Module 6

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 91

PETE 609 - Module 6

Improved Water Flooding Processes

Module 6 Improved Waterflooding Processes


Estimated Duration: 2 weeks
Improved Waterflooding Processes.
Polymer Flooding.
Rheology of Polymer Solutions.
Polymer Adsorption and Retention.
Micellar-Polymer or Microemulsion Flooding.
Properties of Surfactants and Cosurfactants.
Surfactant-Brine-Oil Phase Behavior.
Caustic Flooding.
Performance evaluation.
Suggested reading: MAB, L, R24

Learning Objectives
After completion of this module, you will be able to:

List typical polymer types and compare their physical properties.

Understand the rheological behavior of polymers.

Understand the importance of salinity, temperature, polymer concentration, and


shear rate in determining the polymer solution properties.

Understand the principles of permeability reduction.

Summarize polymer retention mechanisms.

Evaluate polymer adsorption.

Learn the major steps involved in designing a polymer flood.

Class Notes for PETE 609 Module 6


Author: Dr. Maria Antonieta Barrufet Fall, 2001

Page 1/91

PETE 609 - Module 6


Improved Water Flooding Processes

Compare polymer versus waterflood using a Polymer Flood Simulation program


provided in class.

Analyze various production strategies in polymer floods (slug size, concentration,


injection rates, etc.).

Recognize the importance of lowering the interfacial tension (Micellar-Polymer


flooding, and caustic flooding).

Identify the terminology and identify process variations.

Define the chemical basis for surfactants.

Identify types of phase behavior for different oil/surfactant/brine systems.

Define and use optimal salinity.

Summarize caustic flooding mechanisms.

Evaluate fractional flows in caustic flooding processes.

Improved Waterflooding Processes


This module covers a variety of methods classifies as improved waterflooding
processes. These are implemented and analyzed in the same way as a waterflood, but
with some variations and complexities. Improved waterflooding processes consist in
changing the properties of the water (brine) used by adding some chemicals. The target
of these chemicals is to increase the water viscosity to improve mobility ratios, or to
lower the IFT between oil and water by using surfactants, or a combination of both.
These surfactants are either added to the water (micellar flooding), or generated in-situ
by chemical reactions of acidic components of the oil with a caustic solution.
In this module, we will cover,

Polymer Flooding,

Micellar- Polymer Flooding,

Alkaline Flooding.

Class Notes for PETE 609 Module 6


Author: Dr. Maria Antonieta Barrufet Fall, 2001

Page 2/91

PETE 609 - Module 6


Improved Water Flooding Processes

There are other improved waterflooding processes such as foam flooding which will not
be covered in this course due to time constraints. Most of the material for this section
has been taken from Lake (1989).

Polymer Flooding
Polymer flooding consists of adding polymer to the water of a waterflood to lower its
mobility. The resulting increase in viscosity, as well as a decrease in aqueous phase
permeability that occurs with some polymers, causes a lower mobility ratio.
A lower mobility ratio increases the efficiency of the waterflood by increasing the
volumetric sweep efficiency. Although polymer addition does not normally decrease
residual oil saturation it reduces the water-cut as oil is recovered.
It has also been reported that the addition of some polymer reduces rock permeability to
the polymer-water solution, thereby favorably reducing water mobility and increasing
both areal and vertical reservoir sweep efficiency.
The greater recovery efficiency constitutes the economic incentive for polymer flooding
when applicable. Generally, a polymer flood will be economic only when the

waterflood mobility ratio is high,

the reservoir heterogeneity is high,

a combination of these two occurs.

Polymers have been used in oil production in three modes.


1. As near-well treatments to improve the performance of water injectors or
watered-out producers by blocking off high-conductivity zones.
2. As agents that may be cross-linked in situ to plug high-conductivity zones
at depth in the reservoir.
3. As agents to lower water mobility or water-oil mobility ratio.

Class Notes for PETE 609 Module 6


Author: Dr. Maria Antonieta Barrufet Fall, 2001

Page 3/91

PETE 609 - Module 6


Improved Water Flooding Processes

The first mode is not truly polymer flooding since the effect is to lower the water
production, and due to this lower water/oil ratio the oil production may become
profitable.
The second mode requires that polymer be injected with an inorganic metal cation that
will cross-link subsequently injected polymer molecules with ones a lready bound to solid
surfaces.
Most polymer EOR projects have been in the third mode, the one we emphasize here.
We discussed how lowering the mobility ratio affects displacement and volumetric
sweep efficiency in Module 3.
Figure 1 shows a schematic of a typical polymer flood injection sequence along with
typical concentrations of chemicals and pore volumes injected.

Chase
Water

Taper

Mobility buffer
3

250-2500g/cm
polymer
Biocide

Mobility
Buffer

Slug

Preflush

Slug

Preflush

Polymer Solution

Electrolyte (Na+ , Ca++ ,


etc.)

(2500-10000)g/cm 3
5-20% Vpf

0-100% Vpf

Sacrificial chemicals
0-100% Vpf

Figure 1 - Schematics of a typical polymer flood injection sequence.

The oil bank at the front end is being pushed by a preflush, which usually consists of a
low-salinity brine to condition the reservoir. Next follows the slug of the polymer solution
itself, followed by a freshwater buffer to protect the polymer solution from backside
dilution. Many times the buffer contains polymer in decreasing amounts (a grading or
taper) to lessen the unfavorable mobility ratio between the chase water and the polymer
solution, finally the sequence ends with the chase or drive water.
Because of the driving nature of the process polymer floods are always done through
separate sets of injection and production wells.

Class Notes for PETE 609 Module 6


Author: Dr. Maria Antonieta Barrufet Fall, 2001

Page 4/91

PETE 609 - Module 6


Improved Water Flooding Processes

Mobility is lowered in a polymer flood by injecting water that contains a high molecular
weight, water-soluble polymer. Since the water is usually a dilution of oil-field brine,
interactions with salinity are important, particularly for certain classes of polymers.
Salinity is quantified by the total dissolved solids (TDS) content of the aqueous phase.
Chemical flooding properties depend on the concentrations of specific ions rather than
salinity only. The aqueous phase's total divalent cation content is known as hardness is
usually more critical to chemical flood properties than the same TDS concentration.
These divalent cations are mostly Ca++ and Mg++ .
Salinities from representative oil-field brines can range from as low as a few thousand
ppm to over 250,000 ppm. While hardness, which is expressed as the ion concentration
of Ca++ and Mg ++ may be from 100 to about 20,000 ppm. (Lake, 1989).
Because of the high molecular weight the water soluble polymers used in this EOR
technique (1 to 3 million), only a small amount (about 500 g/m3) of polymer will bring
about a substantial increase in water viscosity. Further, several types of polymers have
been reported to lower the mobility by reducing the water relative permeability in
addition to increasing the water viscosity. In reality the oil permeability is also altered
and there is a reduction of both oil and water, but this effect is more pronounced in the
water relative permeability.
The mechanisms by which polymer lowers mobility, and the interactions with salinity,
can be qualitatively illustrated by discussing polymer chemistry, and polymer rheology.

Polymer Characteristics

Polymers Used
Some of the polymer types that have been considered for polymer flooding include:

Xanthan gum (biopolymer)

Hydrolyzed polyacrylamide (HPAM)

Class Notes for PETE 609 Module 6


Author: Dr. Maria Antonieta Barrufet Fall, 2001

Page 5/91

PETE 609 - Module 6


Improved Water Flooding Processes

Copolymers (a polymer consisting of two or more different types of monomers) of


acrylic acid and acrylamide

Copolymers of acrylamide and 2-acrylamide 2-methyl propane sulfonate


(AM/AMPS)

Hydroxyethylcellulose (HEC)

Carboxymethylhydroxyethylcellulose (CMHEC)

Polyacrylamide (PAM)

Polyacrylic acid

Glucan

Dextran polyethylene oxide (PEO)

Polyvinyl alcohol

Modified starches

Virtually all the commercially attractive polymers fall into two generic classes:

polyacrylamides

polysaccharides (biopolymers)

Figure 2 shows representative molecular structures for these two major types.

Class Notes for PETE 609 Module 6


Author: Dr. Maria Antonieta Barrufet Fall, 2001

Page 6/91

PETE 609 - Module 6


Improved Water Flooding Processes

Figure 2 - Molecular structures of polyacrylamides and polysaccharides (Willhite and


Dominguez, 1977). Taken from Lake (1992).

Polyacrylamides
These are polymers whose monomeric unit is the acrylamide molecule. As used in
polymer flooding, polyacrylamides have undergone partial hydrolysis, which causes
anionic (negatively charged) carboxyl groups (-COO-) to be scattered along the polymer
chain. The polymers are called partially hydrolyzed polyacrylamides (HPAM) for this
reason.

Class Notes for PETE 609 Module 6


Author: Dr. Maria Antonieta Barrufet Fall, 2001

Page 7/91

PETE 609 - Module 6


Improved Water Flooding Processes

Typical degrees of hydrolysis are 30%-35% of the acrylamide monomers; hence the
HPAM molecule is negatively charged, which accounts for many of its physical
properties.
This degree of hydrolysis has been selected to optimize certain properties such as
water solubility, viscosity, and retention. If the level of hydrolysis is too small, the
polymer will not be water soluble. If it is too large, its properties will be too sensitive to
salinity and hardness, loosing part of the functionality for which has been selected.
The viscosity increasing feature of HPAM lies in its large molecular weight.
This feature is accentuated by the anionic repulsion between polymer molecules and
between segments on the same molecule. The repulsion causes the molecule in solution to elongate and snag on others similarly elongated, an effect that accentuates the
mobility reduction at higher concentrations.
If the brine salinity or hardness is high, this repulsion is greatly decreased through ionic
shielding since the freely rotating carbon-carbon bonds allow the molecule to coil up.
This feature is illustrated in Figure 3. The shielding causes a corresponding decrease in
the effectiveness of the polymer since snagging is greatly reduced. Virtually all HPAM
properties are sensitive to salinity and hardness, an obstacle to using HPAM in many
reservoirs. On the other hand, HPAM is inexpensive and relatively resistant to bacterial
attack, and it exhibits permanent permeability reduction.

Class Notes for PETE 609 Module 6


Author: Dr. Maria Antonieta Barrufet Fall, 2001

Page 8/91

PETE 609 - Module 6


Improved Water Flooding Processes

-C-C-C-C-C-C-C-C-C-C-C-C-C-C-C-C-C-C-C-C-C-C-C-C-C-C-C-C-

Electrostatic
Repulsion

Na+C-CC-C- C-CNa+C-C-+C-CNa

Coiling in
Brines

Figure 3- Schematics of the structure on HPAM in fresh water and in brines.

