolidated cores. The other brine containing 2.52% TDS (2.3% NaCl and 0.22%

NaHCO3) was used when dynamic retentions were measured in consolidated sandstone

cores. Both brines were filtered through 0.45- m filters before application.

The rheology of HPAM polymer was determined in an Anton Paar rheometer

(Xanthan rheology will be shown in Chapter 4). As shown in Fig. 3.1, at concentration

below 320 ppm, it behaves like a Newtonian fluid within the broad range of shear rates

between 1 to 1,000 s-1

, i.e., polymer viscosity is almost independent of shear rate.

Polymer solutions with concentration of 640 and 1,000 ppm, at shear rate less than 10 s-1

,

they show Newtonian behavior. For shear rate greater than 10 s-1

, they show slightly

shear thinning. No shear thickening behavior is observed in a viscometer. The correlation

of viscosity at shear rate of 7.3 s-1

with polymer concentration is shown by Fig. 3.2. The

viscosity of 20 ppm is about 1 cp, increasing to 3.8 cp at concentration of 1,000 ppm.

Tracer. 40 ppm Potassium iodide (KI) was added into the polymer solution as a tracer.

Its concentration in the effluent was monitored by a Tunable Absorbance Detector

(Waters 486) as an indicator of brine propagation through porous rock.

35

Fig. 3. 1-Rheology of HPAM polymer in a viscometer.

Fig. 3. 2-Viscosity vs. concentration at shear rate of 7.3 s-1.

0

2

4

6

1 10 100 1000 10000

Vis

cosi

ty, c

p

Shear rate, 1/s

20 ppm

80 ppm

320 ppm

640 ppm

1,000 ppm

HPAM 6-8 million MW 30% hydrolysis 2.52% TDS 25 °C

0

1

2

3

4

0 200 400 600 800 1000 1200

Vis

cosi

ty, c

p

Polymer concentration, ppm

HPAM 6-8 million MW 30% hydrolysis 2.52% TDS 25 °C

36

Sand Preparation. Sand grains with particle sizes between 106–180 m were prepared

as the adsorbent by crushing and sieving Berea sandstone core cuttings. To reduce the

presence of very fine particles and carbons released by the sands, special processes were

undertaken for the treatment of these disaggregated sands. First, sands were put into a

bottle with brine and rotated at 300 rpm for 8 hours on an IKA KS 4000 shaker, (Fig. 3.3).

The purpose of mixing sands with brine is to minimize the release of carbon with the

sands themselves. Then, the upper mud-like phase was separated from the sand. Next, the

sand was washed with distilled water to remove newly-generated fine particles and

residual salt until the upper water phase was totally clear. Finally, the sand was dried at

110 °C.

Fig. 3. 3-Sand shaker (IKA KS 4000).

Porous Media. Disaggregated sands prepared as described above were used for static

adsorption measurements. To determine dynamic and hydrodynamic polymer retention in

37

porous media, both consolidated sandstone cores and high permeability sandpacks were

used. Four consolidated sandstone cores were used in our tests, among them, three

rectangular Dundee cores and one cylindrical Berea sandstone core. Dundee sandstone

cores cast in epoxy resin (Core #1, #2 and #3) have permeability of 347 mD, 449 mD,

and 1.9 darcies respectively. The fourth core (Core #3) is a Berea sandstone core with a

permeability of 71 mD. It is a cylindrical core with a section area of 11.4 cm2. This core

was cast in the metal before being assembled in the Hassler-type core holder. All cores

were 15 cm long except core #4, which is 12.8 cm long. Two internal pressure taps

divided the Dundee cores into three sections with lengths of 2.5 cm, 10 cm, and 2.5 cm.

For the Berea core, these three sections were 1.5 cm, 10 cm, and 1.5 cm in length.

Sandpacks with high permeability and porosity were prepared from the same sands used

for static measurements. Sandpacks were 6.35 cm long and 14.5 cm2 in cross section.

Table 3.1-Core Properties.

Core No. L, cm A, cm2 PV, ml , % k, mD Note

1 15 14.5 49 22.5 347

Dundee 2 15 14.5 51 23.4 449

3 15 14.5 52.4 24.1 1,900

4 13 11.4 27 18.2 71 Berea

Experimental Setup. Figure. 3.4 shows the schematic diagram of the experimental unit

for determining polymer retention in porous media. It can be divided into three major

sections based on their functionalities: polymer injection, tracer concentration

determination and effluent polymer concentration determination. All these three parts

were assembled in series.

38

1) The first part deals with polymer injection into the core. The polymer flow rate can be

accurately controlled by an ISCO syringe pump (Model 500HP). The pressure drop

across the core is indicated by a Honeywell pressure transducer, which is connected

to the two internal pressure taps on the core.

2) The second part is setup to determine tracer concentration in the effluent. In this part,

the fluid exiting the core outlet flows through the absorbance detector (Waters 486)

and the concentration of KI can be measured versus pore volume injected at light

wavelength of 232 nm. Note that a 7 m Swagelok metal filter is attached in the flow

line between the core and the absorbance detector to prevent any large particulates

from flowing downstream.

3) The third part is used for the determination of polymer concentration in the effluent.

As shown in Fig. 3.4, the effluent leaving the absorbance detector is first collected in

a container and then fills the small accumulator beneath the container via the second

ISCO pump every 10 or 20 minutes. Then, the fluid is forced to flow through a 10 m

Millipore filter combination at a constant flow rate (controlled by another ISCO pump)

and the pressure drop across the filter is recorded. For HPAM polymer concentration

higher than 150 ppm, the filter combination can be connected directly to the core.

39

Fig. 3. 4-Schematic diagram of polymer retention determination system.

Note: 1. ISCO syringe pump #1 (Model 500HP); 2. Core; 3. Pressure transducer #1; 4. 7

m Swagelok filter; 5. Absorbance detector (Waters 486); 6. Beaker; 7. 10 m Millipore

filter combination; 8. Pressure transducer #2; 9. Fluid container; 10. Accumulator; 11.

ISCO syringe pump #2 (Model 500D).

3.3 Experimental Procedures

Static Measurement. After mixing known concentration solution and known mass sands,

bottles containing both sand and polymer solution were tied on a roller which rotate at a

speed of 6 RPM for 1 hr to complete the adsorption process. One hour contact was

considered to be sufficient because polymer adsorption on the rock surface is believed to

be instantaneous (the adsorption kinetic will be discussed in Chapter 4). To reduce the

effect of carbon released by sands themselves, a blank sample only containing sands and

brine was prepared.

1) Polymer solutions with known concentration were prepared in 2% NaCl brine;

2) Known amounts of sand grains were added into the polymer solutions;

40

3) The bottle containing both polymer solution and sand grains was mounted on a roller

(Fig. 3.5). The system was rotated at 6 rpm for 1 hr;

4) After the rotation, the upper polymer solution was transferred to a plastic tube to be

centrifuged at speed of 3,000 rpm for about 1 hr to separate the residual polymer

solution from sand particles;

5) Equilibrium polymer concentration was determined by Total Organic Carbon (TOC)

Analyzer. Dilution was needed for high concentration cases;

6) Polymer adsorption for each concentration case is calculated:

0( ) /eq p sgC C V W

where, is polymer adsorption, g/g sand, C0 and Ceq are initial and equilibrium

polymer concentrations, ppm. Vp is polymer volume, ml. Wsg is the weight of sand

grains, g. Both polymer and brine density were assumed to be 1 g/ml.

Fig. 3. 5-Roller for static measurement.

41

Determination of Polymer Concentration. A TOC analyzer (Shimaduz Model TOC—

VCSH, Fig. 3.6) was used to determine polymer concentration before and after

adsorption. This is based on the excellent linearity between total organic carbon content

and polymer concentration. As seen from Fig. 3.7, a very good correlation exists between

these two parameters. To reduce systematic error, samples with concentration higher than

200 ppm were diluted with the same brine for polymer preparation before examination.

The reason for diluting high polymer concentration solutions is because a range of 0 to

200 ppm TOC calibration curve was used.

Fig. 3. 6-Total Organic Carbon (TOC) analyzer for concentration determination.

42

Fig. 3. 7-Correlation between TOC and polymer concentration.

Dynamic Measurement. The detailed procedures of this method are described below

and the determinations of polymer retention and IAPV are shown graphically by Fig. 3.8:

1) Core was initially saturated with degassed brine after evacuation and then the

permeability to water was determined. Based on the amount of brine saturated and the

core dimension, total pore volume (PV) and porosity were obtained.

2) Polymer solution was injected with KI tracer until polymer concentration Cp in the

effluent achieved the injected concentration C0. Polymer concentration and tracer

concentration (relative to the injected concentrations, Cp/C0) were plotted versus pore

volumes injected.

y = 1.8611x + 0.706 R² = 0.994

0

20

40

60

80

100

120

0 10 20 30 40 50 60

Po

lym

er c

on

cen

trat

ion

, pp

m

TOC, ppm

43

3) Core was flushed with brine until polymer was not detectable in the effluent. In our

tests, 60 to100 PV of brine were injected with intervening periods of no flow.

4) A second polymer and tracer bank were injected with the same concentration until Cp

achieved the injected concentration C0. Again polymer and tracer concentrations

(Cp/C0) were plotted versus pore volumes injected.

5) Retention is given by the area between the two plots of polymer concentration

breakout curves.

6) IAPV is given by the area between the second polymer and tracer concentration

breakout curves.

Fig. 3. 8-Polymer retention and inaccessible pore volume (IAPV) determination.

44

Note: all the tests are conducted at room temperature around 25 °C. Unless otherwise

specified, polymer solution is injected at a flow rate of 60 ml/hr which is about 1

meter/day or 3.3 ft/day.

Attention should be paid when this method is utilized to determine polymer retention

and IAPV. Firstly, any reversibly retained polymer should be flushed from the core

during the extensive brine injection. The reoccurrence of this retention during the second

polymer injection will make the effluent concentration profile shift closer to the first one.

As a consequence, reversible retention is excluded or only irreversible retention is

measured via this dynamic method. Secondly, IAPV measurement will be affected if

substantial reversible retention occurs. Both of these two situations were encountered in

our tests and will be discussed in the later sections.

Determination of Polymer Concentration in the Effluent. How to accurately monitor

polymer concentration in the effluent is very important for this test. After some

experimentation, we established a rheological method based on shear-thickening

behaviors of HPAM polymer flowing through porous media. HPAM polymers show

Newtonian behavior at low flow rate in the porous rock, while, at moderate to high flux,

they become definitely shear thickening, i.e., the resistance factor increases with

increased flux. Results from Seright et al. (2011) show that shear thickening behavior of

HPAM solution with a concentration even as low as 25 ppm is evident in porous media.

45

Fig. 3. 9- p vs. Cp when HPAM flowing through a 10 m filter.

Based on this finding, we established a filter combination that uses a Millipore

AP10TM

filter pad upstream of a 10 m Sterlitech membrane filter to mimic porous media.

When polymer solutions flow through this filter combination at constant flow rate,

pressure drops given by the Honeywell pressure transducer correlate well with polymer

concentrations. Fig. 3.9 is one standard curve that shows the sound linear relationship

between polymer concentration and pressure drop.

3.4 Polymer Injection at Different Concentrations

To investigate the effect of polymer concentration on retention, a series of polymer

solutions with concentration from 20 ppm to 1,000 ppm were injected into the Dundee

sandstone core. As described previously, for one specific concentration, two identical

polymer solution banks with tracer were injected which were separated by a brine slug

y = 0.0082x + 0.207 R² = 0.99

0

2

4

6

8

10

0 200 400 600 800 1000 1200

Pre

ssu

re d

rop

, psi

Polymer concentration, ppm

HPAM 6-8 million Daltons Mw 30% hydrolysis 2.52% TDS 40 ppm KI 300 ml/hr 10 micron filter 25 C

46

injection. The tracer and polymer concentrations in the effluent were determined by

Absorbance Detector and Millipore filter combination. By plotting Cp/C0 versus pore

volumes injected for the first and second polymer slugs and second tracer slug, both

polymer retention and inaccessible pore volume can be determined. Again, polymer

retention is given by the area between the two polymer concentration curves and IAPV is

given by the area between the second tracer and second polymer concentration curves.

This process was repeated for the next concentration case. The tests were run in the

sequence of low concentration to high concentration. All the tests were performed under

room temperature, which is approximately 25°C.

3.5 Polymer Injection at Different Flow Rates.

Hydrodynamic retention was measured by injecting polymer solutions in core #3 and #4

at various flow rates. Again, two polymer banks were injected separated by large pore

volumes of brine injection, and effluent polymer concentrations were recorded. To

determine whether this flow-related retention significantly affects polymer rheology in

porous media, both HPAM and xanthan polymers were tested. Polymer pressure drops

across the core during polymer injection and subsequent brine injection for these cases

were recorded. These pressure drops were used to calculate polymer resistance factors

and residual resistance factors.

47

CHAPTER 4. RESULTS AND DISCUSSIONS

4.1 Introduction

This chapter deals with the experimental results and discussions. Mainly, two sections are

included here. First, is an account of the investigation of the dependence of retention on

polymer concentration. Porous media, such as disaggregated sands, high permeability

sandpacks and low permeability sandstone cores were used. Based on the observed

retention behaviors, a concentration-related retention model is proposed. The second part

of this chapter focuses on hydrodynamic retention of both HPAM and xanthan polymers.

A method is established to show, as flow rate increases, the amount of total incremental

retention and how to distinguish irreversible and reversible retention. To clarify whether

this flow-induced retention has strong impact on polymer rheology in porous media, the

rheology and retention of both HPAM and xanthan polymer in porous media were

examined.

The evaluation of reversibility of retention under different flow rates and rock

permeabilities and the effect of reversibility of polymer retention on IAPV is also

discussed. In this section, we will also address the question of how permeability

48

reduction (residual resistance factor) is affected by polymer retention and under what

circumstances, the increase of polymer retention will reduce rock permeability.

4.2 Dependence of Retention on HPAM Concentration

Both static and dynamic measurements were used to estimate the impact of concentration

on retention. Porous media employed include disaggregated sand grains, high

permeability sandpacks and low permeability consolidated sandstone cores. Furthermore,

retention in pre-contacted porous media was also studied before the proposition of a

concentration-related adsorption model. Here, the term “pre-contacted porous media”

refers to the porous media whose retention is previously satisfied at low concentration.

4.2.1 Static Measurements

Adsorption Kinetics. The kinetics of polymer adsorption were first explored by mixing

100-ppm polymer solution with disaggregated sands. The bottle containing both polymer

solution and sands was mounted to a roller and rotated at 6 RPM for 20 hrs. Liquid

samples were taken periodically from the upper phase for polymer concentration

determination using TOC analyzer. The adsorption calculated based on mass balance was

plotted as a function of time. As shown by Fig. 4.1, adsorption reached the maximum

plateau (about 40 g/g sand) within about three minutes and then leveled off. This

indicates that the long chain HPAM molecule shows a very high adsorption tendency on

rock surface and that the adsorption can be completed instantaneously. In our tests, the

contact of polymer solution and sands lasted 1 hr to ensure adsorption equilibrium.

49

Fig. 4. 1-Kinetics of polymer adsorption on sand.

Desorption Test. After the completion of adsorption, desorption tests were carried out to

estimate the amount of polymer that can be removed from the surface. In this test, excess

polymer solution was decanted from the top of the sand after sufficient contact. Then

fresh brine was added and again the bottle containing both sand and brine was rotated at 6

RPM for 1 hr. After the sands settled, the upper phase was sampled for polymer

concentration determination. The residual polymer adsorption was calculated using mass

balance. This procedure was repeated until no more desorbed polymer was detected.

Figure. 4.2 shows the results for 100-, 500-, and 1,000-ppm HPAM. Calculations show

that the percentage of the reversible adsorption for these three cases was 6.6%, 2.4%, and

2.9%, respectively. This result was similar to that from Chauvetear and Kohler (1974),

Deng et al. (2006). Because EOR polymers have high molecular weights and extended

chains, many polar groups along the polymer chain will attach to many different polar

points on the rock surface. It is statistically very unlikely that a polymer molecule would

1

10

100

1 10 100 1000 10000

Ad

sorp

tio

n,

g/g

san

d

Time, min

HPAM 6-8 million Daltons Mw 106-180 µm sands 100 ppm Rotated at 6 rpm 2% NaCl 25°C

50

release all points of attachment at the same time. Therefore, polymer adsorption on the

sand surface can be treated as almost irreversible.

Fig. 4. 2-Desorption tests for 100-, 500-, and 1,000-ppm HPAM.

Effect of Polymer Concentration. To investigate the effect of polymer concentration,

retention from polymer solutions with concentrations from 10 ppm through 6,000 ppm

was examined. The results are illustrated in Fig. 4.3, which suggests three distinct

concentration-related retention behaviors. First, in the very low concentration region

(from 10 ppm to about 100 ppm), polymer retention stabilized approximately at a value

of 20 g/g. In the intermediate-concentration region (from 100 ppm to about 4,000 ppm),

polymer retention increased from 35 to 420 g/g, increasing almost linearly with polymer

concentration. In the very high concentration region (above 4,000 ppm), nearly constant

retention (~ 420 g/g) was achieved.

0

50

100

150

200

250

300

0 1 2 3 4

Res

idu

al a

dso

rpti

on

, g/

g sa

nd

Brine/sand ratio, (weight/weight)

100 ppm 500 ppm 1000 ppm

51

These results (especially the concentration-dependent observation) agree with the

previous findings where most of the measurements were made in the intermediate

concentration region (Mungan 1969; Espinasse and Siffert 1979). Our findings indicate

that polymer retention does not fit the Langmuir isotherm, which is commonly used to

describe the reversible adsorption of small molecules such as surfactants and gas. For

EOR polymers with high molecular weights and extended chains, their adsorption on

rock shows little reversibility (see Fig. 4.2). It is postulated that at very low concentration,

polymer molecules continue to be adsorbed until the maximum coverage is reached.

During this process, few adsorbed polymer molecules are likely to detach from the

surface. Therefore, unlike the adsorption described by the Langmuir isotherm, polymer

adsorption at very low concentration approaches a constant non-zero value.

Fig. 4. 3-Adsorption isotherm of HPAM using static method.

Re-Adsorption Test. Fresh sands were used for each case to generate the adsorption

isotherm shown in Fig. 4.3, which illustrates the concentration-related adsorption

1

10

100

1000

1 10 100 1000 10000

Ad

sorp

tio

n,

g/g

san

d

Polymer concentration, ppm

106-180 m sand grains Rotated at 6 rpm for 1 hr 2% NaCl 25 °C

52

behaviors. After the desorption tests described in Fig. 4.2, 1,000-ppm polymer solution

was added to the sands previously contacted with 100-ppm and 500-ppm polymer

solution to check if polymer re-adsorption occurred. The results (Fig. 4.4) show little

additional polymer was adsorbed onto the used sands. For instance, for the 100-ppm

concentration case, the retention increased from 32.4 to 35.8 g/g, increasing by 10.3%.

For the 500-ppm case, retention rose from 132.9 to 141.1 g/g—merely a 6.1% increase.

Compared to the adsorption of 243 g/g at 1,000 ppm, a substantial retention difference

existed between the fresh sands and pre-contacted sands whose retention was previously

satisfied by low-concentration polymer. Apparently, even though adsorption was

relatively small at low concentration, the surface was already fully covered by adsorbed

polymer molecules, and no vacant sites were available for further attachment.

Fig. 4. 4-Comparsion of retention on fresh sands and used sands.

32.4 35.8

132.9 141.1

243.0

0

50

100

150

200

250

300

1 2 3 4 5 6 7

Ad

sorp

tio

n,

g/g

san

d

106-180 µm sand grains Rotated at 6 rpm for 1 hr 2% NaCl 25 °C

Adsorption at 100 ppm

Re-adsorption at 1,000 ppm

Adsorption at 500 ppm

Re-adsorption at 1,000 ppm

Adsorption at 1,000 ppm

53

4.2.2 Dynamic Measurements

Retention in Sandpacks. Dynamic measurements were performed in sandpacks made

from the same sand source that was used for static measurements. Sandpacks were

constructed by tapping sands into a cylindrical Teflon-made container with both ends

sealed by O-rings. A pair of stainless steel meshes with pore size of 15 microns was

placed on both ends of the container to prevent any downstream clogging and 0.125-in

OD tubing was welded onto end caps. Procedures were followed as described in Chapter

3 for dynamic measurement. For each concentration case, a new sandpack was used.

As shown in Table 4.1, these sandpacks had very similar properties. Permeability

ranged from 4.69 to 5.51 darcies, and porosity ranged from 0.43 to 0.44. Due to the high

sandpack permeability, it was suggested that adsorptive retention dominated the retention

for these cases (Szabo and Corp 1975, Huh et al. 1990). Polymer solutions with

concentrations of 20, 50, 100, 500, 1,000 and 2,000 ppm were investigated at an injection

rate of 120 cm3/hr (6.6 ft/day flux). The results were shown in Table 4.1 and Fig. 4.5.

Retention was around 5 g/g at low concentrations from 20 ppm to 100 ppm. With the

increase of concentration from 100 ppm to 2,000 ppm, retention increased from 5.71 to

27.8 g/g—increasing by a factor of nearly 5.

After completion of measurements for 100- and 500- ppm cases (Sandpacks #4 and

#3 in Table 4.1), a 1,000-ppm solution was injected. Retention increases of 5.6% and

7.3% , respectively, were detected in two used sandpacks. This agrees with the results

from the static measurements on used sands, which also confirms that the adsorbed

54

molecules occupied almost all the vacant sites on the sand surface and prevented further

attachment.

Table 4.1-Dynamic Retention in Sandpacks

SP No. L, cm

A, cm

2

Wsd, g PV, cm

3 k, D Cp, ppm Rpret, g/g

sand

1

6.25 14.19

140.4 39.1 0.441 5.51 2,000 27.8

2 139.7 39.0 0.440 5.04 1,000 14.3

3 140.2 38.7 0.436 4.88 500 10.2

4 141.4 38.3 0.432 4.69 100 5.71

5 141.0 38.7 0.436 5.03 50 4.85

6 140.6 39.1 0.441 5.37 20 4.63

Fig. 4. 5-Adsorption isotherm using dynamic method (fresh sandpacks used for each case).

Among this series of tests, two examples are given here. Figure. 4.6 shows the first

and second effluent polymer concentration profiles for 50 ppm HPAM injection. The area

between these two breakout curves is an output of polymer retention in porous media.

The same test was carried out for 500-ppm polymer solution, and with the result shown in

Fig. 4.7. Retentions calculated for 50 ppm and 500 ppm are 4.85 g/g sand and 10.2 g/g

1

10

100

10 100 1000 10000

Ad

sorp

tio

n,

g/g

san

d

Polymer concentration, ppm

HPAM 6-8 million Daltons Mw 30% hydrolysis 2% NaCl, 120 ml/hr 25 °C

55

sand, respectively. We can see that when polymer concentration rises from 50 ppm to 500

ppm, retention increases by about 110%.

Fig. 4. 6-Retention determination for 50 ppm HPAM.

Fig. 4. 7-Retention determination for 500 ppm HPAM.

0

0.2

0.4

0.6

0.8

1

1.2

0 1 2 3 4

Co

nce

ntr

atio

n r

atio

(C

p/C

0)

Pore volumes injected, pv

1st polymer injection

2nd polymer injection

5.03 Darcies sandpack 50 ppm HPAM

0

0.2

0.4

0.6

0.8

1

1.2

0 1 2 3

Co

nce

ntr

atio

n r

atio

(C

p/C

0)

Pore volumes injected, pv

1st polymer injection

2nd polymer injection

4.88 Darcies sandpack 500 ppm HPAM

56

Static measurements show that no significant re-adsorption occurred in used sands

(Fig. 4.4). To check whether this is also true for dynamic measurement, 1,000 ppm

polymer solution was injected through the sandpack whose retention was originally

satisfied at 500 ppm (SP No. 4 in Table 4.1). Approximately 0.75 g/g sand (Fig 4.8) or

7.3% additional retention was determined in this sandpack. Again, the re-occurrence of

retention is negligible for pre-contacted porous media, even though the injected solution

had much higher concentration than the previous one.

Fig. 4. 8-Retention of 1,000 ppm in pre-treated sandpack with 500 ppm.

Retention in Sandstone Cores. To date, polymer retention in high permeability

sandpacks has been evaluated where adsorptive retention is believed to dominate.