Polysaccharides
These polymers are formed from the polymerization of saccharide molecules Figure 2
b), a bacterial fermentation process. This process leaves substantial debris in the
polymer product that must be removed before the polymer is injected (Wellington,
1980). The polymer is also susceptible to bacterial attack after it has been introduced
into the reservoir. These disadvantages are offset by the insensitivity of polysaccha ride
properties to brine salinity and hardness.
Figure 2 b shows the origin of this insensitivity. The polysaccharide molecule is
relatively nonionic and, therefore, free of the ionic shielding effects of HPAM.
Polysaccharides are more branched than HPAM, and the oxygen-ringed carbon bond
does not rotate fully; hence the molecule increases brine viscosity by snagging and
adding a more rigid structure to the solution. Molecular weights of polysaccharides are
generally around 2 million.
HPAM is usually less expensive per unit amount than polysaccharides, but when
compared on a unit amount of mobility reduction, particularly at high salinities, the costs
are close enough so that the preferred polymer for a given application is site specific.
Class Notes for PETE 609 Module 6
Author: Dr. Maria Antonieta Barrufet Fall, 2001

Page 9/91

PETE 609 - Module 6


Improved Water Flooding Processes

Historically, HPAM has been used in about 95% of the reported field polymer floods
(Manning et al., 1983). Both classes of polymers tend to chemically degrade at elevated
temperatures.

Polymer Preparation Methods


Polymers to prepare polymer solutions are p urchased in three different forms:

Powders

Broths

Emulsions

Powders, the oldest of the three methods, can be readily transported and stored with
small cost. They are difficult to mix because the first water contacting the polymer tends
to form very viscous layers of hydration around the particles, which greatly slow
subsequent dissolution.
Broths are aqueous suspensions of about 10 wt.% polymer in water which are much
easier to mix than powders. They have the disadvantage of being more expensive
because of the need to transport and store large volumes of water. Broths are quite
viscous so they can require special mixing facilities. In fact, it is this difficulty which limits
the concentration of polymer in the broth.
Emulsion polymers, the newest polymer form, contain up to 35 wt. % polymer solution,
suspended through the use of a surfactant, in an oil-carrier phase. Once this water-in-oil
emulsion is inverted (next lectures) the polymer concentrate can be mixed with make-up
water to the desired concentration for injection.

Polymer Properties
In this section, you will see qualitative trends, quantitative relations, and representative
data on the following properties:

Polymer Rheology

Viscosity relations

Class Notes for PETE 609 Module 6


Author: Dr. Maria Antonieta Barrufet Fall, 2001

Page 10/91

PETE 609 - Module 6


Improved Water Flooding Processes

Non-Newtonian effects

Polymer transport

Inaccessible pore volume

Permeability reduction

Chemical and biological degradation

Mechanical degradation.

Polymer Rheology
To understand the behavior of the polymer viscosity we must introduce some
rheological concepts.
Rheology is the science that studies the deformation and flow of matter by describing
the manner in which materials respond to applied stress and strain.
Viscosity, one of the properties that affects flow the most, is needed in polymer flooding
processes, in the design of drilling fluids, and in transportation trough pipes and facilities
design.
Before classifying a fluid in terms of their rheological behavior we have to define some
terminology.
Stress is the ratio of force over area. Forces can be perpendicular (normal) to the area,
parallel, or combined. This gives rise to normal stresses (tensile or compressive) and to
shear stresses.
Figure 4 illustrates the concept of normal and shear stresses, as well as combined
stresses that induce bending.

Class Notes for PETE 609 Module 6


Author: Dr. Maria Antonieta Barrufet Fall, 2001

Page 11/91

PETE 609 - Module 6


Improved Water Flooding Processes

compressive force

tensile force

Shear forces parallel to the


surface they act
Combined forces inducing
bending
Figure 4 - Illustration of forces and stresses.

Stresses are a tensorial quantity and 9 nine separate quantities are required to describe
completely the state of stress in a material as indicated in Figure 5.

i j

Direction of face upon the


force acts

y
xy

xz
xx

2
3

Direction of the force

12

rr 11
= 21 22

31 12

13
13

33

z
Figure 5 - Tensorial nature of stresses.

Class Notes for PETE 609 Module 6


Author: Dr. Maria Antonieta Barrufet Fall, 2001

Page 12/91

PETE 609 - Module 6


Improved Water Flooding Processes

Another important definition is the Strain, which is defined as a relative deformation.


Mathematically is

l
l

(1)

and the instantaneous rate of strain or shear rate is defined as

& =

d
dt

(2)

Rheograms are plots of the variation of the shear stress (scalar) versus the shear rate.
The behavior of these plots gives rise to the classification of fluids in terms of their
rheological behavior. Figure 6 sketches typical responses that serve to classify fluids
accordingly.

Herschel-Bulkley

Bingham

Shear Stress

Shear- thinning

Newtonian

Shear - thickening
Shear Rate, 1/s

Figure 6 - Shear stress/shear rate behavior for fluids.


Class Notes for PETE 609 Module 6
Author: Dr. Maria Antonieta Barrufet Fall, 2001

Page 13/91

PETE 609 - Module 6


Improved Water Flooding Processes

For a Newtonian fluid the viscosity, which is defined as the slope of the shear stress
versus shear rate curve is constant. That is the viscosity is independent of the shear
rate. This slope is always positive and it can be very small for gases at standard
conditions, to very high for highly viscous fluids such as heavy oil.
For a shear-thinning fluid an apparent viscosity can be defined at the various shear
rates and this decreases with shear rate. Most polymers fall within this category.
For a shear-thickening fluid an apparent viscosity can be defined at the various shear
rates and this increases with shear rate.
For a Bingham-plastic fluid, the behavior is similar to a Newtonian fluid (constant
viscosity) but a certain shear stress must be applied before the fluid begins to flow.
For a Hershchel-Bulkley fluid, the behavior is similar to a shear-thinning fluid
(decreasing viscosity) but a certain shear stress must be applied before the fluid begins
to flow.

Non-Newtonian Effects
Polymers are non-newtonian fluids - which means viscosity is not a constant. They are
defined as pseudoplastic under most conditions, and their viscosity decreases with
increasing shear rates, which accompany increasing flow rates.
At high flow rates encountered in injection wells polymers sometimes deviate from
pseudoplastic behavior and exhibit viscoelastic effects. This results in an increasing
apparent viscosity, caused by the polymer rapidly moving through expansions and
contractions within the rock matrix.
Additional variables influence the polymer apparent viscosity and can be evaluated for
specific polymers and water available for mixing. These include mix water composition,
polymer molecular weight, degree of hydrolysis and concentration.
The freshest economically available water compatible with the rock, and the lowest
polymer concentration that furnishes the desired mobility should be used.
Figure 7 sketches the behavior of most polymer solutions used for oil recovery. Note the
different parameters that affect the viscosity.

Class Notes for PETE 609 Module 6


Author: Dr. Maria Antonieta Barrufet Fall, 2001

Page 14/91

PETE 609 - Module 6


Improved Water Flooding Processes

Figure 7- Typical viscosity response of polymer solutions versus shear rate. (From Core
laboratories manual).

Variable flow rate tests on core samples yield data that reflect polymer viscosity. Data
are referred to as Reciprocal Relative Mobility, and vary with polymer concentration.
Figure 8 shows these tests on a commercial polymer determined by Core Lab
laboratories.
Pore sizes and distribution influence polymer flow. Polymer molecules do not move into
pores below some critical size, resulting in reduced adsorption in non contacted rock.
(A certain permeability may be needed for a specific polymer to prevent filtration on the
well bore injection face.)
Class Notes for PETE 609 Module 6
Author: Dr. Maria Antonieta Barrufet Fall, 2001

Page 15/91

PETE 609 - Module 6


Improved Water Flooding Processes

Figure 8 - Viscoelastic effects for a commercial polymer (Core laboratories manual).

Figure 9 shows polymer solution viscosity

versus shear rate

laboratory viscometer at fixed salinity. At low shear rates,


the solution behaves like a Newtonian fluid. At higher

& ,

&

measured in a

is independent of

& , and

decreases, approaching a

limiting value. A fluid whose viscosity decreases with increasing

&

is shear thinning.

The shear thinning behavior of the polymer solution is caused by the uncoiling and
Class Notes for PETE 609 Module 6
Author: Dr. Maria Antonieta Barrufet Fall, 2001

Page 16/91

PETE 609 - Module 6


Improved Water Flooding Processes

unsnagging of the polymer chains when they are elongated in shear flow. Below the
critical shear rate, the behavior is part reversible.

Figure 9 - Shear thinning behavior of various Xanflood polymer concentrations (from


Lake, 1992).

Figure 10 shows a viscosity-shear-rate plot at fixed polymer concentration with variable


NaCl concentration for an AMPS polymer. The sensitivity of the viscosity to salinity is
profound. As a rule of thumb, the polymer solution viscosity decreases a factor of 10 for
every factor of 10 increase in NaCl concentration. The viscosity of HPAM polymers and
HPAM derivatives are even more sensitive to hardness, but viscosities of
polysaccharide solutions are relatively insensitive to both. This behavior is favorable
because, for the bulk of a reservoir's volume,

&

is usually low (about 1-5 s-1), making it

possible to attain a design mobility ratio with a minimal amount of polymer. But near the
injection wells,

&

can be quite high, which causes the polymer injectivity to be greater

than that expected based on nominal viscosity.

Class Notes for PETE 609 Module 6


Author: Dr. Maria Antonieta Barrufet Fall, 2001

Page 17/91

PETE 609 - Module 6


Improved Water Flooding Processes

Figure 10 - Effect of salinity o n polymer solution viscosity (take from Lake, 1992)

The relative magnitude of this enhanced injectivity effect can be estimated once
quantitative definitions of shear rate in permeable media, and shear-rate-viscosity
relations are given.

Viscosity Relations
Figure 11 shows a plot of Xanflood viscosity versus polymer concentration. This type of
curve has traditionally been modeled by the Flory-Huggins equation (Flory, 1953).

2
3
p = b 1 + a1Cpw + a2Cpw
+ a3Cpw
+K

(3)

Where Cpw is the polymer concentration in the aqueous phase, b is the brine (solvent)
viscosity, and ai (i =1, 2, ) are constants depending upon the polymer used.

Class Notes for PETE 609 Module 6


Author: Dr. Maria Antonieta Barrufet Fall, 2001

Page 18/91

PETE 609 - Module 6


Improved Water Flooding Processes

The usual polymer concentration unit is g/m3 of solution, which is approximately the
same as ppm. A useful conversion to recall is that 1,000 g/m3 = 0.1% weight
approximately.
The linear term in Equation (3) accounts for the dilute range where the polymer
molecules act independently (without entanglements). For most purposes, this equation
can usually be truncated at the cubic term.

Figure 11 - Polymer (Xanflood) viscosity as a function of polymer concentration (1%


NaCl brine) at fixed shear rates (and Lake, 1992).