However, besides adsorption on rock surface, mechanical entrapment occurs

simultaneously in pore throat constrictions and dead-end spaces when consolidated cores

are used (Huh et al. 1990, Ranjbar et al. 1991). To investigate the dependence of polymer

0

0.2

0.4

0.6

0.8

1

1.2

0 1 2 3

Co

nce

ntr

atio

n r

atio

(C

p/C

0)

Pore volumes injected, pv

1st polymer injection

2nd polymer injection

4.88 Darcies Sandpack Pre-treated with 500 ppm HPAM 1,000 ppm HPAM

57

retention on concentration in less permeable porous media, retention of HPAM with

different concentrations was measured in several consolidated sandstone cores. The first

measurement was performed using 347 mD Dundee sandstone core (see the description

of this core in Chapter 3). In this experiment, a series of polymer solutions were injected

sequentially through this core at a rate of 60 ml/hr. Concentrations of 20, 40, 80, 100, 200,

400, 700 and 1,000 ppm were considered. A higher flow rate was used when dealing with

the hydrodynamic effect on retention; this will be discussed in the next section.

For this series of tests, the same sandstone core was repeatedly used and retention

measurement followed the concentration sequence from low to high. For instance, the

retention at 40 ppm was measured only after the retention at 20 ppm (the lowest

concentration) was completed. Accordingly, the retention for the highest concentration

case (1,000 ppm for this series of tests) was carried out in the final run. The results are

shown in Fig. 4.9. From 20 ppm through 1,000 ppm, all the retention values fall into the

range of approximately 15—16 g/g sand with a relative error of 1.1%. When polymer

concentration was greater than 100 ppm, the retention reached a plateau, with the

maximum retention of approximately 16 g/g rock. Therefore, if one core was repeatedly

used, polymer concentration showed almost no influence on polymer retention. The result

was consistent with other studies where the same core was used repeatedly (Shah et al.

1978, Zheng et al. 1998).

Fig. 4.10 shows two effluent concentration profiles from the following 80 ppm

polymer solution injection. The small area between these two curves represents the

increased retention, which is only 0.7 g/g sand.

58

Fig. 4. 9-Effect of concentration on polymer retention, 347 mD core.

Fig. 4. 10-Retention determination for 80 ppm HPAM, 347 mD core.

0

4

8

12

16

20

10 100 1000

Po

lym

er r

eten

tio

n,

g/g

san

d

Polymer concentration, ppm

Dundee sandstone core 347 mD Porosity: 23%

0

0.2

0.4

0.6

0.8

1

1.2

0 1 2 3 4 5

Co

nce

ntr

atio

n r

atio

(C

p/C

0)

Pore volumes injected, PV

1st polymer injection

2nd polymer injection

Dundee sandstone core Porosity: 23% 347 mD

59

The above results were obtained from relatively high permeability core (Dundee

sandstone core with permeability of 347 mD). For the retention in low permeability

porous rock, a 71 mD Berea sandstone core was used. The same strategy was followed to

determine the polymer retention at different polymer concentrations. As shown in Fig.

4.11, the retention increases a little bit with the increase of polymer concentration, but

compared with the concentration increase from 20 ppm to 1,000 ppm which increases by

50 times, the increase of polymer retention is very small, only about 1.6 times. Figure.

4.12 demonstrates the effluent concentration profile for the first and second polymer slug

injections. For the 20 ppm case, around 16 PV of polymer were injected to satisfy the

polymer retention. Nevertheless, for the following 40 ppm case, shown by Fig. 4.13,

fewer than 4 PV of polymer are needed to reach the maximum retention.

Fig. 4. 11-Retention isotherm of HPAM in 71 mD core.

0

5

10

15

20

10 100 1000

Po

lym

er r

eten

tio

n,

g/g

san

d

Polymer concentration, ppm

Berea sandstone core 71 mD Porosity: 18.5%

60

Fig. 4. 12-Retention determination for 20 ppm HPAM.

Fig. 4. 13-Retention determination for 40 ppm HPAM.

0

0.2

0.4

0.6

0.8

1

1.2

0 2 4 6 8 10 12 14 16 18 20

Co

nce

ntr

atio

n r

atio

(C

p/C

0)

Pore volumes injected, PV

1st polymer injection

2nd polymer injection

0

0.2

0.4

0.6

0.8

1

1.2

0 2 4 6

Co

nce

ntr

atio

n r

atio

(C

p/C

0)

Pore volumes injected, PV

1st polymer injection

2nd polymer injection

61

The two tests conducted on both high and low permeability sandstone cores indicate

that polymer concentration shows very little effect on retention in porous media if the

same core is repeatedly used. These results are consistent with our static measurements

and also agree with findings by other researchers (Shah et al. 1978, Zheng et al. 1998).

To make a comparison, a similar Dundee sandstone core with 23.4% porosity and 449

mD permeability was used. Polymer solution with concentration of 1,000 ppm was

injected and 56.5 g/g rock retention was measured. Again, for higher concentration,

more retention occurs if fresh core used. In addition, polymer retention shows strongly

concentration-dependent behavior.

Table 4.2-Retention on Sandstone Cores

Core No.

L, cm A, cm2 PV, ml

, % k, mD Polymer Conc, ppm

Retention,

g/g

1 15 14.5 49 22.5 347 20-1,000 16.1

2 15 14.5 51 23.4 449 1,000 56.5

3 12.8 11.4 27 18.5 71 20-1,000 9.77

When static and dynamic retentions were compared, we found that at the same

polymer concentration, static measurement showed much higher retention. For instance,

at concentration of 2,000 ppm, static retention of 335 g/g was detected on disaggregated

sand grains (Fig. 4.3). Nevertheless, dynamic retention was only about 27.8 g/g in 5.51

Darcy sandpack (Table 4.1). The usually high retention from static measurement is

attributed to the large surface area exhibited by disaggregated sand grains, which provide

more vacant sites for polymer molecule attachment. The other dynamic retention by 449

md fresh sandstone core is about 56.5 g/g at concentration of 1,000 ppm (the second

core in Table 4.2). The difference between these two dynamic retentions can be explained

62

by the effect of rock permeability. Research shows that polymer retention strongly

depends on the permeability of porous media. Polymer retention usually increases with

decrease of rock permeability (Vela et al. 1976).

4.2.3 Proposed Adsorption Model

Based on the experimental results, a polymer concentration-related retention model is

proposed, which accounts for the observed retention behaviors. It is well known that

polymer molecules may interact with each other in solution and the degree of interaction

depends greatly on polymer concentration. Three concentration regimes were proposed

(de Gennes 1979; Ying and Chu 1987) as dilute (C < C*), semidilute (C* < C < C**), and

concentrated (C** < C), where C* is the overlap concentration crossover from dilute to

semidilute regimes and C** is the overlap concentration crossover from semidilute to

concentrated regimes. More specifically, as shown in Fig. 4.14, in the dilute regime,

polymer molecules exist in solution as free coils where little interaction occurs. In the

semidilute regime where polymer concentration is greater than the overlap concentration,

C*, macromolecules start to contact each other and intermolecular interactions occur.

With further increase in concentration (especially when the concentration is above C**),

the intermolecular entanglements dominate the interaction, resulting in the formation of a

network structure (Ferry 1948). For our HPAM, brine, and temperature, we measured C*

to be 300 ppm and C** to be 3000 ppm (see the following section). When dealing with

polymer retention on sand surfaces, this concentration-based interaction among polymer

molecules in solution may be used to explain the adsorption mechanism.

63

In the dilute regime, polymer molecules exist in solution as free coils but tend to take

a flat orientation when they adsorb onto the rock surface. In this configuration, most, if

not all of the molecular segments are in contact with the surface. It was called two-

dimensional adsorption (Peterson and Kwei 1961). In this regime, two-dimensional

adsorption dominates retention, and polymer molecules continue to be adsorbed until the

maximum coverage is reached. As shown by Region A in Fig. 4.15, adsorption is

independent of polymer concentration. For practical purposes, the retention in the dilute

regime indicates the minimum amount of polymer needed to occupy the available vacant

sites. In field applications of polymer and chemical floods, reduced polymer retention

may be achieved by first injecting a low-concentration polymer bank.

In the semidilute regime, the intermolecular interaction in solution will result in a

mixed adsorption, i.e., some molecules will be adsorbed with all the segments in contact

with the surface, while others will be adsorbed with only partial segments in contact with

the surface. The latter orientation will be labeled as three-dimensional adsorption.

Increasing the polymer concentration will increase the three-dimensional adsorption as

well as the total adsorption, as shown by Region B in Fig. 4.15. Polymer retention is

concentration-dependent in the semidilute regime.

In the concentrated regime, the molecular entanglement in solution causes the three-

dimensional adsorption to dominate, i.e., most polymer molecules are adsorbed with

segments partially attached to rock surface. Put another way, only one end of the polymer

molecule is attached to the surface, while the majority of the molecule dangles free in the

solution. In this case, almost no additional polymer molecules can be adsorbed with

increasing concentration because all sites have already been occupied. As shown by

64

Region C in Fig. 4.15, the adsorption is concentration-independent. This concentration-

related adsorption model is also summarized in Table. 4.3.

Fig. 4. 14-Polymer molecule interaction at different concentrations.

Fig. 4. 15-Proposed polymer adsorption mechanism on the rock surface.

65

Table 4.3-Summary of Adsorption vs. Polymer Concentration

Polymer

concentration

Molecule interaction

in solution

Dominant adsorption

type on rock surface

Adsorption vs.

concentration

Dilute (Cp<C*) Little Two-dimensional Concentration

independent

Semidilute

(C*<Cp<C**) Intermediate Mixed adsorption Concentration

dependent

Concentrated

(C**<Cp) Entanglement Three dimensional Concentration

independent

This model also explains why no significant adsorption occurred during subsequent

exposure to a high-concentration solution after sand was first contacted with dilute

polymer solution. After the surface maximum coverage is achieved, no more vacant sites

are available.

The above explanation accounts for the dependence of adsorption on polymer

concentration. Nevertheless, besides adsorption on rock surface, mechanical entrapment

occurs simultaneously in pore throat constrictions and dead-end spaces when

consolidated cores are used (Huh et al. 1990, Ranjbar et al. 1991). Hence, the impact of

polymer concentration on mechanical entrapment should also be addressed. Based on the

experimental results depicted in Figs. 4.9 and 4.11, polymer retention that encompasses

both adsorption and entrapment shows concentration-independent behavior. Since

additional adsorption proves to be insignificant when used sands are employed (Fig. 4.4),

the mechanical entrapment, at least the irreversible accumulation of polymer molecules in

small pores and dead-end spaces shows little variation with increase of polymer

concentration. In other words, suppose two identical fresh cores are used. If a high

polymer concentration is injected, this will result in a higher adsorption relative to low

66

concentration injection, but the trapped polymer will remain about the same for the high-

and low-concentration cases.

4.2.4 Overlap Concentration (C* and C**) Measurement

As mentioned, two overlap concentrations exist in polymer solution. One is the overlap

concentration (C*) crossover from the dilute to the semidilute region, and the other is the

overlap concentration (C**) crossover from semidilute to concentrated region. These two

concentrations act as an indicator of molecular interaction in solution. Many approaches

have been proposed to measure C* (Ke et al. 2009). Here, one convenient method,

viscometry, is employed to estimate the overlap concentration for the HPAM polymer we

used.

Method 1: Linearity Deviation. In both low concentration and high concentration

regions, a linear relationship exists between viscosity and concentration. The point

derivates from the linearity should be polymer overlap concentration, C* (Grossman and

Soane 1991). Therefore, by measuring viscosity for polymer solutions from 10 ppm

through 2,000 ppm, Fig. 4.16 was generated which plots viscosity as a function of

concentration. As shown by Fig. 4.16, two line cross point gives the overlap

concentration C*, which is approximately equal to 310 ppm.

Method 2: Intrinsic Viscosity [ ]. Intrinsic viscosity is widely accepted to be correlated

with overlap concentrations:

10**C ……………………………………………………………………………………………………………..(4.1)

67

where n is a constant, and [ ] is the intrinsic viscosity of a polymer in dl/g. Slightly

different values of n have been proposed. For instance, Graessley (1974) came up with n

equal to 1, while, Viovy and Duke (1993) suggested n equaled 0.6. To obtain intrinsic

viscosity, similarly, viscosities of brine as well as polymer solutions with concentration

from 10 ppm to 2,000 were measured. Specific viscosity sp is defined as:

solution solvent

sp

solvent……………………………………………………………………………….(4.2)

where, solution and solvent are viscosities of polymers solution and solvent, respectively.