For a 1,000 g/m3 Xanflood solution at 0.1 s-1 in 1 wt % NaCl brine at 24 oC,
the viscosity is 70 mPa-s (70 cp). Compared to the brine at the same conditions, this is
a substantial increase in viscosity brought about by a relatively dilute concentration,
therefore Xanflood at these conditions is a n excellent thickener. However, note that a
shear rate of 5 s-1, the viscosity dropped to 10 cp.
A more fundamental way of measuring the thickening power of a polymer is through its
intrinsic viscosity which is defined as

Class Notes for PETE 609 Module 6


Author: Dr. Maria Antonieta Barrufet Fall, 2001

Page 19/91

PETE 609 - Module 6


Improved Water Flooding Processes

[] =

p b
limit

C pw 0 bC pw

(4)

From its definition, [] is a measure of the polymer's intrinsic thickening power. It is


insensitive to the polymer concentration. The intrinsic viscosity for the Xanflood polymer
under the conditions given above is 70 dl/g, the units being equivalent to reciprocal
weight percent. Intrinsic viscosity is the same as the a1 term in Equation (3).
For any given polymer-solvent pair, the intrinsic viscosity increases as the molecular
weight of the polymer increases according to the following equation (Flory, 1953):

= K' Mw

(5)

The exponent a varies between about 0.5 and 1.5 and is higher for good solvents such
as freshwater. K' is a polymer-specific constant.
These relationships are useful for characterizing the polymer solutions. For example,
the size of the polymer molecules in solution can be estimated from Flory's (1953)
equation for the mean end-to-end distance

d p = 8(Mw )

1/ 3

(6)

This is an empirical equation that requires certain units; [] must be in dl/g, and dp is
returned in Angstroms (10-10 m).
This measure of polymer size is useful in understanding how these very large molecules
propagate through the small pore openings of rocks. The molecular weight of Xanthan
gum is about 2 million.
Using Equation (5) for Xanthan gum gives a dp of about 0.4 m. This is the same size
as many of the pore throats in a low-to-moderate permeability sandstone. As a result,
we would expect to, and in fact do, observe many polymer-rock interactions.

Class Notes for PETE 609 Module 6


Author: Dr. Maria Antonieta Barrufet Fall, 2001

Page 20/91

PETE 609 - Module 6


Improved Water Flooding Processes

The relationship between polymer-solution viscosity and shear rate may be described
by a power-law model

p = K pl ( & )

n pl 1

(7)

where Kpl and npl are the power-law coefficient and exponent, respectively. For shear
thinning fluids, 0 < npl < 1; for Newtonian fluids, npl = 1, and Kpl becomes the viscosity,

&

is always positive.

Equation (7) applies only over a limited range of shear rates: Below some low shear
rate, the viscosity is constant at
viscosity is also constant at

= 0p

= p

).

) and above the critical shear rate, the

Another useful model is the Meter model (Meter and Bird, 1964)

p = p

0p p
&
1 +

&

1/ 2

where nM is an empirical coefficient, and


average of

0p

and

(8)

nM 1

& 1/ 2

is the shear rate at which

is the

p .

As with all polymer properties, all empirical parameters are functions of salinity,
hardness, and temperature.
When applied to permeable media flow, the above general trends and equations
continue to apply.
shear rate

& eq

is usually called the apparent viscosity

app

and the effective

is based on capillary tube concepts.

Class Notes for PETE 609 Module 6


Author: Dr. Maria Antonieta Barrufet Fall, 2001

Page 21/91

PETE 609 - Module 6


Improved Water Flooding Processes

For power-law fluids, the apparent viscosity of a flowing polymer solution is (Hirasaki
and Pope, 1974).

app = H plu

n pl 1

= K pl &

npl 1

(9)

where,

Hpl

1 + 3npl
= K pl
npl

n pl 1

(8kw w )(1 n pl )/ 2

(10)

Combining Equation (9) with Equation (10) yields the equivalent shear rate for flow of a
power-law fluid

1 + 3n pl
u

& eq =
n pl 8k
w w

The aqueous phase permeability


and the absolute permeability,

(11)

kw

is the product of the phase's relative permeability

the aqueous phase porosity defined as Sw .

Polymer Transport
Polymers are retained in permeable media because of adsorption onto solid surfaces or
trapping within small pores. Polymer retention varies with polymer type, molecular
weight, rock composition, brine salinity, brine hardness, flow rate, and temperature.

Class Notes for PETE 609 Module 6


Author: Dr. Maria Antonieta Barrufet Fall, 2001

Page 22/91

PETE 609 - Module 6


Improved Water Flooding Processes

Field-measured values of retention range from 7 to 150 g polymer/cm3 of bulk volume,


with a desirable retention level being less than about 20 g/cm3.
Retention causes the loss of polymer from solution, which can also cause the mobility
control effect to be lost - particularly at low polymer concentrations.

Evaluation of Polymer Retention


Polymer adsorption can be represented empirically using a Langmuir type isotherm

Cpr =

where

a pCp
1 + bpCp

Cp

and

Cpr

(12)

are the polymer concentrations in the aqueous and on the rock

phases.
The units of adsorption can take on a variety of forms, but mass of polymer per mass of
rock is most common. Typical shapes of polymer adsorption isotherms are indicated in
Figure 12.

Class Notes for PETE 609 Module 6


Author: Dr. Maria Antonieta Barrufet Fall, 2001

Page 23/91

PETE 609 - Module 6


Improved Water Flooding Processes

Figure 12 - Typical Langmuir isotherm shapes (Lake, 1989).

Polymer adsorption does increase with increasing salinity and hardness and it is
unknown wether adsorption is reversible. Typical polymer adsorption isotherms are
quite steep; that is, they attain their plateau value at very low concentrations .
Equation (12) is a general isotherm function. The specific form depends on the units of
the retention; unfortunately, no standard form exists for this. Common ways to report
retention are,

Mass Polymer / Mass Solid ** (more common)

Class Notes for PETE 609 Module 6


Author: Dr. Maria Antonieta Barrufet Fall, 2001

Page 24/91

PETE 609 - Module 6


Improved Water Flooding Processes

Mass Polymer / Surface Area

Mass Polymer / Bulk Volume

Mass Polymer / Pore Volume

Inaccessible Pore Volume


Offsetting the delay caused by retention is an acceleration of the polymer solution
through the permeable medium caused by inaccessible pore volume (IPV). The most
common explanation for IPV is that the smaller portions of the pore space will not allow
polymer molecules to enter because of their size. Thus a portion of the total pore space
is uninvaded or inaccessible to polymer, and accelerated polymer flow results.
IPV depends on polymer molecular weight, medium permeability, porosity, and pore
size distribution and becomes more pronounced as molecular weight increases and the
ratio of permeability to porosity (characteristic pore size) decreases. In extreme cases,
IPV can be 30% of the total pore space.

Permeability Reduction
For many polymers, viscosity-shear-rate data derived from a viscometer (viscosity
versus shear rate) and those derived from a flow experiment viscosity versus equivalent
shear rate) will yield essentially the same curve. But for HPAM, the viscometer curve
will be offset from the permeable medium curve by a significant and constant amount.
The polymer evidently causes a degree of permeability reduction that reduces mobility
in addition to the viscosity increase.
Permeability reduction is only one of three measures in permeable media flow
(Jennings et al., 1971). The resistance factor RF is the ratio of the injectivity (mobility) of
brine to that of a single-phase polymer solution flowing under the same conditions:
either constant flow rate or constant pressure drop.

kw p
w
RF =
= w app =

p
k

p w
Class Notes for PETE 609 Module 6
Author: Dr. Maria Antonieta Barrufet Fall, 2001

(13)

Page 25/91

PETE 609 - Module 6


Improved Water Flooding Processes

For constant flow rate experiments, RF is the inverse ratio of pressure drops; for
constant pressure drop experiments, RF is the ratio of flow rates.
RF is an indication of the total mobility lowering contribution of a polymer. To describe
the permeability reduction effect alone, a permeability reduction factor Rk is defined as

k w w
Rk =
=
R
k p p F

(14)

A final definition is the residual resistance factor RRF, which is the mobility of a brine
solution before (wb) and after (w a) polymer injection

RRF =

wb
wa

(15)

RRF indicates the permanence of the permeability reduction effect caused by the
polymer solution.
RRF is the primary measure of the performance of a channel-blocking application of
polymer solutions. For many cases, Rk and RRF are nearly equal, but RF is usually much
larger than Rk because it contains both the viscosity-enhancing and the permeabilityreducing effects.
Figure 13 shows some typical resistance factors measured by Core Labs on a
commercial polymer. Note the viscoelastic effects at high injection rates.
The most common measure of permeability reduction is Rk , which is sensitive to
polymer type, molecular weight, degree of hydrolysis, shear rate, and permeable media
pore structure. Polymers that have undergone even a small amount of mechanical
degradation seem to lose most of their permeability reduction effect. For this reason,
qualitative tests based on screen factor devices are common to estimate polymer
quality.
Class Notes for PETE 609 Module 6
Author: Dr. Maria Antonieta Barrufet Fall, 2001

Page 26/91

PETE 609 - Module 6


Improved Water Flooding Processes

Figure 13 - Typical resistance factors for a commercial polymer (Core Lab, laboratory
manual).

The screen factor device is simply two glass bulbs mounted into a glass pipette as
shown in Figure 14. Into the tube on the bottom of the device are inserted several fairly
coarse wire screens through which the polymer solution is to drain.

Class Notes for PETE 609 Module 6


Author: Dr. Maria Antonieta Barrufet Fall, 2001

Page 27/91

PETE 609 - Module 6


Improved Water Flooding Processes

To use the device, a solution is sucked through the screens until the solution level is
above the upper timing mark. When the solution is allowed to flow freely, the time
required to pass from the upper to the lower timing mark td is recorded.

Polym er Solutio n

2R

30 m l

Timing Marks

h1

h2
Five 1 00 mesh screens
0.25 inches in d iameter

Figure 14-Screen factor device.

The screen factor for the polymer solution is then defined as

SF =

td
t ds

(16)

where tds is the similar time for the polymer-free brine.

Class Notes for PETE 609 Module 6


Author: Dr. Maria Antonieta Barrufet Fall, 2001

Page 28/91

PETE 609 - Module 6


Improved Water Flooding Processes

Screen factors are particularly sensitive to changes in the polymer molecule itself. One
definition of polymer quality is the ratio of the degraded to the undegraded screen
factors. This use is important for screen factor devices, particularly in locations that
prohibit more sophisticated equipment.

Chemical and Biological Degradation


The average polymer molecular weight can be decreased, to the detriment of the overall
process, by chemical, biological, or mechanical degradation. We use the term chemical
degradation to denote any of several possible processes such as thermal oxidation, free
radical substitution, hydrolysis, and biological degradation.
For a given polymer solution, there will be some temperature above which the polymer
will actually thermally crack. Although not well established for most EOR polymers, this
temperature is fairly high, on the order of 260oF. Since the original temperature of oil
reservoirs is almost always below this limit, of more practical concern for polymer
flooding is the temperature other degradation reactions occur at.
The average residence time in a reservoir is typically very long, on the order of a few
years, so even slow reactions are potentially serious. Reaction rates also depend
strongly on other variables such as pH or hardness. At neutral pH, degradation often will
not be significant, whereas at very low or very high pH, and especially at high
temperatures, it may be. In the case of HPAM, the hydrolysis will destroy the carefully
selected extent of hydrolysis present in the initial product.
To prevent biological reactions the following chemicals are used (commonly used
bactericides)

Acrolein

Formaldehyde

Sodium dichlorophenol

Sodium pintachiorophenol

To prevent oxidation reactions the following chemicals are used

Class Notes for PETE 609 Module 6


Author: Dr. Maria Antonieta Barrufet Fall, 2001

Page 29/91

PETE 609 - Module 6


Improved Water Flooding Processes

Oxygen scavengers

Hydrazine

Sodium bisulfite

Sodium hydrosulfite

Sulfor dioxide

Oxidation or free radical chemical reactions are usually considered the most serious
source of degradation. Therefore, oxygen scavengers and antioxidants are often added
to prevent or retard these reactions. These chemicals are strong reducing agents and
have the additional advantage of reducing iron cations from the + 3 to the +2 state.
They, in turn, help prevent gelation, agglomeration, and other undesirable effects that
can cause wellbore plugging and reduced injectivity.
Wellingnton (1980) found that alcohols such as isopropanol and sulfur compounds such
as thiourea make good antioxidants and free radical inhibitors.
Biological degradation can occur with both HPAM and polysaccharides, but is more
likely with the latter. Variables affecting biological degradation include the type of
bacteria in the brine, pressure, temperature, salinity, and the other chemicals present.
As in waterflooding, the preventive use of biocide is highly recommended. Often too
little biocide is used or it is started too late, and the ensuing problems become almost
impossible to correct.