In Fig. 4.17, sp

C vs. C is plotted, where C is polymer concentration in g/dl. Intrinsic

viscosity equals sp/c when C approaches zero:

0

/lim

sp

c

C

C………………………………………………………………………………………………..…(4.3)

Therefore, intrinsic viscosity can be acquired graphically, i.e., the ratio of sp

Cwhen

C equals to zero or the intercept of the line on the vertical axis gives [ ]=19 dl/g. Hence,

C* = 526 ppm if n=1, or C*=316 ppm if n=0.6. The latter value is very close to that

obtained from the linearity deviation method.

Grassley also suggested C** should be close to 10

**C ; therefore, for the

polymer we used, C** can be approximated to be 3,160 ppm.

68

Fig. 4. 16-Overlap concentration (C*) determination by linearity deviation.

Fig. 4. 17-Overlap concentration (C*) determination by intrinsic viscosity.

0.1

1

10

100

1 10 100 1000 10000

Vis

cosi

ty, c

p

Polymer concentration, ppm

HPAM 6-8 million Daltons Mw 30% hydrolysis 2% NaCl 25 °C

1

10

100

0.001 0.01 0.1

sp/C

(d

l/g)

Polymer concentration, g/dl

HPAM 6-8 million Daltons Mw 30% hydrolysis 2% NaCl 25 °C

69

4.3 Effect of Flow Rate on Polymer Retention

Polymer retention in porous media not only depends on concentration; previous

research also shows that it may be affected by flow rate (Maerker 1973, Dominguez and

Willhite 1976, Aubert and Tirrell 1980, Zaitoun and Kohler 1987, Huh et al 1990). For a

practical application, when polymer solution is injected through an unfractured well (e.g.,

radial flow), the flow rate varies significantly as the polymer penetrates into the reservoir.

In a 20-acre well pattern with formation height of 20 ft, porosity of 20 % and polymer

injected via an openhole injector at a rate of 2,000 bbls/day, Fig. 4.18 demonstrates that

the velocity at which polymer travels through the matrix decreases inversely with the

increase of distance or the invading radius (as shown by Eq. 4.4). For instance, the

velocity (v) is about 67 ft/day at the distance of 1 ft but it decreases sharply to 0.67 ft/day

at the distance of 100 ft. Though the majority of the flow rate in a deep reservoir may be

relatively low, the flow pattern in the near-wellbore area may have tremendous impact on

polymer retention. Therefore, it is of great importance to address this issue.

67.019v

r……………………………...…………………………………...………4.4

where, r is the invading radius of polymer front, ft.

70

Fig. 4. 18-Polymer retention in the near wellbore region, radial flow.

4.3.1 Method Established to Detect Hydrodynamic Retention

Actually, the similar strategy can be applied to the laboratory measurement of

hydrodynamic retention. As shown by Fig. 4.19, if the measurement starts with a fresh

core, then polymer retention at low injection rate is first determined by the same

approach introduced previously, e.g., retention is given by the area between the two

effluent breakout curves. After the completion of retention at a low injection rate, two

high flow rate polymer bank injections are conducted, which are also intercepted with

large volumes of brine injection to flush out the mobile polymers in the porous media.

These two polymer breakout curves are plotted together with the second polymer

breakout curve at a low flow rate as a function of pore volumes injected, (see Fig. 4.19).

There are three curves in Fig. 4.19. To make the explanation simpler, the second breakout

curve at low flow rate is called Curve A, the first and second breakout curves at high flow

y = 67.019x-1 R² = 1

0.1

1

10

100

1 10 100 1000

Vel

oci

ty, f

t/d

ay

Invading radius, ft

20-ac 5-spot spacing rw=4.5 in

h=20 ft, =0.2 QInj=2,000 bbls/day Radial flow

71

rate are called Curve B, and Curve C, respectively. The total retention is given by the

area between the second breakout curve at low flow rate and the first breakout curve at

high flow rate, i.e., area between Curve A and Curve B. In addition to the total retention,

this method is also able to distinguish the irreversible retention from the reversible

retention. The principle behind this approach can be summarized as following:

a) If the area between Curve A and Curve B is zero, it means no incremental

retention is induced by flow rate variation. Theoretically, for a purely

homogeneous core, if no retention occurs with increase of flow rate, these three

breakout curves are supposed to overlap one another. Fig. 4.20 demonstrates the

retention of 40 ppm KI (used as a tracer) in 1.9 Dundee sandstone core. As flow

rate increases from 3.26 ft/day through 104.26 ft/day, almost no retention

difference is noticeable. For this case, the retention of KI in porous media shows

little dependence on flow rate.

b) If the area between Curve A and Curve B is not zero, but Curve B and Curve C

overlap each other or the area between these two curves is zero, this means there

is no incremental irreversible retention. In other words, all the retention can be

considered reversible.

c) If the area between Curve A and Curve B is not zero, but Curve B and Curve A

overlap each other or the area between Curve B and Curve A is zero, this means

there is no incremental reversible retention. In other words, all the retention can

be considered irreversible.

d) If the area between Curve A and Curve B is not zero and Curve C falls in the

middle of the other two curves (Curves A and B), then the area between Curves A

72

and C gives the incremental reversible retention and the area between Curves C

and B gives the incremental irreversible retention. Of course, the total incremental

retention is the addition of these above two.

Note: this method is valid to estimate flow rate induced retention based on the

assumption that polymer IAPV is independent of flow rate. If at high injection rate or

pressure gradient, polymer solution may penetrate into the unswept region which is

already occupied by brine (not polymer), then polymer IAPV will decreases with increase

of flow rate. If this is true, the phenomenon of IAPV decrease will also delay polymer

propagation somewhat and a scenario similar to that shown in Fig. 4.19 will be observed.

This question will be addressed in a later section.

Fig. 4. 19-Mehod to determine hydrodynamic retention.

73

Fig. 4. 20-Effect of flow rate on KI (tracer) retention.

4.3.2 Hydrodynamic Retention in 1.9 Darcy Dundee Sandstone Core

500 ppm HPAM Polymer. A Dundee sandstone core with relatively high permeability

of 1.9 darcies and porosity of 24.1% was used to estimate how flow rate affected polymer

retention (Table 4.4). As mentioned earlier, retention at low flow rate of 3.26 ft/day was

first measured, which was about 20.35 g/g rock. After that, the same 500 ppm HPAM

polymer solution was injected into thehigh permeability core at elevated injection rates

ranging from 6.52 ft/day to 52.14 ft/day. Again, 100 PV of brine were injected after each

polymer injection. The results are shown in Fig. 4.21 and Table 4.5. Because the area

between the second breakout curve at 3.26 ft/day and the increased flow rates is not zero,

the total retentions as a result of flow rate increase were estimated to be 2.69, 5.68, 7.97,

8.67, and 9.10 g/g rock as flow rates increases from 3.26 ft/day to 6.52, 13.03, 26.07,

52.13, and 104.26 ft/day, respectively.

0

0.2

0.4

0.6

0.8

1

1.2

0 1 2 3 4

Co

nce

ntr

atio

n r

atio

(C

p/C

0)

Pore volumes injected

Tracer at 3.26 ft/day

Trace at 6.52 ft/day

Tracer at 13.03 ft/day

Tracer at 26.07 ft/day

Tracer at 104.26 ft/day

1.9 D sandstone core 40 ppm KI tracer 2% NaCl 25 °C

74

Table 4.4-Properties of the Dundee Sandstone Core.

L, cm A, cm2 PV, cm3 ,% k, D Retention @ 500 ppm, 3.26

ft/day

15 14.5 52.4 24.1 1.9 20.35

Table 4.5-Retention Summary.

Q, ft/day Total, g/g rock Incremental Irrev,

g/g

Incremental Rev, g/g rock

RRF

3.26 20.35 0 1.89

6.52 23.03 0 2.69 1.87

13.03 25.72 -0.31 5.68 1.92

26.07 28.52 0.20 7.97 1.86

52.13 28.45 -0.57 8.67 1.89

104.26 29.50 0.061 9.10 1.90

Notes: Total, Irrev, and Rev, refer to total, irreversible and reversible retention, respectively.

Fig. 4. 21-Polymer retention at flow rates from 3.26 ft/day to 52.16 ft/day.

0

0.2

0.4

0.6

0.8

1

1.2

0 1 2 3 4

Co

nce

ntr

atio

n r

atio

(C

p/C

0)

Pore volumes injected

3.26 ft/day

6.52 ft/day

13.04 ft/day

52.16 ft/day

75

Retention vs. Flow Rate. We plotted incremental retention (percentage) as a function of

flux in Fig 4.22. Careful examination of Fig. 4.22 reveals that different trends are

observed in different flow regions. For instance, the curve slope of low region from 6.52

ft/day to 26.07 ft/day is approximately 19 times greater than that from 26.07 ft/day to

104.26 ft/day. This implies the retention increases drastically in the low region and when

the flow rate reaches a certain value, the retention rises much more gradually with the

increase of flow rate.

Fig. 4. 22-Incremental retention of HPAM vs. flux.

Reversibility of Hydrodynamic Retention. For each case, a second polymer bank was

injected after 100 PV of brine injection with intervening periods of no flow to check the

reversibility of this flow rate-related retention. Fig. 4.23 and Fig. 4.24 give the results for

13.03 ft/day and 52.16 ft/day injection, respectively. The area between the first and

10

100

1 10 100 1000

Incr

emen

tal r

eten

tio

n, %

Flux, ft/day

76

second breakout curves at same flow rate for both cases is close to zero. Based on the

principle explained earlier, it can be inferred that no incremental irreversible retention is

detected. In other words, all the incremental retention turns out to be reversible. These

results confirm that 100 PV brine injection in our study is sufficient to displace these

mobile polymers in the core. Otherwise, a non-zero area between these two breakout

curves should be recognized. They also prove that the same core can be repeatedly used

to measure retentions at other flow rates as long as sufficient brine injection is guaranteed.

Impact of Hydrodynamic Retention on Residual Resistance Factor. How does

hydrodynamic retention affect permeability? It can be predicted that, if all the flow-

induced retention is reversible as indicated by Figs. 4.23 and Fig. 4.24, there should be no

permeability reduction caused by this hydrodynamic retention. This postulation is

confirmed by residual resistance factor measurement after core exposure to 100 PV brine

injection for each case. As shown in Table 4.5, the residual resistance factors prove to be

quite stable for all these cases with values close to 1.9. Little dependence of residual

resistance factor on polymer injection rate also verifies that almost all the incremental

retention associated with flow rate increases can be flushed out of the core during the

brine postflush. These results agree with findings by Maerker (1973), Dominguez and

Willhite (1976). However, our results display a much better picture of retention

reversibility.

77

Fig. 4. 23-Determination of irreversible retention at 13.04 ft/day.

Fig. 4. 24-Determination of irreversible retention at 52.16 ft/day.

0

0.2

0.4

0.6

0.8

1

1.2

0 1 2 3 4

Co

nce

ntr

atio

n r

atio

(C

p/C

0)

Pore volumes injected

1st polymer injection

2nd polymer injection

HPAM 6-8 million Daltons Mw 30% hydrolysis 500 ppm 2% NaCl

1.9 D Dundee sandstone core

0

0.2

0.4

0.6

0.8

1

1.2

0 1 2 3 4

Co

nce

ntr

atio

n r

atio

(C

p/C

0)

Pore volumes injected

1st polymer injection

2nd polymer injection

HPAM 6-8 million Daltons Mw 30% hydrolysis 500 ppm 2% NaCl

1.9 D Dundee sandstone core

78

HPAM Polymer Rheology in Porous Media. Polymer resistance factor is defined as

water mobility (before exposure to polymer) divided by polymer solution mobility. The

resistance factor is a measure of the effective viscosity of a polymer solution in a porous

media, relative to water. Pressure drops across the core during both polymer and brine

injection were recorded and used to calculate resistance factor for each case. As shown

by Fig. 4.25, except for the extremely high flow rate, HPAM polymer shows either

Newtonian or shear thickening behaviors. For instance, at low to moderate injection rates

from 3 ft/day to 10 ft/day, HPAM shows Newtonian behavior, i.e., apparent viscosity is

independent of flow rate or shear rate. While, with further increase of flow rates from 10

ft/day to 50 ft/day, it becomes shear thickening, i.e., apparent viscosity increases with

flow rate or shear rate. If flow rate keeps increasing, a decreased resistance factor may be

observed. This is attributed to the viscosity reduction caused by mechanical degradation

at high flux (Maerker 1975, Seright 1983). For a practical perspective, this unusually high

flow rate or shear rate is only anticipated to exist in the near-wellbore region from where

polymer solution leaves the wellbore and first enters the formation.