Mechanical Degradation
High shear flow rates can break the linear chain and reduce apparent viscosity. The
higher the salinity, the greater the susceptibility to shear degradation - with calcium
solutions more sensitive than sodium.
Mechanical degradation is potentially present under all applications. It occurs when
polymer solutions are exposed to high velocity flows, which can be present in surface
equipment (valves, orifices, pumps, or tubing), downhole conditions (perforations or

Class Notes for PETE 609 Module 6


Author: Dr. Maria Antonieta Barrufet Fall, 2001

Page 30/91

PETE 609 - Module 6


Improved Water Flooding Processes

screens), or the sand face itself. Perforated completions, particularly, are a cause for
concern as large quantities of polymer solution are being forced through several small
holes. For this reason, most polymer injections are done through open-hole or gravelpack completions.
Figure 15 shows typical viscosity reductions using these two types of completions.

% Viscosity Reduction

100
Perforated hole

50
Openhole or Gravel Pack

100

1,000
10,000
100,000
Maximum Shear Rate , 1/s

Figure 15 - Comparison of Viscosity reduction in different completions.

Partial preshearing of the polymer solution can lessen the tendency of polymers to
mechanically degrade. Because flow velocity falls off quickly with distance from an
injector, little mechanical degradation occurs within the reservoir itself.
All polymers mechanically degrade under high enough flow rates. But because of their
ionic nature HPAMs are most susceptible under normal operating conditions,
particularly if the salinity or hardness of the brine is high.

Class Notes for PETE 609 Module 6


Author: Dr. Maria Antonieta Barrufet Fall, 2001

Page 31/91

PETE 609 - Module 6


Improved Water Flooding Processes

Polymer Injectivity
The economic success of all EOR processes is strongly tied to project life or injection
rate, but polymer flooding is particularly susceptible. In many cases, the cost of the
polymer itself is secondary compared to the present value of the incremental oil.
Because of its importance, many field floods are preceded by single-well injectivity
tests.
Lake (1989) gives a simple technique for analyzing injectivity tests based on the
physical properties provided earlier.

The injectivity of a well (volumetric flow rate over pressure drop) is defined as

I=

i
P

(17)

where i is the volumetric injection rate into the well, and P is the pressure drop
between the bottom-hole flowing pressure and some reference pressure. Another useful
measure is the relative injectivity

Ir =

(18)

Iw

where Iw is the water injectivity.

Ir
I

is an indicator of the injectivity decline to be anticipated when injecting polymer. Both


and I r are functions of time, but the long time limit of I r for a Newtonian polymer

solution is simply the viscosity ratio if skin effects are small. However, the ultimate I r for
an actual polymer solution can be higher than this because of shear-thinning.

Class Notes for PETE 609 Module 6


Author: Dr. Maria Antonieta Barrufet Fall, 2001

Page 32/91

PETE 609 - Module 6


Improved Water Flooding Processes

This treatment is limited to several simplifying assumptions as follows

The well, of radius Rw, whose injectivity we seek, is in a horizontal,


homogeneous, circular drainage area of radius Re.

The pressures at Re, and Rw are Pe and Pwf , respectively. Pe is constant (steadystate flow), but Pwf can vary with time.

The fluid flowing in the reservoir is a single aqueous phase, at residual oil
saturation, which is incompressible with pressure-independent rheological
properties.

Dispersion and polymer adsorption are negligible although the polymer can
exhibit permeability reduction.

The flow is one-dimensional and radial.

Finally, the entire shear rate range in the reservoir lies in the power-law regime; hence
Equation (7) describes the apparent viscosity.
Subject to these assumptions, the continuity equation in radial coordinates reduces to

d
(ru r ) = 0
dr

(19)

where ur is the radial volumetric flux. This equation implies the volumetric rate is
independent of r and equal to i since

i = 2rH t ur

(20)

Equation (19) is a consequence of the incompressible flow assumption; however, i is


not independent of time.
Let us substitute Darcy's law for ur in Equation (20)

Class Notes for PETE 609 Module 6


Author: Dr. Maria Antonieta Barrufet Fall, 2001

Page 33/91

PETE 609 - Module 6


Improved Water Flooding Processes

i
dP k p
= ur =
A
dr app

(21)

using the definition of Rk

i=

2rH t k p dP
2rH t k b dP
=
n 1
app dr
H pl ur pl Rk dr

(22)

This equation has been defined so that i is positive.

Now replace

app from Equation (7). The permeability reduction factor Rk is introduced

through to replace kp.


Eliminating ur yields an ordinary differential equation, which may be integrated between
the arbitrary limits of P1, r1 and P2, r2.

i
P2 P1 =
2Ht

n pl

H pl Rk
1 n
1 n
r1 pl r2 pl
kb (1 n pl )

The Newtonian flow limit n pl = 1 = Rk and

Hpl = app

(23)

of this equation is the familiar

steady-state radial flow equation,

P2 P1 =

i b
r
ln 1
2k b Ht r2

Class Notes for PETE 609 Module 6


Author: Dr. Maria Antonieta Barrufet Fall, 2001

(24)

Page 34/91

PETE 609 - Module 6


Improved Water Flooding Processes

We now apply these equations to the polymer flood injectivity.

At some time tp during the injection, the polymer front (assumed sharp) is at radial
position Rp where

tp

idt = ( R

2
p

Rw2 )Ht (1 Sor )

(25)

The left side of this equation is the cumulative volume of polymer solution injected.
Therefore, Equation (23) applies in the region Rw < r < R p, and Equation (24) applies in
the annular region Rp < r < R e. With the appropriate identification of variables, we have
for the second region

Rp

Pe =

R
ib
ln e
2kbHt Rp

(26)

and for the first

Pwf P R p

where

P Rp

i
=

n pl

Hpl Rk
1 n
1 n
Rp pl Rw pl
k b (1 npl )

(27)

is the pressure at the polymer-water front. Adding these two equations

gives the total pressure drop from Rw to Re.

Class Notes for PETE 609 Module 6


Author: Dr. Maria Antonieta Barrufet Fall, 2001

Page 35/91

PETE 609 - Module 6


Improved Water Flooding Processes

i
Pwf Pe =

npl

ib
+
2kbHt

H pl Rk

kb ( 1 npl
R
ln e
R p

() R

1 npl
p

1n pl

Rw

)
(28)

+ sw

where SW , the intrinsic skin factor of the well, has been introduced to account for well
damage.
Equation (28)substituted into the injectivity definition Equation (17) gives

i
=
2 Ht

n pl

H pl Rk

ik b (1 n pl

) (R

1 n pl
p

1 n pl

Rw

)+ 2k H ln RR
b

+ sw

(29)

The water injectivity Iw is evaluated from Equation (24) with r1 = Rw and r2 = Re.
This and I, yield an expression for Ir as it will be evaluated in an exercise.
Both I and I r relate to the cumulative polymer solution injection (or to time).

Figure 16 Illustrates these two regions.

Class Notes for PETE 609 Module 6


Author: Dr. Maria Antonieta Barrufet Fall, 2001

Page 36/91

PETE 609 - Module 6


Improved Water Flooding Processes

Rw
Ht

Rp
Re

Rw

Rp
Re

Figure 16 - Schematics of the regions used to evaluate injectivity in radial geometry.

Fractional Flow in Polymer Floods


The fractional flow treatment of polymer floods is the same as seen in the previous
module (Here we will use a Polymer Flood software from DOE). The only major
complications are the addition of terms for polymer retention and inaccessible pore
volume (IPV).

Expected results are that:


1. The oil bank breakthrough time (reciprocal of the oil bank specific velocity)
increases as the oil saturation increases, suggesting polymer floods will be
more economic if they are begun at low initial water saturation. Of course, the

Class Notes for PETE 609 Module 6


Author: Dr. Maria Antonieta Barrufet Fall, 2001

Page 37/91

PETE 609 - Module 6


Improved Water Flooding Processes

lower Sw , the higher the mobile oil saturation, also a favorable indicator for
polymer floods.
2. Adsorption causes a delay of all fronts. This can be large if the porosity is low,
the retention is high, or the injected polymer concentration is low.
3. Inaccessible pore volume causes an acceleration of all fronts, exactly opposite to
retention. In fact, retention and IPV can exactly cancel so that the polymer
front and the denuded water front travel at the same velocity.

Figure 17 indicates the laboratory procedures commonly used to screen polymer flood
candidates.

Class Notes for PETE 609 Module 6


Author: Dr. Maria Antonieta Barrufet Fall, 2001

Page 38/91

PETE 609 - Module 6


Improved Water Flooding Processes

Polymer Floods

Reservoir Rock
Characterization

Crude Oil
Characterization

Thin Sections for


Mineral Content

Crude Oil
Equivalent
Molecular Wt.

X-Ray for Clay


Identification

IFT Tests to
Define
Equivalent Alkane
Carbon No.

Microemulsion
Formulation
and Core Testing
Phase Behavior of
Crude+Brine+Surfactant+
Cosurfactant (middle
phase desired)
Microemulsion
viscosity vs.
Cosurfactant
composition
& concentration

SEM for Clay


Location

Polymer Buffer
Selection

Relative Permeability

Micellar Adsorption
Drainage

Imbibition
Core Displacement
Studies to Define Sor

Capillary Number

Water Permeability

Waterflood core
to residual oil
Inject micellar
slug
Inject polymer

Figure 17 - Laboratory screening procedures for designing a polymer flood.

Elements of a Polymer Flood Design


Polymer flood design is a complex subject. But most of the complexity arises from
reservoir-specific aspects of a particular design. In this section, we deal in generalities
that apply to all types of polymer flooding.

Class Notes for PETE 609 Module 6


Author: Dr. Maria Antonieta Barrufet Fall, 2001

Page 39/91

PETE 609 - Module 6


Improved Water Flooding Processes

A polymer flood design procedure will follow the following steps.


Screen the candidate reservoirs The distinction between technical and
economic feasibility is important. Technical feasibility means a given
reservoir can be polymer flooded regardless of the funds available.
Economic feasibility means the project has a good chance of being
profitable. Technical feasibility screening parameters are only two: the
reservoir temperature should be less than about 170 oF to avoid
degradation, and the reservoir permeability should be greater than about
0.02 m2 to avoid plugging. Economic feasibility can be estimated by
simple hand calculations (as in the fractional flow method) or through using
predictive models (Jones et al. 1984), which requires deciding how the
polymer is to be used.
Decide on the correct mode of polymer use The choices are (a) mobility
control (decrease M), (b) profile control (improve the permeability profile at
the injectors or producers), or (c) some combination of both. We want to
inject an agent that will alter the permeability so that more fluid will go into
the tight rock than into the high-permeability rock. We can do this by using
gels, polymers, and solids and by using selective perforation. When
selective perforation is ineffective or incompletely effective, we use
chemical agents or solids.
Select the polymer type. The requirements for EOR polymers are severe.
An outline of the principal ones is as follows:
Good thickening. This means high mobility reduction per unit cost.
High water solubility. The polymers must have good water solubility
under a wide range of conditions of temperature, electrolyte
composition, and in the presence of stabilizer additives.
Low retention. All polymers adsorb on reservoir rocks to various
degrees. Retention may also be caused by plugging, trapping,
phase separation, and other mechanisms. Low is less than 20 g/g.
Shear stability. During flow through permeable media, stress is applied
to the polymer molecules. If this is excessive, they may mechanically break apart or permanently degrade, resulting in less
viscosity. HPAM is especially subject to shear degradation.

Class Notes for PETE 609 Module 6


Author: Dr. Maria Antonieta Barrufet Fall, 2001

Page 40/91

PETE 609 - Module 6


Improved Water Flooding Processes

Chemical stability. Polymers, like other molecules, can chemically react,


especially at high temperature and in the presence of oxygen.
Antioxidants are used to prevent this
Biological stability. Both HPAM and polysaccharides can be degraded
by bacteria, but the latter are more susceptible. Biocides are
required to prevent this.
Good transport in permeable media. This means the ability to propagate
the polymer through the rock intact and without excessive pressure drop or
plugging. Good transport also means good injectivity and no problems with
microgels, precipitates, and other debris. Obviously, no one polymer can
universally meet these requirements for all reservoir rocks. Thus we must
tailor the polymer to the rock to some extent. Some general guidelines are
possible for minimum standards, but the ultimate criterion must be
economics.
Estimate the amount of polymer required. The amount, the total mass in
kilograms to be injected, is the product of the slug size, the pore volume,
and the average polymer concentration. Ideally, the amount would be the
result of an optimization study that weights the present value of the
incremental oil against the present value of the injected polymer.
Design polymer injection facilities. Getting a good quality solution is, of
course, important, but the cost of the injection facilities is usually small
corn- pared to well and chemical costs. The three essential ingredients are
mixing facilities, filtration, and injection equipment. The type of mixing
apparatus depends on the polymer. For solid polymers, a skid-mounted
solid mixer is required. Concentrates or emulsion polymers require
somewhat less sophistication although the latter may require some
emulsion breaking. Filtration largely depends on the success of the mixing,
but ordinarily it is no more stringent than what is required by waterflooding.
But if exotic and difficult filtration is required, the complexity and cost can
become significant. Injection equipment is the same as that for
Waterflooding. All surface and downhole equipment should be modified to
avoid all forms of degradation.
Consider the reservoir. Little is required here beyond the usual waterflood
considerations such as the optimal well pattern and spacing, completion

Class Notes for PETE 609 Module 6


Author: Dr. Maria Antonieta Barrufet Fall, 2001

Page 41/91

PETE 609 - Module 6


Improved Water Flooding Processes

strategy. pattern allocation (balance), reservoir characterization, and


allowable injection rates.
Optimal values of these quantities imply precise values that will result in the
maximum rate of return on investment. Since several quantities are
involved, it is usually not possible to perform optimizations on everything.
Hence most of the parameters must be fixed by other considerations (such
as striving for a target mobility ratio). But for the most sensitive quantities,
optimization is required.

Micellar-Polymer Flooding
Most of this material is taken from Lake (1992).
An injected surfactant lowers interfacial tension (IFT) between oil and water, and
thereby recovers the residual oil that normally remains after water flood. This is
followed by polymer to efficiently displace the surfactant and mobilized oil. Most
projects inject a surfactant slug, and the laboratory tests assist in formulation of the slug
and the polymer solution to push it. Displacement tests furnish surfactant oil recovery
data in the reservoir rock.
Mobility design and control is an essential part of field application. Without proper
mobility, surfactant will finger through the oil bank formed, or polymer drive water will
finger through the surfactant bank. Fingering reduces oil recovery.
Capillary forces due to large IFT between oil and water resist the externally applied
viscous forces and cause the injected water to bypass the oil.
Micellar-Polymer (MP) flooding is the predominant EOR technique to lower the IFT. We
have already seen that very low IFTs, of the order of 1N/m, are required to
substantially lower the residual oil saturation. These very low IFTs can be achieved by
using highly surface active chemicals.
Recall that the competition between viscous forces and capillary forces was given by
the capillary number.

Class Notes for PETE 609 Module 6


Author: Dr. Maria Antonieta Barrufet Fall, 2001

Page 42/91

PETE 609 - Module 6


Improved Water Flooding Processes

The following figure illustrates laboratory recoveries as a function of the capillary


number for two types of rock.

Figure 18 - Residual oil saturation as a function of Capillary Number and rock type.

Other names for the MP process are: detergent, surfactant, low tension, chemical, and
microemulsion flooding. The difference with alkaline flooding is that the surfactant is
injected, while in alkaline flooding it is generated in-situ.

Class Notes for PETE 609 Module 6


Author: Dr. Maria Antonieta Barrufet Fall, 2001

Page 43/91

PETE 609 - Module 6


Improved Water Flooding Processes

MP is the most complex EOR process from the view point of alternatives involved in its
design. In this section our emphasis is to present the fundamentals so that you can
understand the basis for design and reasons for how it performs.
Figure 19 shows an idealized sketch of a MP seque nce. The process can be applied
when the oil production from waterflood is at the expense of very high water-oil ratios.

The sequence involves:

Preflush - Brine is injected to change (usually to lower) the salinity of the


resident brine so that mixing with the surfactant will not cause to lower the
interfacial activity. Preflushes can be as high at 100% of the floodable pore
volume Vp of a reservoir. In some cases a sacrificial agent (cheap material) is
added to minimize the surfactant retention in the rock.

Micellar-Polymer Slug - This volume, ranging from 5 to 20% Vp in field


applications contains the main oil-recovery agent (the surfactant). Typical
surfactant concentrations range from 1 to 20% on a volumetric basis. Among
the types of MP slugs are: aqueous slugs A (surfactant dissolved in water),
oleic slugs O (surfactant dissolved in oil) and variations between these two
extremes. Other chemicals are necessary to design an optimum oil recovery.
The nature and purpose of these chemicals will be discussed later.

Mobility Buffer - This fluid is a dilute solution of a water-soluble polymer and


its purpose is to drive the MP slug and banked-up fluids to the production
wells.

Mobility Buffer Taper - This consists of a volume of brine containing a


concentration of polymer grading between the one of the mobility buffer at the
front and zero at the back. The gradual concentration decrease mitigates the
adverse mobility ratio between the mobility buffer and the chase water.

Chase Water - The objective of this is to reduce the costs of injecting polymer
continuously. If the taper and mobility buffer have been designed properly, the
MP slug will be produced before it is invaded by the chase water.

Class Notes for PETE 609 Module 6


Author: Dr. Maria Antonieta Barrufet Fall, 2001

Page 44/91

PETE 609 - Module 6


Improved Water Flooding Processes

Chase
Water

Taper

Mobility
Buffer

Slug

Preflush

Mobility buffer

Slug

Preflush

250-2500g/cm3
polymer

1-20% Surfactant

Electrolyte (Na+ , Ca++ ,


etc.)

0-1% Alcohol
Stabilizers
Biocide
0-100% Vpf

0-5% Alcohol
0-5% Cosurfactant
0-90% Oil

Sacrificial chemicals
0-100% Vpf

Polymer
5-20% Vpf

Figure 19 - Sketch of a Micellar-Polymer sequence.

The success of a MP flooding process depends upon meeting certain criteria.


First, the surfactant slug must be propagated in its interfacial active mode. This is
accomplished through the chemical formulation steps.
Second, the amount of surfactant injected mus t be enough to overcome the retention by
the porous media. This is accomplished by using some sacrificial agents, scale-up
studies, laboratory experiments, and numerical simulation.
Third, the MP displacement must be designed such that dissipation due to dispersion
and channeling are minimized.

Surfactants Used
To understand the role of the surfactants in MP flooding, we will define their physical
properties, how they affect phase behavior, and how phase behavior and interfacial
tension are correlated.

Class Notes for PETE 609 Module 6


Author: Dr. Maria Antonieta Barrufet Fall, 2001

Page 45/91

PETE 609 - Module 6


Improved Water Flooding Processes

Molecular Structure
A typical surfactant monomer is composed of a polar (hydrophilic) portion, and a nonpolar (lypophilic) portion. The entire monomer is sometimes called amphiphile because
of this dual nature.
Surfactants are classified in 4 groups based on their polar groups. These are:
Anionics: The monomer is associated with an inorganic metal (a cation, which is
usually sodium). In an aqueous solution the molecule dissociates into free
cations (positively charged), and the anionic monomer (negatively charged).
The solution is electroneutral, which means positive and negative charges
balance. Anionic surfactants are the most common in MP flooding because
they are good surfactants, relatively resistant to retention, stable, and can be
made relatively cheap.
Cationic: In this case the surfactant molecule contains a an inorganic anion to
balance the charge. In solution it ionizes into a positively charged monomer,
and the anion. Cationic surfactants are highly adsorbed by clays and
therefore have not much use in MP flooding.
Non-ionic: This class of surfactant does not have ionic bonds, but when
dissolved in aqueous solutions, exhibits surfactant properties mainly by
electronegativity contrasts among its constituents. Non-anionic surfactants
are much more tolerant to high salinities than anionic, but they are poorer
surfactants. The non-ionic surfactants are used extensively in MP floods
mainly as co-surfactants.
Amphoteric: This class of surfactant has not been used in oil recovery. They
contain aspects of two or more of the previous classifications.

Table 1 contains ranges where MP has been applied.

Class Notes for PETE 609 Module 6


Author: Dr. Maria Antonieta Barrufet Fall, 2001

Page 46/91

PETE 609 - Module 6


Improved Water Flooding Processes

Property

Range

Depth [ft]

350 - 4550 (107 - 1387 m)

Reservoir Temperature [F]

55 - 200 (12.8 - 93 C)

Porosity [%]

13 - 32

Permeability [md]

7 - 300+ {avg.}

Type of Reservoir

Unconsolidated to well cemented


sandstones, limestones

Formation Water [ppm TDS]

3000 - 160,000 (3000 - 160,000


mg/kg)

Hardness [ppm Ca, Mg, Fe]

25 - 5000 (25 - 5000 mg/kg)

Crude Gravity [API at 60F]

15 - 45 (0.965 - 0.801 g/cm3)

Crude Viscosity [cp]

3 - 31.7 (3 - 31.7 m Pa*s)

Crude Type

Aromatic-Paraffinic-Naphtenic

Table 1 - Range of oil field characteristics to which microemulsion flooding has been
applied.

Figure 20 shows some examples of these surfactants.