The rheology of HPAM in porous, especially shear thickening behavior, was

attributed to the viscoelastic character of HPAM and elongational flow field in porous

media (Durst et al. 1982; Macosko 1993). The rheology of HPAM in porous media can

be summarized as the following: In a typical brine solvent, polymer molecules usually

take a coiled configuration and significant energy needed to untangle this coil (i.e.,

expand and elongate this molecule). At low flow rate, polymer molecules have sufficient

time to react to the hydrodynamic force acting upon them and become somewhat

stretched. The main principle here is that a stretched molecule always shows less

79

resistance to flow than a coiled one. Therefore, the resistance factor exhibited in porous

media at low flow rate or shear rate is consistent with the viscosity measured in a

viscometer. However, at high flow rate or shear rate, polymer molecules do not have

sufficient time to become stretched before they flow through a constricted pore structure.

Hence, more energy (or pressure gradient) needed to force them through these small pore

throats, which results in a high resistance factor at increased flow rate or shear rate. This

is the reason why shear thinning behavior of HPAM exhibited in a viscometer will not be

observed in a porous media. In contrast, double-helix, rod-like xanthan polymer

molecules are already expanded even at low flow rate or shear rate. When they flow

through constricted pore throat, regardless of flow rate, no significant coil-to-stretched

transition occurs. Therefore, unlike HPAM polymer, xanthan always show shear thinning

behavior in both the viscometer and porous media.

Fig. 4. 25-Resistance factor of HPAM at different flow rates.

0

2

4

6

8

10

1 10 100 1000

Res

ista

nce

fac

tor

(RF)

Flux, ft/day

HPAM 6-8 million Daltons Mw 30% hydrolysis 500 ppm 2% NaCl 25 °C

1.9 D Dundee sandstone core

80

4.3.3 Is HPAM Shear Thickening Behavior Caused by Hydrodynamic Retention?

As polymer retention in porous media increases with increase of flow rate or shear rate,

simultaneously, the resistance factor of HPAM solution in porous media also increases.

The question can be raised: is this shear thickening behavior a result of retention-related

permeability reduction instead of elongational flow? To address this question, another

water-soluble EOR polymer, xanthan was tested by following the same procedures.

Before the retention measurement, both HPAM and xanthan rheology was measured in a

viscosity (Anton Paar rheometer). Figure 4.26 reveals that HPAM and xanthan show

similar rheology in the viscometer except for the low concentration HAPM (150 ppm),

which behaves more like a Newtonian fluid, i.e., viscosity depends little on shear rate.

150 ppm and 500 ppm xanthan solutions show obvious shear thinning behavior, while

500 ppm HPAM displays a mild shear thinning property. Figure 4.26 also demonstrates

that no shear thickening behavior of HPAM, which is displayed in porous media can be

observed in a viscometer.

81

Fig. 4. 26-Rheology of HPAM and xanthan polymers in a viscometer.

Hydrodynamic Retention of Xanthan in 1.9 Darcy Core. The same method was tried

to estimate how xanthan retention differs with flow rate. Unfortunately, no consistent

results could be attained. Analysis shows that pressure drops across the 10 micron filter

combinations were much less sensitive to effluent concentration, compared to HPAM

solution. To minimize experimental error, we switched to another approach for effluent

concentration determination. Effluent samples were collected periodically for total

organic carbon (TOC) content determination, as we did for static adsorption measurement.

Figure 4.27 plots two normalized concentration ratios as a function of pore volumes

injected. One is the breakout curve at low flow rate of 3.26 ft/day, and the other is from

polymer injection at a high flow rate of 26.07 ft/day. The area between these two curves

gives the incremental retention of only 0.81 g/g rock, which is much lower than that

from HPAM (7.97 g/g rock). This difference in retention reveals that different

1

10

100

1 10 100 1000 10000

Vis

cosi

ty, c

p

Shear rate, 1/s

500 ppm Xanthan

500 ppm HPAM

150 ppm HPAM

150 ppm xanthan

82

functional groups or molecular size among these two polymers may significantly affect

their retention behaviors in the same porous media. Since no appreciable amount of

xanthan is retained in this quite permeable core, a lower permeability core probably is

needed to study xanthan retention.

Fig. 4. 27-Hydrodynamic retention of 150 ppm xanthan in 1.9 Darcy core.

Rheology of Xanthan in Porous Media. Resistance factors of xanthan polymer were

also measured. As shown by Fig. 4.28 and 4.26, xanthan consistently shows shear

thinning behavior in both porous media and in a viscometer. For instance, resistance

factor is 8.5 at flux of 3.26 ft/day. It declines to 5.5 when flow rate rises to 52.13 ft/day.

These results agree with other findings (Seright et al. 2009, 2010).

0

0.2

0.4

0.6

0.8

1

1.2

0 1 2 3 4 5

Co

nce

ntr

atio

n r

atio

(C

p/C

0)

Pore volumes injected

Cp/Co at 3.26 ft/day

Cp/Co at 26.07 ft/day

Xanthan

500 ppm 2% NaCl, 60 ml/hr 25 °C

1.9 D Dundee sandstone core

83

Fig. 4. 28-Resistance factor of xanthan at different flow rates.

4.3.4 Hydrodynamic Retention in 71 mD Berea Sandstone Core

As discussed above, a relatively high permeability core may not be a good candidate to

capture the retention variation with flow rate for a xanthan polymer. A 71 md Berea

sandstone core was selected whose retention has been satisfied with large pore volumes

of HPAM polymer injection. Experiments were carefully designed to attack two kinds of

problems. 1) Is rheology an intrinsic property of a polymer solution itself, or it is closely

related to polymer retention in porous media? 2) Does polymer IAPV decrease with

increase of flow rate or pressure gradient? This second question has been mentioned

previously.

Experiment Procedures. To answer these two questions, the following procedures were

designed:

1

10

1 10 100

Res

ista

nce

fac

tor

(RF)

Flux, ft/day

Xanthan 500 ppm 2% NaCl 25 °C

1.9 D Dundee sandstone core

84

a) Inject 100 PV of brine with intervening periods of no flow to flush out all the

mobile polymer molecules in the porous media;

b) Inject polymer solution at a low flow rate (Qa) till the effluent concentration is the

same as injected. Note: from this point, during polymer injection in each step,

record pressure drop across the core and collect effluent samples periodically for

concentration determination via TOC analyzer;

c) Without any pause, raise flow rate abruptly to a level several times higher than the

previous one and continually inject polymer solution at this flow rate (Qb) until

several pore volumes have been injected;

d) Either pause for a while or without pause, reduce flow rate back to the lower level

(Qa) and continue polymer injection until several pore volumes have been injected;

e) Raise flow rate again from Qa to Qb and complete several pore volumes of

solution injection;

f) Plot normalized polymer concentration and polymer resistance factor vs. pore

volumes injected.

Retention and Rheology of 150 ppm HPAM in 71 mD Core. Injection of 150 ppm

HPAM was analyzed first. Figure 4.29 indicates that when a sudden increase of flow rate

from 4.14 ft/day to 16.57 ft/day was made after around 4 PV injection, an immediate

effluent polymer concentration decrease was detected. It indicates that more polymers are

stripped from the solution due to hydrodynamic force increase. This phenomenon agrees

with the finding mentioned previously in a 1.9 darcy core. After around 4 PV of injection,

effluent polymer concentration approached the initial polymer concentration. Next, the

abrupt flow rate decrease from 16.57 ft/day back to 4.14 was accompanied with a rise of

85

effluent concentration. Again, this demonstrates that the incremental retention caused by

applying more hydrodynamic force on polymer molecules will be released back to the

solution when flow rate is lowered. Finally, when flow rate once again increases to 16.57

ft/day, an approximately similar amount of polymer is retained by the porous media.

These results confirm our former finding that almost all hydrodynamic retention is

reversible. Comparing the three areas marked A, B and C in Fig. 4.29, within

experimental error, they are nearly the same, which means that almost all the retention is

reversible. Another significance of this test suggests that polymer IAPV varies little with

increase of flow rate. Otherwise, we would not expect to see these effluent concentration

variations with flow rate change.

Rheology of 150 HPAM in this 71 md core was continuously monitored during the

whole injection period. As shown by Fig. 4.29, the resistance factor is almost the same at

both injection rates of 4.14 ft/day and 16.57 ft/day with a value around 12.5 except at the

time when flow rate adjustment was just made. This Newtonian-like rheology is also

displayed in a viscometer shown in Fig. 4.26 (the solid triangle). Actually, a similar

phenomenon was observed from the previous tests for HPAM in 1.9 darcy core. When

flow rate increased from 3.26 ft/day to 6.52 ft/day and 13.03 ft/day, 2.56 g/g rock and

5.68 g/g rock incremental retention was detected (Table 4.5), respectively, but all the

resistance factors stabilized at 3.8 (Fig. 4.25) and no resistance factor increase was

observed with retention increase. Therefore, the rheology of polymer in porous media

should be its intrinsic property. Otherwise, the increase in hydrodynamic retention will

have a significant impact on permeability reduction, and as a consequence, the resistance

factor at elevated flow rate would be much higher than that from low injection rate.

86

Fig. 4. 29-Hydrodynamic retention and resistance factor for HPAM.

Retention and Rheology of 150 ppm Xanthan in 71 mD Core. Similar tests were

performed to 150 ppm xanthan polymer. The results are displayed in Fig. 4.30. Again, as

flow rate rises from 4.14 ft/day to 33.16 ft/day, more retention in porous media is

observed, i.e., the concentration ratio (Cp/C0) is less than 1. After several pore volumes of

polymer injection, the system was shut down for 2 hr to eliminate any hydrodynamic

force acting on the molecules. This allowed the deformed molecules to have sufficient

time to relax and take random-coil configurations. When flow was resumed, an effluent

concentration higher than the initial concentration was detected (Cp/C0>1). The highest

concentration ratio (Cp/C0) after the flow resumed was as high as 1.9.

1

10

100

1000

0

0.2

0.4

0.6

0.8

1

1.2

1.4

0 5 10 15

Res

ista

nce

fac

tor

Co

nce

ntr

atio

n r

atio

(C

p/C

0)

Pore volume injected

4.14 ft/day 16.58 ft/day 4.14 ft/day 16.58 ft/day

A

B C

87

Again, when we checked the polymer rheology in porous media, we found xanthan

exhibits shear thinning behavior. For example, the resistance factor at low flow rate of

4.14 ft/day is 8.04. In contrast, the resistance factor decreases to 5.9 at high flow rate of

33.16 ft/day (see the solid triangle in Fig. 4.30). This rheology is consistent with that

from a viscometer (Fig. 4.26).

Fig. 4. 30-Hydrodynamic retention and resistance factor for xanthan.

If the incremental retention does not significantly impact polymer rheology in porous

media by reducing rock permeability, then, one may ask, where do these newly-retained

polymers go? Based on the experimental results, we postulate that they are either trapped

in the part of the porous media that contribute little to rock permeability before polymer

1

10

0

0.5

1

1.5

2

2.5

3

0 2 4 6 8 10 12

Res

ista

nce

fac

tor

Co

nce

ntr

atio

n r

atio

(C

p/C

0)

Pore volumes injected

4.14 ft/day 33.16 ft/day 4.14 ft/day

88

injection (such as dead end pores or crevices), or in a structure whose permeability is not

altered much with the increased accumulation of polymer molecules.

4.4 Polymer Inaccessible Pore Volume (IAPV)

Several mechanisms can be envisioned to IAPV for polymers (Liauh et al. 1979, van

Domselaar and Fortmuller 1992). First, pores or parts of pore spaces may be large enough

to accommodate small molecules (such as solvent, salts or tracers) but too small to allow

entry of polymer molecules (Dawson and Lantz 1972). The presence of clay may also

contribute IAPV, if small molecules can freely penetrate the clay but large polymer

molecules cannot.