Class Notes for PETE 609 Module 6


Author: Dr. Maria Antonieta Barrufet Fall, 2001

Page 47/91

PETE 609 - Module 6


Improved Water Flooding Processes

Anionics

Cationics

Noionics

Amphoterics

Quaternary ammonium
Alkyl-, alkyl- aryl-,
organics, pyridinum,
acyl-, acylamindo- Aminocarboxylic
imidazonlinium,
acyl- aminepolyglycol,
acids
piperidinium, and
and polyol ethers
sulfononium compounds

Sulfonates,
Sulfates,
Carboxylates,
Phosphates

O
C
C

C
C

C
C

C
C

O-Na+

(a) Sodium dodecyl sulfate


C
C

C
C

C
C
C

C
C
C

C
C

O
S

O-Na +

(b) Texas No. 1 sulfonate


O
R

O-Na+

Figure 20 - Examples of surfactants for the different classifications. (Taken from Lake,
1992).

The following Table shows some typical properties of commercial sulfonates. Typical
molecular weights range from 350 to 450 kg/kmole. Lower MW has better water
solubilities. Part of the purchased surfactant is inactive in the sense that may contain
impurities for example unreacted oil from the sulfonation step and water from the
neutralization. In same calculations the concentration of surfactant is expressed at
Equivalent weights (which is the Mw/charge i.e. number of polar active sites).

Class Notes for PETE 609 Module 6


Author: Dr. Maria Antonieta Barrufet Fall, 2001

Page 48/91

PETE 609 - Module 6


Improved Water Flooding Processes

Table 2 - Properties of commercial surfactants (from Lake, 1992).

Critical Micelle Concentration


If an anionic surfactant is dissolved in an aqueous solution the surfactant dissociates
into a cation and a monomer. If the surfactant concentration is increased, the lypophilic
portions (moieties) of the surfactant begin to cluster among themselves to form
aggregates, or micelles, containing several monomers each. A plot of the surfactant
monomer concentration versus the total surfactant concentration, as indicated in Figure
21, shows a curve that begins at the origin, increases monotonically with init slope, then
levels off at the so called critical micelle concentration CMC. Above the CMC, all further
increases in surfactant concentration cause increases in the micelle concentration only.
CMCs are typically small, and the surfactant in oil applications is in the micellar state.
That is the reason of the name Micellar flooding. Typical CMC values are
10-5 10-4 kg-mol/m3. And the size of the micelles is 10-4 to 10-6 mm.

Class Notes for PETE 609 Module 6


Author: Dr. Maria Antonieta Barrufet Fall, 2001

Page 49/91

PETE 609 - Module 6


Improved Water Flooding Processes

Micelles

Surfactant Monomer
Concentration

Monomers

Critical Micelle
Concentration
(CMC)

Total Surfactant Concentration


Figure 21 - Surfactant monomer concentration versus total surfactant concentration.

When the surfactant solution contacts an oleic phase, the surfactant tends to
accumulate at the interface. The lypophilic tail dissolves in the oil phase, and the
hydrophilic end dissolves in the aqueous phase. The surfactant prefers the interface
over the micelle. Now it becomes clear the purpose of the dual nature of the surfactant
since its accumulation at the interface will lower the IFT between the oleic and the
aqueous phase. This interface blurs in the same manner as as do interfaces in vaporliquid-equilibrium (VLE) near a critical point. We need to design the surfactant to
maximize the solubility in this interface, however brines affect greatly the surfactant
behavior. Therefore we need to analyze the interactions surfactant-oil-brine. Depending
on the salinity, micelles may form either with water or oil as the external phase as
indicated in Figure 22.

Class Notes for PETE 609 Module 6


Author: Dr. Maria Antonieta Barrufet Fall, 2001

Page 50/91

PETE 609 - Module 6


Improved Water Flooding Processes

Surfactant-Brine-Oil Phase Behavior


Here we use a ternary diagram to indicate this behavior as indicated in Figure 23.
Conventionally the surfactant is indicated at the top of the diagram, the lower left
indicates the brine, while the lower right indicates the oil. Following Lakes (1992)
nomenclature we associate numbers to the phases and species, as indicated in Table
3.

WATER

OIL

WATER

WATER

(W)

OIL

MOLECULAR
DISPERSION
IN WATER

(S1 )

(S 2)

WATER
EXTERNAL

OIL
EXTERNAL

(O)
MOLECULAR
DISPERSION
IN OIL

Figure 22 - Different types of micelles structures.

Class Notes for PETE 609 Module 6


Author: Dr. Maria Antonieta Barrufet Fall, 2001

Page 51/91

PETE 609 - Module 6


Improved Water Flooding Processes

Figure 23 - Conventional representation of surfactant/oil/brine phase behavior. Salinity


is fixed.

Species

Concentration Unit

Phase

Water

Volume Fraction

Aqueous

Oil

Volume Fraction

Oleic

Surfactant

Volume Fraction

Microemulsion

Polymer

Weight percent or g/m3

Table 3 - Numbering of phases and species according to Lake, 1992.

Class Notes for PETE 609 Module 6


Author: Dr. Maria Antonieta Barrufet Fall, 2001

Page 52/91

PETE 609 - Module 6


Improved Water Flooding Processes

Type II (-) Behavior


At low brine salinity, a typical MP surfactant will exhibit good aqueous phase solubility
and poor oil phase solubility. Therefore, an overall composition near the brine -oil
boundary of the ternary plot will split into two phases: an excess oil phase which is
essentially pure oil, and a water external microemulsion phase that contains brine
surfactant, and some solubilized oil. This oil occupies the central core of the swollen
micelles. The tie lines within the two-phase envelope have a negative slope. The plait
point PR in this system is located closer to the oil apex. Any overall composition above
this envelope (also called binodal curve) is a single phase (all components are
miscible). Figure 24 is an schematic representation of this type of behavior.

Class Notes for PETE 609 Module 6


Author: Dr. Maria Antonieta Barrufet Fall, 2001

Page 53/91

PETE 609 - Module 6


Improved Water Flooding Processes

Figure 24 - Schematic representations of the Type II (-) system, from Lake, 1984.

Type II (+) Behavior


For high brine salinities, electrostatic forces decrease the surfactant solubility in the
aqueous phase drastically. An overall composition within the two-phase region will now
split into an excess brine phase and a microemulsion phase (oil external) which
contains most of the surfactant and some solubilized brine. The brine is solubilized
through the formation of inverted swollen micelles, with brine globules at their cores.
Class Notes for PETE 609 Module 6
Author: Dr. Maria Antonieta Barrufet Fall, 2001

Page 54/91

PETE 609 - Module 6


Improved Water Flooding Processes

The plait point PL, is now located closer to the brine apex. The tie lines here have a
positive slope. Figure 25 illustrates this concept.

Figure 25 - Schematic representation of high-salinity type Ii (+) system, from Lake 1984.

Type III Behavior


At intermediate salinities there must be a continuous change between Type II (-) and
Type II (+) systems. The intuitive change of a plait point migration to the top and
Class Notes for PETE 609 Module 6
Author: Dr. Maria Antonieta Barrufet Fall, 2001

Page 55/91

PETE 609 - Module 6


Improved Water Flooding Processes

horizontal tie lines is incorrect. There is no salinity for which the solubilities of the
surfactant for the brine-oil and oil-rich phases are exactly the same. What occurs under
these conditions is the formation of three phases (therefore the name Type III).
An overall concentration within the three phase region separates into excess oil and
brine phases as in Type II (-) and Type II(+) environments, and into a microemulsion
phase whose composition is represented by an invariant point. Now there are two IFTs
between the microemulsion and the oil
brine

( mo ) and between the microemulsion and the

( mw ). Figure 26 illustrates this behavior.

Class Notes for PETE 609 Module 6


Author: Dr. Maria Antonieta Barrufet Fall, 2001

Page 56/91

PETE 609 - Module 6


Improved Water Flooding Processes

Figure 26 - Schematic representation of optimal-salinity type III system, from Lake,


1984.

Figure 27 illustrates the three types of behaviors just described.

Class Notes for PETE 609 Module 6


Author: Dr. Maria Antonieta Barrufet Fall, 2001

Page 57/91

PETE 609 - Module 6


Improved Water Flooding Processes

Figure 27 - Simplified version of phase diagrams for a microemulsion system.

Changes in Phase Environments


Figure 28 shows the entire progression of phase environments from Type II (-) to Type
II (+), over the type III salinity range. The invariant point M migrates from near the oil
apex to near the brine apex before disappearing at the critical tie line. As the migration
takes place, the surfactant concentration in the microemulsion phase goes through a
minimum, where the brine-oil ratio at the invariant point becomes one. Type III
environments are those at which all the IFTs are the lowest.
There are parameters other than salinity that which can cause phase environment
shifts. In general, changing any condition that enhances the surfactants oil solubility will
cause a shift from type II (-) to type II (+). Some of these include

Decreasing the temperature

Increasing the surfactant molecular weight

Decreasing the oil specific gravity

Increasing the concentration of high molecular weight alcohols.

Class Notes for PETE 609 Module 6


Author: Dr. Maria Antonieta Barrufet Fall, 2001

Page 58/91

PETE 609 - Module 6


Improved Water Flooding Processes

Decreasing the surfactants oil solubility will cause the reverse change.

Figure 28 - Pseudoternary or 'tent' diagram representation of micellar-polymer phase


behavior, from Lake, 1984.

Phase Behavior and IFTs


IFTs vary with the types and concentration of surfactant, co-surfactant, electrolytes, oil,
polymer, and temperature. However, one of the most significant advances in MP
technology is that all IFTs correlate directly with the MP phase behavior. One of the
biggest advantages of this is that difficult measurements of IFTs can be supplanted by
easier phase behavior measurements.

Solubilization Parameters
Lake (1992) defines the volume fractions of oil, brine, and surfactant in the
microemulsion phase as: V o, Vw , Vs respectively. The solubilization parameters
between the microemulsion-oleic phase for type II(-) and type III behavior, and between
the microemulsion-aqueous phases for type II (+) and type III are defined as.

Class Notes for PETE 609 Module 6


Author: Dr. Maria Antonieta Barrufet Fall, 2001

Page 59/91

PETE 609 - Module 6


Improved Water Flooding Processes

Smo =

Vo
Vs

or

C 23
C 33

(30)

Vw
Vs

or

C13
C 33

(31)

and

Smw =

The IFTs between the corresponding phases are only functions of theses solubilization
parameters. Other nomenclatures in use, particularly by Lake (1992), are introduced
here to understand the following figures taken from his book.
He uses two numbered subscripts one defines the component and the other the phase.
For example

Cij

i = component
j = phase
as indicated in Table 3 : (1) stands for aqueous phase and brine, (2) stands for oleic
phase and oil, and (3) stands for surfactant and microemulsion.

The solubilization parameters are defined as the ratios indicated in the following figure

Class Notes for PETE 609 Module 6


Author: Dr. Maria Antonieta Barrufet Fall, 2001

Page 60/91

PETE 609 - Module 6


Improved Water Flooding Processes

Figure 29 - Correlation of solubilization parameters with interfacial tensions (from


Glinsmann, 1979; and Lake, 1992).