Fig. 4. 31-IAPV determination for 100 ppm HPAM, 60 ml/hr, 347 mD core.

0

0.2

0.4

0.6

0.8

1

1.2

0 1 2 3 4 5

Co

nce

ntr

atio

n r

atio

(C

p/C

0)

Pore volumes injected, PV

2nd polymer injection

2nd tracer injection

Dundee sandstone core Porosity: 23% 347 mD

89

In our method, Fig. 4.31 is a typical curve for calculating IAPV. It shows both

effluent concentration breakout curves of the second HPAM slug and the tracer slug. The

first polymer injection is supposed to satisfy polymer retention in the porous rock so no

irreversible polymer retention occurs during the second polymer slug injection. Therefore,

if the reversible retention is negligible, due to the presence of IAPV, the second polymer

front should travel faster than tracer front. This is the principal of IAPVestimation in our

tests.

Of course, if there is any reversible retention, these movable polymer molecules will

be flushed out during the following large pore volumes of brine injection, and then this

part of retention during the second polymer bank injection will reoccur. If this reversible

retention is significant, the two concentration breakout curves may be very close to each

other or their positions may even be switched, resulting in a very small or even a negative

IAPV. This is especially true with the increase of polymer injection rate and decrease of

core permeability, where reversible retention becomes severe. As shown by Fig. 4.32,

where 1,000 ppm polymer is injected into 347 mD Dundee core at a flow rate of 1,000

ml/hr, the second polymer concentration curve is below that of the tracer, which gives a

negative inaccessible pore volume. The same phenomenon was found by Vela et at

(1976). Nevertheless, if there is any reversible retention occurring, these curves could be

used as a very good indicator of the reversibility of polymer retention. This will be

discussed in the following section.

90

Fig. 4. 32-IAPV determination for 20 ppm at flow rate of 1,000 ml/hr.

Inaccessible pore volume calculated from Fig. 4.33 is about 0.22 PV. For this case, 20

ppm HPAM polymer was injected into 347 mD Dundee sandstone core at the flow rate of

60 ml/hr (around 3 ft/day). When 100 ppm HPAM solution was injected into the same

core at the same flow rate, IAPV determined as shown in Fig. 4.33 decreased to 0.15 PV.

Shah et al (1978) attributed this phenomenon to the decrease of hydrodynamic

dimensions of polymer macromolecules with increase of polymer concentration. They

postulated that the decrease of polymer dimensions makes it possible for some polymer

molecules to penetrate into some small pores that might not be contacted at low

concentrations. As a result, a smaller IAPV could be formed. However, further

examination is needed to determine how much this theory can be relied on. Instead, using

reversible retention theory seems more plausible, because at the same flow rate, the

increase of polymer concentration will lead to an increase in hydrodynamic force, as a

0

0.2

0.4

0.6

0.8

1

1.2

0 2 4 6

Co

nce

ntr

atio

n r

atio

(C

p/C

0)

Pore volumes injected, PV

2nd polymer injection

2nd tracer injection

Dundee sandstone core Porosity: 23% 347 mD

91

consequence, the reversible hydrodynamic retention increases. As explained earlier, this

will lead to a decreased inaccessible pore volume.

Fig. 4. 33-IAPV determination for 20 ppm HPAM at flow rate of 60 ml/hr.

Although IAPV was estimated in our tests, we should be aware of the experimental

errors and limitations responsible for unrealistic IAPV values. Sorbie (1991) pointed that

the diameter of an EOR HPAM in a 3% NaCl brine is typically 0.5—0.8 m. Also, an

XMT analysis of 470-mD Berea sandstone revealed pores that were highly connected and

that 98% of the pores had an effective diameter greater than 26 m and a pore throat

diameter over 6.7 m (Seright et al. 2006). Thus, a typical EOR HPAM molecule in

solution is small compared to the pore sizes and throat sizes, so it should have access to

most spaces in the porous media. It seems likely that experimental errors and limitations

are responsible for the observed IAPV variations. Therefore, more research needs to be

done to clarify this discrepancy before reaching any decisive conclusions about polymer

IAPV.

0

0.2

0.4

0.6

0.8

1

1.2

0 2 4 6

Co

nce

ntr

atio

n r

atio

(C

p/C

0)

Pore volumes injected, PV

2nd polymer injection

2nd tracer injection

Dundee sandstone core Porosity: 23% 347 mD

92

Fig. 4. 34-2nd injection of 160 ppm HPAM and tracer at 60 ml/hr, 71 mD core.

Fig. 4. 35-2nd injection of 1,000 ppm HPAM and tracer at 240 ml/hr, 71 mD core.

0

0.2

0.4

0.6

0.8

1

1.2

0 2 4 6

Co

nce

ntr

atio

n r

atio

(C

p/C

0)

Pore volumes injected, PV

2nd polymer injection

2nd tracer injection

Barea sandston core Porosity: 18.5% 71 mD

0

0.2

0.4

0.6

0.8

1

1.2

0 1 2 3 4 5

Co

nce

ntr

atio

n r

atio

(C

p/C

0)

Pore volumes injected

2nd polymer injection

2nd tracer injection

Barea sandston core Porosity: 18.5% 71 mD

93

4.5 Effect of Polymer Retention on Permeability Reduction

Literature shows that polymer retention reduces the permeability or increases the residual

resistance factor of porous rock. This phenomenon becomes more severe when rock

permeability decreases (Vela et al 1976; Seright 1992; Seright and Martin 1993).

Residual resistance factors (RRF) are determined by conducting brine injection before

and after polymer retention until the pressure drop across the core stabilizes. Equation 4.4

is used to calculate the residual resistance factor.

( )

( )

A

B

PRRF

P………………………………………………………………………..….4.4

where, ( P)A ( P)B are pressure drops for brine injection after and before polymer

retention, respectively. Of course, the same brine injection rate should be maintained.

Figure 4.36 plots the relationship of the residual resistance factor vs. polymer

concentration for a high permeability core (347 mD Dundee core). As illustrated, the

residual resistance factor is only about 1.4 and it varies little with increase of polymer

concentration. Based on previous findings (Rowland 1963; Rowland and Eirich 1966),

the retention of polymer will form a monolayer on the solid surface, reducing the

effective pore size. As a consequence, rock permeability decreases. The constant residual

resistance factor with concentrations from 20 ppm to 1,000 ppm correlates well with the

amount of polymer retention, as indicated by Fig. 4.9, which is also independent of

polymer concentration.

94

Fig. 4. 36-Effect of polymer concentration on residual resistance factor, 347 mD core.

Fig. 4. 37-Effect of polymer concentration on residual resistance factor, 71 mD core.

0

1

2

3

4

10 100 1000

Res

idu

al R

esis

tan

ce F

acto

r (R

RF)

Polymer concentration, ppm

Dundee sandstone core 347 mD Porosity: 23%

0

2

4

6

8

10

10 100 1000

Res

idu

al R

esis

tan

ce F

acto

r (R

RF)

Polymer concentration, ppm

HPAM 6-8 million Daltons MW 30% hydrolysis 2.52% TDS 60 ml/hr 25 C

Barea sandston core 71 mD Porosity: 18.5%

95

For low permeability core (71 mD Berea core), the residual resistance factor is much

higher compared with that from high permeability core. As shown by Fig. 4.37, the value

of the residual resistance factor is in the range of 4.5—7.2, which is also consistent with

the result of polymer retention shown in Fig. 4.11.

4.6 Steady-State Flow in Porous Media

Literature (Zitha et al 2001) shows that an unsteady-state (continuous build-up of

polymer injection pressure) may be reached when polymer is flowing through porous

media if these conditions are satisfied: (a) the velocity gradient is large enough to induce

a coil-stretch transition, (b) the polymer adsorbs on the pore surfaces and (c) the length of

stretched macromolecules is larger than the effective pore throat diameter. During our

tests, we recorded the pressure drop for both high permeability (347 mD) and low

permeability core (71 mD). As shown in Figs. 4.38 and 4.39, regardless of the flow rate

and permeability, all the pressure drops across the core stabilized at around 1 PV polymer

injection. For these cases, no brine injection was conducted after polymer injection. This

means a steady-state flow was reached and no continual pressure build-up was observed.

In other words, no significant pore bridging phenomenon took place in our tests

regardless of rock permeability.

96

Fig. 4. 38-Pressure drop during polymer injection in 347 mD core.

Fig. 4. 39-Pressure drop during polymer injection in 71 mD core.

1

10

100

0 2 4 6

Pre

ssu

re d

rop

, psi

Pore volumes injected, PV

60 ml/hr

180 ml/hr

1000 ml/hr

1

10

100

1000

0 2 4

Pre

ssu

re d

rop

, psi

Pore volumes injected, PV

60 ml/hr

120 ml/hr

600 ml/hr

97

The increase of pressure drop with increased flow rate shown by these two figures

can be attributed to HPAM shear thickening rheology in porous media; i.e., the HPAM

resistance factor (effective polymer viscosity relative to brine viscosity) increases with

the flow rate (Jennings et al. 1971, Southwick and Manke 1988, Seright et al. 2010). For

instance, the pressure drop at injection rate of 60 ml/hr for 347 mD core was about 2 psi;

by comparison, the value rose to around 10 psi at 180 ml/hr. This trend was also true for

the 71 mD core where pressure drop was about 60 psi and 200 psi at flow rate of 60 ml/hr

and 120 ml/hr, respectively.

98

CHAPTER 5. CONCLUSIONS

5.1 Conclusions

This study focuses on EOR polymer retention in porous media. Factors such as polymer

concentration and flow rate were specially investigated. Retention reversibility and the

impact of retention on rock permeability reduction and flow rheology were also estimated.

The following conclusions were drawn from our experiments:

1. The adsorption of HPAM polymers on rock surface can be considered

instantaneous and irreversible. These results indicate that these long chain

molecules have high adsorption tendency. It is impossible for the adsorbed

molecules to lose all of their segment contact with rock surface at the same time.

2. Different polymer retention behaviors are observed in dilute, semi-dilute and

concentrated regions in the absence of oil saturation. In both dilute and

concentrated regions, polymer retention is basically concentration-independent. In

contrast, in the semi-dilute region, polymer retention is concentration-dependent.

3. If a porous medium is first contacted with dilute HPAM solution to satisfy the

retention, no significant additional retention occurs when exposed to higher

concentrations. In field applications of polymer and chemical floods, reduced

99

polymer retention may be achieved by first injecting a low-concentration polymer

bank.

4. Based on the experimental results, a concentration-related retention mechanism is

proposed that considers the orientation of the adsorbed polymer molecules and the

interaction between molecular coils in solution. More specifically, in dilute

regions, where there is little or no interaction between polymer molecules, the

polymer molecules will be adsorbed individually and take a flat orientation on the

rock surface with full segment attachment. For this case, two-dimensional

adsorption dominates and retention shows concentration-independence. In the

concentrated region, strong interaction between molecules exists in the solution

and the dominant adsorption will be three-dimensional which means almost no

molecules are in full segment contact. The retention in this concentrated region

also shows concentration-independent. In contrast, in the semi-dilute region, the

intermediate degree of molecular interaction leads to mixed retention, and three-

dimensional retention increases with the increase of polymer concentration,

therefore, polymer retention shows concentration dependent and increases with

concentration.

5. Polymer retention proves to be affected by flow rate. Retention of 500 ppm

HPAM in 1.9 sandstone core increases by 13.2% and 39.2% as flow rate rises

from 3.26 ft/day to 6.52 ft/day and 52.13 ft/day, respectively. Hydrodynamic

retention shows strong flow dependence. In low flow regions, the retention

increases abruptly. By comparison, in high flow regions, the increase becomes

much more gradual. Almost all the hydrodynamic retention proves to be

100

reversible. In other words, incremental irreversible retention associated with flow

rate increase is negligible.

6. In porous media, HPAM polymers exhibit Newtonian behavior at low to moderate

flow rate and shear thickening behavior and high flow rate, while, xanthan

polymers show shear thinning behavior. Results show rheology of EOR polymers

exhibited in porous media should be an intrinsic property. The hydrodynamic

retention has limited effect on flow behaviors.

7. Negative polymer inaccessible pore volume (IAPV) is observed with increase of

flow rate and decrease of permeability. This presence of negative IAPV is caused

by reversible polymer retention. Adsorption on rock surface was proved to be

almost irreversible, therefore, the mechanical entrapment and hydrodynamic

retention account for this phenomenon.

8. The residual resistance factors (RRF) determined after the polymer injection

increases with decrease of rock permeability as a result of polymer retention.

Residual resistance factors of 1.4 and over 4 were detected for high permeability

core (347 mD) and low permeability core (71 mD), respectively. Nevertheless, the

reversible hydrodynamic retention does not lead to permeability reduction.

5.2 Discussions and Future Work

This research focused on the effect of polymer concentration on retention and

measurements were carried out in monophasic conditions (in the absence of oil phase).

Previous work shows polymer retention is usually lower with residual oil saturation than

that without oil present—suggesting that the presence of residual oil may reduce HPAM

somewhat (Szabo 1975, Hughes et al. 1990, Broseta et al. 1995). Polymer retention

101

causes rock permeability reduction by forming a thin layer on rock surface (residual

resistance factor). The RRF increases with decrease of permeability (Jennings et al. 1971,

Hirasaki and Pope 1974, Seright 1993). In the future, the dependence of permeability

reduction and oil saturation on polymer concentration should be investigated further,

especially with regard to the concentration-dependence of polymer retention that was

reported in this paper.

As mentioned earlier, field applications may be interested in our finding that polymer

retention is low if the rock is first contacted with a low polymer concentration. Polymer

losses due to retention may be reduced considerably if a dilute polymer bank precedes the

main mobility-control bank. Whether or not this approach is practical will depend on

economics, timing, and the specific polymer retention levels that occur in the particular

field application.

Should our model and results replace the existing Langmuir isotherm formulations for

polymer retention in simulators? The Langmuir isotherm was always incorrect

mechanistically for polymer retention (because most polymer retention is irreversible), so

one could argue that it should never have been used. However, to be more supportive of

previous simulators, using the Langmuir isotherm may not result in grossly incorrect

predictions if the Langmuir plateau is set to be reached at a very low polymer

concentration, if the polymer front is sufficiently sharp, and if the injected polymer

concentration is relatively high. If these conditions are not met, there should be value in

incorporating our model into the polymer flooding simulator.

102

Accurate estimation of polymer IAPV should be addressed in the future. Controversial

results on IAPV are shown in petroleum literature. To better understand polymer

propagation through reservoirs, in addition to polymer retention, the magnitude of the

IAPV for a specific kind of reservoir should also be investigated.

103

NOMENCLATURE

A = core section area, cm2

a1 = constant in Langmuir isotherm adsorption equation

b1 = constant in Langmuir isotherm adsorption equation

C = solute concentration in Langmuir isotherm adsorption equation

C* = polymer overlap concentration crossover from dilute to semidilute regime

C** = polymer overlap concentration crossover from semidilute to concentrated

regimes

C0 = initial polymer concentration, ppm

Ceq = equilibrium polymer concentration, ppm

Cp = polymer concentration in effluent, ppm

Cpoly = polymer concentration, ppm

eH = effective hydrodynamic thickness of the adsorbed layer, m

h = formation height, ft

k = permeability, mD or D

kro = relative permeability to oil

krw = relative permeability to water

L = core lenth, cm

M = mobility ratio

Pinj = mass of polymer injected, lbm

104

Pre = mass of polymer retained in the reservoir, lbm

PVret = pore volume delay per pore volume injected

Qa, Qb = polymer injection rates

QInj = polymer injection rate, bbls/day

Rpret = polymer retention, g/g sand

r = invading radius, ft

rw = wellbore radius, in

Sor = residual oil saturation

Sw = water saturation

Swr = residual water saturation

v = velocity, ft/day.

Vp = volume of polymer solution, cm3

Wsg = weight of sand, g

( P)A = pressure drops for brine injection after polymer retention, psi

( P)B = pressure drops for brine injection before polymer retention, psi

[ ] = intrinsic viscosity, dl/g

p = pressure drop, psi

= solute adsorption in Langmuir isotherm

= porosity

= shear rate, s-1

c = critical shear rate, s-1

solution = solution viscosity, cp

solvent = solvent viscosity, cp

sp = specific viscosity

105

d = mobility of the displaced phase, mD/mPa·s

D = mobility of the displacing phase, mD/mPa·s

o = viscosity of crude oil, cp

p = viscosity of polymer solution, cp

w = viscosity of water, cp

g = density of sand grains, g/cm3

rock = rock density, g/cm3

Acronyms

BET = Brunauer, Emmett, and Teller

CAT = cationic polyacrylamide

EHT = effective hydrodynamic thickness

EOR = enhanced oil recovery

HPAM = partially hydrolyzed polyacrylamide

IAPV = inaccessible pore volume

Mw = molecular weight

OOIP = original oil in place

PAM = polyacrylamide

PV = pore volume

RPM = revolution per minute

RRF = residual resistance factor

SP = sandpack

TDS = total dissolved solids

TOC = total organic carbon

106

REFERENCES

Aubert, J.H. and Tirrell M. 1980. Flows of Dilute Polymer Solutions through Packed

Porous Chromatographic Columns, Rheologica Acta, 19 (4): 452–461.

Broseta, D., Medjahed, F., Lecourtier, J., and Robin, M. 1995. Polymer

Adsorption/Retention in Porous Media: Effects of Core Wettability and Residual Oil.

SPE Advanced Technology Series, 3 (1), 103-112.

Camilleri, D., Engelsen, S., Lake, L.W, et al. 1987. Description of an Improved

Compositional Micellar/Polymer Simulator, SPE Res Eng, 2 (4): 427–432.

Castagno, R.E., Shupe, R.D., Gregory, M.D., and Lescarboura, J.A. 1987. Method for

Laboratory and Field Evaluation of a Proposed Polymer Flood, SPE Res Eng, 2 (4):

452–460.

Chauvetear, G. and Kohler, N. 1974. Polymer Flooding: the Essential Elements for

Laboratory Evaluation. Paper SPE 4745 presented at SPE/AIME Improved Oil

Recovery Symposium, Tulsa, OK. April 22-24.

Chauveteau, G., Denys, K., and Zaitoun, A. 2002. New Insight on Polymer Adsorption

under High Flow Rates. Paper SPE 75183 presented at the 2002 SPE/DOE Improved

Oil Recovery Symposium, Tulsa, OK, April 13-17.

Chiappa, L., Mennella, A., and Burrafato, G. 1998. Polymer/Rock Interactions in

Polymer Treatments for Water-cut Control. Paper SPE 39619, presented at SPE/DOE

Improved Oil Recovery Symposium, Tulsa, OK, April 19-22.

Cohen, Y. and Christ, F.R. 1986. Polymer Retention and Adsorption in the Flow of

Polymer Solutions through Porous Media. SPE Res Eng, March, 113-118.

Dabbous, M.K. 1977. Displacement of Polymers in Waterflooded Porous Media and Its

Effects on a Subsequent Micellar Flood. SPE J (October), 358-368.

Dang, C.T.Q., Chen, Z., Nguyen, N.T.B., et al. 2011. Development of Isotherm

Polymer/Surfactant Adsorption Models in Chemical Flooding. Paper SPE 147872

presented at the 2011 SPE Asia Pacific Oil and Gas Conference and Exhibition,

Jakarta, Indonesia, September 20–22.

Dawson, R. and Lantz, R. B. 1972. Inaccessible Pore Volume in Polymer Flooding. SPE

J, 24 (5): 448–452.

de Gennes, P.G. 1979. Scaling Concepts in Polymer Physics: Ithaca, NY: Cornell

University Press.

Delamaide, E., Zaithoun, A., Renard, G., and Tabary, R. 2013. Pelican Lake Field: First

Successful Application of Polymer Flooding in a Heavy Oil Reservoir. SPE paper

165234 presented at SPE Enhanced Oil Recovery Conference, Kuala Lumpur,

Malaysia, July 02-04.

107

Delshad, M., Kim, D.H., Magbagbeola, O.A., Huh, C., Pope, G.A., and Tarahhom, F.

2008. Mechanistic Interpretation and Utilization of Viscoelastic Behavior of Polymer

Solutions for Improved Polymer-Flood Efficiency. Paper SPE 113620 presented at the

2008 SPE/DOE Improved Oil Recovery Symposium held in Tulsa, OK, April 19-23.

Deng Y., Dixon, J.B., and White, G.N. 2006. Adsorption of Polyacrylamide on Smectite,

Illite and Kaolinite. Soil Sci. Soc. AM. J., 70 (January-February): 297–304.

Dominguez, J.G. and Willhite, G.P. 1977. Retention and Flow Characteristics of Polymer

Solutions in Porous Media. SPE J. 17 (2): 111–121.

Durst, F., Haas, R., and Interthal, W. 1982. Laminar and Turbulent Flows of Dilute

Polymer Solutions: A Physical Model. Rheologica Acta, 21 (4-5): 572-577.

Espinasse, P. and Siffert, B. 1979. Acetamide and Polyacrylamide Adsorption onto Clays:

Influence of the Exchangeable Cations and the Salinity of the Medium. Clays and

Clay Minerals, 27 (4): 279–284.

Ferry, J. D. 1948. Viscoelastic Properties of Polymer Solutions. J of Research of the

National Bureau of Standards, 41 (1): 53–61.

Foshee, W.C., Jennings, R.R., and West, T.J. 1977. Preparation and Testing of Partially

Hydrolyzed Polyacrylamide Solutions. Paper presented at 28st

Annual Technical

Meeting of the Petroleum Society of CIM, Edmonton, May 30-June 3.

Gogarty, W.B. 1967. Mobility Control with Polymer Solutions. SPE J, 7 (2): 161–170.

Graessley, W.W. Adv. Polym.Sci, 1974, 16,1.

Green, D.W. and Willhite, G.P. 1998. Enhanced Oil Recovery, Vol. 6, 107–110.

Richardson, TX: Textbook Series, SPE.

Grossman, P.D. and Soane, D.S. Biopolymers, 1991, 31 (10): 1221.

Gupta, S.P. and Trushenski, S.P. 1978. Micellar Flooding-the Propagation of the Polymer

Mobility Buffer Bank. SPE J, 18 (1): 5-12.

He, Q., Young, T., Willhite, G.P., and Green D.W. 1990. Measurement of Molecular

Weight Distribution of Polyacrylamides in Core Effluents. SPE Res Eng, August, 333-

338.

Hirasaki, G.J. and Pope, G.A. 1974. Analysis of Factors Influencing Mobility and

Adsorption in the Flow of Polymer Solution through Porous Media. SPE J, 14 (4):

337-346.

Huang, Y. and Sobie, K.S. 1993. Scleroglucan Behavior in Flow through Porous Media:

Comparison of Adsorption and In-Situ Rheology with Xanthan. Paper SPE 25173

presented at the International Symposium on Oilfield Chemistry, New Orleans, LA,

March 2–5.

Hughes, D.S., Teeuw D., Cottrell, C.W., et al. 1990. Appraisal of the Use of Polymer

Injection to Suppress Aquifer Influx and to Improve Volumetric Sweep in a Viscous

Oil Reservoir. SPE Res Eng, 5 (1): 33–40.

Huh, C., Lange, E.A., and Cannella, W.J. 1990. Polymer Retention in Porous Media.

Paper SPE 20235 presented at the SPE/DOE Symposium on Enhanced Oil Recovery,

Tulsa, OK, April 22–25.

Jennings, R.R., Rogers, J.H., and West, T.J. 1971. Factors Influencing Mobility Control

by Polymer Solutions. J. Pet Tec. 23 (3): 391-401.

Ke, F., Mo, X., and Liang, D. 2009. Effect of Overlap Concentration and Persistence

Length on DNA Separation in Polymer Solutions by Electrophoresis. Chinese J of

Poly Sci, Vol. 27 (5), 601-610.

108

Koral, J., Ullman, R., and Eirich, R. 1957. The Adsorption of Polyvinyl Acetate. Ibid,

Vol. 62, 541-550.

Liauh, W.C., Duda, J.L., and Klaus, E.E. 1978. An Investigation of the Inaccessible Pore

Volume Phenomena. Paper SPE 8751 presented at the 84th

National American Institute

of Chemical Engineers Meeting, Atlanta, February.

Liu, He., Li, J., Yang, J, et al. 2009. Successful Practices and Development of Polymer

Flooding in Daqing Oilfield. Paper SPE 123975 presented at the SPE Asia Pacific Oil

and Gas Conference and Exhibition, Jakarta, Indonesia, August 4-6.