Optimal Salinity
The behavior of the IFTs versus salinity is indicated in Figure 30 to Figure 35.
The optimum salinity is when both IFTs are at its minimum, and this is normally
achieved for type III behavior. Since optimal phase behavior salinity translates into a
maximum oil recovery, the objective is to design the proper slug-brine-surfactant
formulation that will achieve this salinity insitu. Dilution effects and adsorption must be
taken into account as well.
Class Notes for PETE 609 Module 6
Author: Dr. Maria Antonieta Barrufet Fall, 2001

Page 61/91

PETE 609 - Module 6


Improved Water Flooding Processes

Unfortunately, for most commercially attractive surfactants in most MP candidate


reservoirs, the optimal salinity is lower than the resident brine salinity. Optimal salinities
can be raised by adding to the slug any chemical, a co-surfactant, which increases the
primary surfactants brine solubility. The addition of co-surfactants to the MP slug
normally increases the optimal IFT as well.

Figure 30 - Interfacial tensions and solubilization parameters (from Reed and Healy,
1977; and Lake, 1992).
Class Notes for PETE 609 Module 6
Author: Dr. Maria Antonieta Barrufet Fall, 2001

Page 62/91

PETE 609 - Module 6


Improved Water Flooding Processes

Figure 31 - Correlation of phase volume and IFT behavior with retention and oil
recovery (from Glinsmann, 1979; and Lake, 1992).

Class Notes for PETE 609 Module 6


Author: Dr. Maria Antonieta Barrufet Fall, 2001

Page 63/91

PETE 609 - Module 6


Improved Water Flooding Processes

Figure 32 - Phase volume diagrams (salinity scans) at three water-oil ratios (from
Englesen, 1981).

Class Notes for PETE 609 Module 6


Author: Dr. Maria Antonieta Barrufet Fall, 2001

Page 64/91

PETE 609 - Module 6


Improved Water Flooding Processes

Figure 33 - Microemulsion as a function of salinity (from Jones, 1981).

Class Notes for PETE 609 Module 6


Author: Dr. Maria Antonieta Barrufet Fall, 2001

Page 65/91

PETE 609 - Module 6


Improved Water Flooding Processes

Figure 34 - Schematic diagram of the effect of increasing salt concentration on phase


volumes of multiphase microemulsion system.

Class Notes for PETE 609 Module 6


Author: Dr. Maria Antonieta Barrufet Fall, 2001

Page 66/91

PETE 609 - Module 6


Improved Water Flooding Processes

Figure 35 - Interfacial tension versus salinity for multiphase microemulsion system.

The velocity of the microemulsion can be tailored by proper additions of alcohols (or cosurfactants). Figure 36 and Figure 37 show the effect of alcohol concentration upon
viscosity.

Class Notes for PETE 609 Module 6


Author: Dr. Maria Antonieta Barrufet Fall, 2001

Page 67/91

PETE 609 - Module 6


Improved Water Flooding Processes

Figure 36 - Viscosity of a microemulsion composed of 12 vol. % petroleum sulfonate,


24.9 vol. % water, and 63.1 vol. % hydrocarbon versus added cosurfactant, isopropyl
alcohol.

Class Notes for PETE 609 Module 6


Author: Dr. Maria Antonieta Barrufet Fall, 2001

Page 68/91

PETE 609 - Module 6


Improved Water Flooding Processes

Figure 37 - Viscosity and stability limits of a microemulsion composed of 11.7 vol. %


petroleum sulfonate, 65.5 vol. % water, and 22.8 vol. % hydrocarbon versus added
cosurfactant, p-hexanol or p-pentanol.

Surfactant Retention
Surfactant retention is probably the most significant barrier to the commercial
application of MP flooding . The surfactant should be designed for good selectivity to the
oil/water interfaces, but poor selectivity to the fluid/solid interfaces.
Some possible mechanisms that explain the retention of surfactants in the rock include:
1. Adsorption on metal oxide surfaces. While the surfactant concentration is lower than
the CMC it will adsorb through hydrogen bonding and ionic bonding to cationic
surface sites.
2. Hard brines cause the formation of surfactant-divalent complexes that have low
solubility in brine. These complexes precipitate out of solution causing retention.
However, when the surfactant encounters the oil phase this effect is lessened due to
the solubility of the surfactant in the oleic phase.

Class Notes for PETE 609 Module 6


Author: Dr. Maria Antonieta Barrufet Fall, 2001

Page 69/91

PETE 609 - Module 6


Improved Water Flooding Processes

3. In reservoirs that contain clays the surfactant can exchange the cations with the
clays and become attached to the surface. The addition of co-surfactants reduces
the density of surfactant molecules at the surface.
A useful way to estimate the volume of surfactant required for a MP slug is defined by
Lake (1992).

1 r as
Ds =

s Cs

(32)

where as is the surfactant retention in mg/g,


concentration in the MP slug, and

and

is the porosity, Cs is the surfactant

s are the rock and surfactant density

respectively.

Ds expresses the volume of surfactant retained at its injected concentration as a


fraction of the pore volume Vp.
For optimal surfactant usage, the volume of surfactant injected should be large enough
to contact the entire Vp, but small enough to prevent excessive production of the
surfactant. Besides wasting an expensive chemical, this could lead to severe production
of emulsions.
According to Lake (1992), the MP slug size should be equal to or a little bit larger (510%) than Ds. The volume of surfactant needed is the product of the slug size and the
surfactant concentration.

Screening Tests for Micellar-Polymer Floods


Screening tests for micellar-polymer suitability vary. The laboratory test sequence
should culminate in reservoir condition floods (or reservoir temperature floods) using the
formulated surfactant and polymer to measure oil mobilized and displaced.

Crude Oil Characterization Tests


Crude oil acid number, equivalent molecular weight and equivalent alkane carbon
number (EACN) are determined. The EACN identifies the crude oil (which is a complex
Class Notes for PETE 609 Module 6
Author: Dr. Maria Antonieta Barrufet Fall, 2001

Page 70/91

PETE 609 - Module 6


Improved Water Flooding Processes

mixture of hydrocarbons) as having equivalent behavior to a single pure component


such as hexane or octane in the presence of surfactants. Knowing the pure component
which matches crude oil behavior yields access to correlations which assist in selecting
surfactants and co-surfactants to mix with the crude oil and water to form stable
microemulsion slugs. Interfacial tension measurements are part of the EACN
evaluation.

Microemulsion Formulation Tests


Complex theories exist to assist in formulations. Reservoir oil and brine are mixed with
surfactant and co-surfactants in bench tests to identify stable microemulsions that
exhibit middle phase (or Type III) behavior. Brine salinity is varied and oil and water
solubility in the middle phase is observed to define optimum formulation.
The co-surfactant concentration is varied to adjust microemulsion viscosity. In some
systems polymer is added to adjust slug viscosity.

Mobility Control Polymer Tests


Measured data include viscosity, screen factors (for polyacrylamides) and filtration tests
(for polysaccharides), injectivity behavior, and mobility (Resistance Factors) in the
formation rock as a function of polymer concentration and rate.
Analyses of formation water, polymer make up water, and drive water are required.
Polymer compatibility with these waters as well as with the microemulsion slug must be
evaluated.

Rock Property Tests


The lithology of reservoir rock influences surfactant floods. Sandstone reservoirs are
the best candidates. Limestones are poor as they adsorb excessive surfactant.
Anhydrite (CaSO4) associated with carbonates can increase the calcium content in flood
waters and reduce surfactant and polymer effectiveness.
Clays exhibit cation exchange capacity and act as ionic exchange media causing rockwater reaction and reduced injectivity. Calcium and magnesium released from clays to
the flood water reduce surfactant effectiveness. Surfactants are also more readily
adsorbed on high surface area clays than on normal silica surfaces, and loss varies with
clay type.

Class Notes for PETE 609 Module 6


Author: Dr. Maria Antonieta Barrufet Fall, 2001

Page 71/91

PETE 609 - Module 6


Improved Water Flooding Processes

Sacrificial agents such as sodium carbonate, sodium silicate in a high pH (>10) solution,
or certain organic compounds have been proposed to reduce surfactant absorption and have been supported by laboratory tests.
X-ray, thin section and SFM tests identify rock minerals (including clays) and location;
rock-water sensitivity tests and relative permeability data are needed for mobility design
of the micellar flood.

Core Floods
Micellar fluids and polymer drive slugs are injected in radial or linear flood tests. Initial
tests are made on Berea sandstone, with final evaluation utilizing actual reservoir rock
at reservoir temperature.
Cores are water flooded to residual oil, and selected micellar and polymer fluids are
subsequently injected to define increased oil recovery and reduced residual oil.
The example illustrates the increase in oil cut in the produced fluids following surfactant
injection, and the tertiary oil recovery as a percentage of residual oil-in-place.
The following figure illustrates the sequence of tests performed to design a MP flood

Class Notes for PETE 609 Module 6


Author: Dr. Maria Antonieta Barrufet Fall, 2001

Page 72/91

PETE 609 - Module 6


Improved Water Flooding Processes

Surfactant-Polymer
Floods
Reservoir Rock
Characterization

Crude Oil
Characterization

Thin Sections for


Mineral Content

Crude Oil
Equivalent
Molecular Wt.

X-Ray for Clay


Identification

If Tests to Define
Equivalent Alkane
Carbon No.

SEM for Clay


Location
Relative Permeability

Microemulsion
Formulation
and Core Testing
Phase Behavior of
Crude+Brine+Surfactant+
Cosurfactant (middle
phase desired)
Microemulsion
viscosity vs.
Cosurfactant
composition
& concentration
Polymer Buffer
Selection
Micellar Adsorption

Drainage

Imbibition
Core Displacement
Studies to Define Sorr

Capillary Number

Water Permeability

Waterflood core
to residual oil
Inject micellar
slug
Inject polymer

Class Notes for PETE 609 Module 6


Author: Dr. Maria Antonieta Barrufet Fall, 2001

Page 73/91

PETE 609 - Module 6


Improved Water Flooding Processes

Alkaline Flooding
Alkaline flooding, is a high pH chemical EOR method which has many similarities with
micellar flooding. The difference is that in micellar flooding the surfactant is injected,
while in alkaline (or caustic) flooding the surfactant is generated in situ.
High pHs indicates large concentrations of the hydroxide anions

COH .

The pH of an

ideal aqueous solution is defined as

pH = log (CH + )

(33)

where the concentration of protons (hydrogen ions) is expressed in kg-moles/m3. A pH


of seven is neutral, lower than that is acid and higher than that is alkaline. As the
concentration ofC

OH

increases, the concentration of C

H+

must decrease since the two are related through the dissociation of water

kw =

COH CH +

(34)

CH2O

and the water concentration in an aqueous phase is nearly constant. This illustrates that
there are two methods of increasing the pH of a reservoir:

(1) By dissociation of a hydroxyl containing species such as NaOH, or KOH.


(2) By adding chemicals that will bind withC

H+

Many chemicals could be used to generate high pHs, but the most common are:
sodium hydroxide, sodium carbonate, ammonia, and sodium orthosilicate. Sodium

Class Notes for PETE 609 Module 6


Author: Dr. Maria Antonieta Barrufet Fall, 2001

Page 74/91

PETE 609 - Module 6


Improved Water Flooding Processes

hydroxide or sodium carbonate dissociate in water generating C

OH

according to the

following dissociation reactions:

Na2CO3 CO3
CO3

+ 2Na

+ 2H2O H2CO3 + 2OH

NaOH Na + + OH

(35)
(36)

The single sided arrow in the last reaction indicates that the reaction is favored to the
right. While two sided arrows indicate that the reaction is reversible. That means the
reaction could go either way depending upon the pH.
High-pH chemicals have been used in field applications in concentrations of up to 5 wt
% (injected pHs 11 to 13) and with slug sizes of up to 0.2 PV.