Lotsch, T., Muller, T., Pusch, G. 1985. The Effect of Inaccessible Pore Volume on

Polymer Core Experiments. Paper SPE 13590 presented at the International

Symposium on Oilfield and Geothermal Chemistry, Phoenix, Arizona, April 9–11.

Macosko, C.W. 1993. Rheology Principles, Measurements, and Applications. John Wiley

& Sons, Inc, Hoboken, NJ. 142-152.

Maerker J.M. 1973. Dependence of Polymer Retention on Flow Rate. J. Petr Tech. 25

(11), 1307.

Maerker, J.M. 1975. Shear Degradation of Partially Hydrolyzed Polyacrylamide

Solutions. SPE J, 15 (4): 311-322.

Martin, F.D. and Sherwood, N.S. 1975. The Effect of Hydrolysis of Polyacrylamide on

Solution Viscosity, Polymer Retention and Flow Resistance Properties. Paper SPE

5339 presented for the Rocky Mountain Regional Meeting of SPE/AIME, Denver,

Colorado, April 7-9.

Meister, John J., Huey Pledger Jr,. Hogen-Esch, Thieo E., and Butler, George B. 1980.

Retention of Polyacrylamide by Berea Sandstone, Baker Dolomite, and Sodium

Kaolinite during Polymer Flooding. Paper SPE 8981 presented at SPE Fifth

International Symposium on Oilfield and Geothermal Chemistry, Stanford, California,

May 28-30.

Mezzomo, R.F., Moczydlower, P., Sanmartin, A.N., et al. 2002. A New Approach to the

Determination of Polymer Concentration in Reservoir Rock Adsorption Tests. Paper

SPE 75204 presented at the SPE/DOE Symposium on Improved Oil Recovery, Tulsa,

Oklahoma, April 13–17.

Mungan, N. 1969. Rheology and Adsorption of Aqueous Polymer Solutions. J. Can. Pet.

Tech, 8 (2): 45–50.

Ogunberu, A.L. and Asghari, K. 2004. Water Permeability Reduction under Flow-

Induced Polymer Adsorption. Paper 2004-236 presented at the Petroleum Society’s 5th

Canadian International Petroleum Conference (55th

Annual Technical Meeting),

Calgary, Alberta, Canada, June 8-10.

Osterloh, W.T. and Law, E.J. 1998. Polymer Transport and Rheological Properties for

Polymer Flooding in the North Sea Captain Field. Paper SPE 39694 presented at the

SPE/DOE Improved Oil Recovery Symposium, Tulsa, Oklahoma, April 19–22.

Pefferkorn, E., Nabzar, L., and Garroy, A. 1985. Adsorption of Polyacrylamide to Na

Kaolinite: Correlation between Clay Structure and Surface Properties. J. Colloid and

Interface Sci, 106 (1), 94-103.

Peterson, C. and Kwei, T.K. 1961. The Kinetics of Polymer Adsorption onto Solid

Surfaces. J. Phys. Che, 65 (8): 1330–1333.

109

Ranjbar, M., Rupp, J., and Pusch, G. 1991. Influence of Pore Radii Distribution on

Polymer Retention in Natural Sandstones. Presented at 6th

European Symposium on

Improved Oil Recovery, Stavanger, Norway, May 21-23.

Ranjbar, M., Rupp, J., Pusch, G., and Meyn, R. 1992. Quantification and Optimization of

Viscoelastic Effects of Polymer Solutions for Enhanced Oil Recovery. SPE paper

24154 presented at the SPE/DOE Eighth Symposium on Enhanced Oil Recovery,

Tulsa, Oklahoma, April 22-24.

Rowland, F.W. 1963. Thickness and Structure of Layers of High Polymer Adsorbed from

Solution onto Surfaces. PhD Dissertation, Polytechnic Inst. of Brooklyn, Brooklyn,

NY.

Rowland, F.W. and Eirich, F.R. 1966. Flow Rates of Polymer Solutions through Porous

Disks as a Function of Solute, II Thickness and Structure of Adsorbed Polymer Films.

J. Poly Sci, Part A-1, Vol. 4, 2401.

RP 63, Recommended Practices for Evaluation of Polymers Used in Enhanced Oil

Recovery Operations. 1990. Washington, DC: API

Satter, A., Shum, Y.M., Adams, W.T., et al. 1980. Chemical Transport in Porous Media

with Dispersion and Rate-controlled Adsorption. SPE J, 20 (3): 99-104.

Seright, R.S. 1983. The Effects of Mechanical Degradation and Viscoelastic Behavior on

Injectivity of Polyacrylamide Solutions. SPE J. 23 (3): 475-485.

Seright, R.S. 1992. Effect of Permeability and Lithology on Gel Performance. Paper SPE

24190 presented at the SPE/DOE Symposium on Enhanced Oil Recovery, Tulsa, OK,

April 22-24.

Seright, R.S. 1993. Effect of Rock Permeability on Gel Performance in Fluid-Diversion

Applications. In Situ, 17 (4): 363-386.

Seright, R.S. 2010. Potential for Polymer Flooding Reservoir with Viscous Oils. SPE Res

Eva & Eng, August, 730-740. SPE-129899-PA.

Seright, R.S. and Martin, F.D. 1993. Impact of Gelation pH, Rock Permeability, and

Lithology on the Performance of a Monomer-Based Gel. Spere, 8 (1), February 43-50.

Seright, R.S., Fan, T., Wavrik, K., et al. 2010. New Insights into Polymer Rheology in

Porous Media. SPE J, 16 (1): 35–42.

Seright, R.S., Prodanovic, M., and Linduist, W.B. 2006. X-Ray Computed

Microtomography Studies of Fluid Partitioning in Drainage and Imbibition before and

after Gel Placement. SPE J, 11(2): 159-170. SPE-89393-PA. doi: 10.2118/89393-PA.

Seright, R.S., Seheult, M., and Talashek, T. 2009. Injectivity Characteristics of EOR

Polymers. SPE Res Eval & Eng, 12 (5): 783–792.

Shah, B.N., Willhite, G.P., and Green, D.W. 1978. The Effect of Inaccessible Pore

Volume on the Flow of Polymer and Solvent through Porous Media. Paper SPE 7586,

presented at the Fall Technical Conference and Exhibition of the SPE of AIME,

Houston, Texas, Oct 1–3.

Shah, S., Heinle, S.A., and Glass, J.E. 1985. Water-Soluble Polymer Adsorption from

Saline Solutions. Paper SPE 13561 presented at the International Symposium on

Oilfield and Geothermal Chemistry, Phoenix, Arizona, April 9-11.

Smith, F.W. 1970. The Behavior of Partially Hydrolyzed Polyacrylamide Solutions in

Porous Media. J of Pet Tech. February, 148-156.

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

110

Southwick, J.G. and Manke, C.W. 1988. Molecular Degradation, Injectivity, and Elastic

Properties of Polymer Solutions. SPE Res Eng, 3 (4), 1193-1201.

Stromberg, R.R., Quasium, A.R., Toner, S.D., and Parker, M.S. 1959. Adsorption of

Polyesters on Glass, Silica, and Alumina. J of Research of the National Bureau of

Standards. Vol. 62 (2), February, 71-77.

Stutzmann, T. and Siffert, B. 1977. Contribution to the Adsorption Mechanism of

Acetamide and Polyacrylamide onto Clays. Clays and Clay Minerals. Vol. 25, 392-

406.

Sydansk, Robert D. and Romero-Zeron, Laura. 2011. Reservoir Conformance

Improvement. Society of Petroleum Engineers. Richardson, Texas: SPE. 57-61.

Szabo, M.T. 1979. An Evaluation of Water-Soluble Polymers for Secondary Oil

Recovery-Part 2. J. Pet Tech, 31 (5): 561–570.

Szabo, M.T. and Corp, C. 1975. Some Aspects of Polymer Retention in Porous Media

Using a C14-Tagged Hydrolyzed Polyacrylamide. SPE J, 15 (4): 323–337.

Taber, J.J, Martin, F.D., and Seright, R.S. 1997a. EOR Screening Criteria Revisited-Part

1: Introduction to Screening Criteria and Enhanced Recovery Field Projects. SPE Res

Eng, 12 (3): 189-198. SPE-35386-PA.

Taber, J.J, Martin, F.D., and Seright, R.S. 1997b. EOR Screening Criteria Revisited-Part

2: Applications and Impact of Oil Prices. SPE Res Eng 12 (3): 199-206. SPE-39234-

PA.

Urbissinova, T.S., Tribedi, J.J., and Kuru, E. 2010. Effect of Elasticity during

Viscoelastic Polymer Flooding: A Possible Mechanism of Increasing the Sweep

Efficiency. J of Can Petr Tech, December, 49 (12): 49-56.

Van Domselaar, H.R. and Fortmuller, C. 1992. On the Transport Properties of a Rod-

Type Polymer in Sand Packs. Paper SPE 25073 presented at the European Petroleum

Conference, Cannes, France, November 16-18.

Vela, S., Peaceman, D.W., and Sandvik, E.I. 1976. Evaluation of Polymer Flooding in a

Layered Reservoir with Crossflow, Retention, and Degradation. SPE J, 16 (2): 82-96.

Viovy, J.L. and Duke, T. Electrophoresis, 1993, 14(4): 322.

Vossoughi, S., Smith, J.E., Green, D.W., and Willhite, G.P. 1984. A New Method to

Simulate the Effects of Viscous Fingering on Miscible Displacement Processes in

Porous Media. SPE J, 24 (1): 56–64.

Wang, D., Xia, H., Liu Z., and Yang, Q. 2001. Study of the Mechanism of Polymer

Solution with Viscoelastic Behavior Increasing Microscopic Oil Displacement

Efficiency and the Forming of Steady “Oil Thread” Flow Channel. Paper SPE 68723

presented at the SPE Asia Pacific Oil and Gas Conference and Exhibition held in

Jakarta, Indonesia, April 17-19.

Wang, D., Xia, H., Yang, S., and Wang, G. 2010. The Influence of Visco-elasticity on

Micro Forces and Displacement Efficiency in Pores, Cores and in the Field. Paper SPE

127453 presented at the EOR Conference at Oil & Gas West Asia, Muscat, Oman,

April 11-13.

Ying, Q., and Chu, B. 1987. Overlap Concentration of Macromolecules in Solution.

Macromolecules, 20 (2): 362–366.

Yuan, C., Delshad, M., and Wheeler, M.F. 2010. Parallel Simulations of Commercial-

Scale Polymer Floods. Paper SPE 123441 presented at the 2010 SPE Western

Regional Meeting, Anaheim, California, May 27–29.

111

Zaitoun, A and Kohler, N. 1987. The Role of Adsorption in Polymer Propagation through

Reservoir Rocks. Paper SPE 16274 presented at the SPE International Symposium on

Oil Chemistry, San Antonio, Texas, February 4-6.

Zhang, Z., Li, J., and Zhou, J. 2010. Microscopic Roles of “Viscoelasticity” in HPAM

Polymer Flooding for EOR. Transp Porous Med, Vol. 86: 199-214.

Zheng, C.G., Gall, B.L., Gao, H.W., et al. 1998. Effects of Polymer Adsorption and Flow

Behavior on Two-Phase Flow in Porous Media. Paper SPE 39632 presented at

SPE/DOE Improved Oil Recovery Symposium, Tulsa, Oklahoma, April 19–22. Zitha, P.L.J. and Botermans, C.W. 1998. Bridging Adsorption of Flexible Polymers in

Low-permeability Porous Media. SPE Prod & Faci, February 15-20.

Zitha, P.L.J., van Os, K.G.S., and Denys, K.F.J. 1998. Adsorption of Linear Flexible

Polymers during Laminar Flow through Porous Media: Effect of the Concentration.

Paper SPE 39675 presented at SPE/DOE Improved Oil Recovery Symposium, Tulsa,

OK. April 19-22.

Zitha, Pacelli L. J., Chauveteau Guy, and Leger Liliane. 2001. Unsteady-state Flow of

Flexible Polymers in Porous Media. J of Col and Int Sci, 234, 269-283.

SI Metric Conversion Factors

ft × 3.048* E−01 = m

in. × 2.54* E+00 = cm

lb × 4.535 9237 E−01 = kg

md × 9.869 233 E−01 = m2

*Conversion factor is exact.