Surfactant formation
OH - by itself is not a surfactant since the absence of a lypophilic tail makes it exclusively
water soluble. However, if the oil contains acidic hydrocarbon components (HAo), some
of it may partition into the aqueous phase as indicated in Figure 38.

Class Notes for PETE 609 Module 6


Author: Dr. Maria Antonieta Barrufet Fall, 2001

Page 75/91

PETE 609 - Module 6


Improved Water Flooding Processes

H 2O
OH-

ROCK

A-

Na+
HAo

OIL

HAo
HAw

NaOH
A- + H+

H2O

Figure 38 - Schematic of alkali recovery process.

We assume that the acid species in the oil is represented by a generic single
component named HAo. This acid component will not be soluble in an aqueous phase
with neutral pH (i.e., 7). However if the pH is increased with a caustic solution the acid
will be extracted from the oil to the aqueous phase.
The exact nature of the acidic component is unknown, b ut it is probably highly
dependent on the crude oil type. The deficiency of protons (high PH) in the aqueous
phase will promote the chemical reactions to the right. The anionic species A - is a
surfactant with many of the properties described in MP flooding.
If no acidic species are present in the crude, no surfactant can be generated. Therefore
to determine the oil characteristics needed for alkaline flooding we must characterize its
acidity. The attractiveness of an oil for alkaline flooding is given by its acid number.
The acid number is the milligrams of potassium hydroxide (KOH) needed to neutralize
one gram of crude oil. To make this measurement, the crude oil is extracted with water
until the acidic species HAo is removed. The aqueous phase is then brought to neutral
pH=7 by adding KOH.
For a meaningful value, the oil must be free of acidic additives such as corrosion
inhibitors and acidic gases such as H2S and CO2.
Class Notes for PETE 609 Module 6
Author: Dr. Maria Antonieta Barrufet Fall, 2001

Page 76/91

PETE 609 - Module 6


Improved Water Flooding Processes

A good alkaline flooding candidate will have an acidic number of 0.5 mg/g or greater.

Displacement Mechanisms
We postulate that the following dissociation reactions may take place with sodium
hydroxide

HAo + NaOH NaA + H2O

(37)

NaA Na + + A

(38)

where A- is a surface active ingredient.

The following reaction represent the extraction of the acidic component from the oil
phase into the aqueous phase.

HAo HAw

(39)

The concept is analogous to vapor-liquid-equilibria.component distribution. The


susbscripts o and w indicate that the acidic component partitions into the oleic and
aqueous phase respectively. A material balance for the acidic component HA, must take
into account the quantities present in the oil phase and in the aqueous phase. If the pH
of the aqueous media is high, this partition will be more favored to the aqueous phase.
The acid distribution between oil and water phases can be quantified through the
distribution coefficient (analogous to k-values).

kD =

CHAo
CHAw

Class Notes for PETE 609 Module 6


Author: Dr. Maria Antonieta Barrufet Fall, 2001

(40)

Page 77/91

PETE 609 - Module 6


Improved Water Flooding Processes

This coefficient is a measure of the solubility of acid in the water phase.


Once the acidic component is in the aqueous solution it may hydrolyze (decompose in
water phase) as indicated by the following equation

HAw H + + A

(41)

The extent of this dissociation is determined by the acid dissociation constant

kA =

C H + C A

(42)

CHAw

This constant controls the pH range for which the surfactant hydrolysis occurs.
Our objective is to find out what is the concentration of the surfactant component in the
aqueous phase and see how this surfactant concentration may alter the fractional flow
curve.
To do this we use the dissociation constants (Equations,(40) and (42)), the water
dissociation constant, and the electro neutrality condition.
Since water concentration is essentially constant, the water dissociation constant can
be expressed as.

kw = kw ' CH2O = CH + COH

(43)

The electro neutrality condition states that positive and negative charges must balance

CNa+ + C H + = CA + COH

Class Notes for PETE 609 Module 6


Author: Dr. Maria Antonieta Barrufet Fall, 2001

(44)

Page 78/91

PETE 609 - Module 6


Improved Water Flooding Processes

Using the equations described above we find an analytical expression to evaluate the
surfactant concentration (Equation (45)). This is left as an exercise.

C A

C
1 CNa +
Na+
=
+
2 1+ k
1+ k

CHAo
CHAo

4k C
+ w HAo

k

(45)

where

k=

kD k w
kA

(46)

At high pH, the concentration of acidic species in the oil phase CHAo is very small and
the concentration of protons may be neglected in the electroneutrality condition
expressed in Equation (44)- Thus Equation (45) can be simplified and the following
expression results

C A =

CNa+
k
1+
CHAo

(47)

Therefore, to determine the concentration of surfactant in the aqueous phase we need


to have values, either experimental or theoretical, for the distribution constants

( kD ,k w ,k A ) .

You can show that k is the inverse of the overall equilibrium constant of Equation (37).

Class Notes for PETE 609 Module 6


Author: Dr. Maria Antonieta Barrufet Fall, 2001

Page 79/91

PETE 609 - Module 6


Improved Water Flooding Processes

Some typical values for these dissociation constants are,

kw = 5 10 14 mol / liter
kD = 10 4
k A = 10 10 mol / dm 3
These constants depend upon temperature.

In addition to the liquid reactions (oil/alkali), we may find reactions between the rock and
the caustic species. The hydroxyl can interact with the rock in several ways depending
upon its mineralogy. For example silicates may dissolve irreversibly, and we may have
precipitation of insoluble solids, although we will not consider these rock dissolution
effects to simplify the model. However, we cannot ignore mineral-base exchange sites
which may have dramatic consequences.

A typical chemical reaction taking place at Liquid-Solid interface is the Na + /H+


reversible exchange

(M ) H + Na

+ OH MNa + H2O

This exchange causes NaOH retention. Figure 39 shows the time

(48)

expressed in pore

volumes for caustic to elute from a Wilmington sand core at various injection caustic
concentrations. It can be observed that lower injected caustic concentrations (low pH)
take longer to breakthrough the core. These predictions have been verified
experimentally.

Class Notes for PETE 609 Module 6


Author: Dr. Maria Antonieta Barrufet Fall, 2001

Page 80/91

PETE 609 - Module 6


Improved Water Flooding Processes

14
Calculated NaOH breakthrough
T = 52.5 oC
NaCl = 1% (wt)

B, Pore Volumes

12
10
8
6
4
2
0
10

11

12

13

14

pH
Figure 39 - Time ( B ) in PV for caustic to breakthrough as a function of pH for
Wilmignton sands.

Displacement Model
The chemical reactions taking place between oil and alkali can be incorporated into a
displacement model. The assumptions here are:

Model a linear, homogeneous porous medium

Temperature is constant

Uniform initial saturation and composition

Mobile, immiscible, and incompressible oil and water phases

Local and instantaneous equilibrium

Negligible dispersion and capillary pressure forces

Stable displacement front without viscous fingering

No emulsion formation

Class Notes for PETE 609 Module 6


Author: Dr. Maria Antonieta Barrufet Fall, 2001

Page 81/91

PETE 609 - Module 6


Improved Water Flooding Processes

Continuous alkali injection

We have 6 unk nowns and two liquid phases (oleic and aqueous) and three equilibrium
relation constants.
These unknowns are,

HAo ,HAw , A ,Na + ,OH ,H +


Define a dimensionless time and a dimensionless distance as

Dimensionless time

ut
L

Dimensionless distance

% =

Local conservation of water demands,

Sw
f
= w

(49)

and a local sodium ion balance is

( SwCNa + )

( fw CNa+ )
1 ( nNa + )
+
=

(50)

The sodium ion exists in the aqueous phase, where their accumulation and convection
must be accounted for. Additionally, the ions exchange with the rock as indicated in
Figure 38.

Class Notes for PETE 609 Module 6


Author: Dr. Maria Antonieta Barrufet Fall, 2001

Page 82/91

PETE 609 - Module 6


Improved Water Flooding Processes

The relation

nNa + = nOH [ =]

moles absorbed
solid rock volume

(51)

gives the adsorption amount of sodium or hydroxide in moles per solid-rock volume. For
the local equilibrium assumption, we have that sodium interchange with rock

( nNa+ )

n
= OH
C
OH

CNa +

(52)

The exchange isotherm for caustic is concave to the abscissa, thus a shock front will
develop when injecting alkali. The concentration velocity of this shock is regulated by
the isotherm chord and it is represented by a pore volume delay parameter.

1 nOH
=

COH

(53)

A differential material balance for the acid species is

Sw (C + CHA ) + ( 1 Sw ) CHA =
A
w
o

fw ( CA + CHAw ) + ( 1 fw ) CHAo
%

Class Notes for PETE 609 Module 6


Author: Dr. Maria Antonieta Barrufet Fall, 2001

(54)

Page 83/91

PETE 609 - Module 6


Improved Water Flooding Processes

To respond to the question of how does the A- generated influence fw , this is


accomplished by altering the residual oil saturation.
The following model can be proposed to describe the change in

Sor

with

C A

C
SOr = c1 + c2 exp A
c3

(55)

where the constants are determined experimentally.

The following graph illustrates the change in residual oil saturation using certain values
for the constants c1 to c3.

0.4

Sor

0.3

0.2
0.1

0.0
0.00

Sor (CA-) = c1 + c2 exp(-CA- / c3)


c1 = 0.05
c2 = 0.35
c3 = 1.18 x 10-3
mol/dm3

0.02

0.04

0.06

0.08

0.10

CA- [mol/dm3]

Figure 40 - Assumed variation of residual oil saturation with surfactant concentration.

To draw the fractional flow curve we may propose the following relative permeability
models

Class Notes for PETE 609 Module 6


Author: Dr. Maria Antonieta Barrufet Fall, 2001

Page 84/91

PETE 609 - Module 6


Improved Water Flooding Processes

k rw = krow S n

(56)

k ro = k roo ( 1 S )

where

S=

Sw Swc
1 Sor Swc

(57)

Figure 41 shows the effect of a given alkali concentration in the fractional flow curve.
The power indexes for the water and oil relative permeability have been taken as 2 and
2.5 respectively.

1
Alkaline Flood
CA- = 5.5(10-4) mol/dm3

0.8
Waterflood
CA- = 0

fw

0.6

0.4
0.2

1-Sor(CA-)

Swc

0
0

0.2

0.4

Sw

0.6

0.8

Figure 41 - Fractional flow for a waterflood and an alkaline flood.

Class Notes for PETE 609 Module 6


Author: Dr. Maria Antonieta Barrufet Fall, 2001

Page 85/91

PETE 609 - Module 6


Improved Water Flooding Processes

Displacement Calculations
This section outlines the procedure to make displacement calculations, and shows the
results obtained for the proposed model. A detailed explanation can be found in
deZabala et al. (1982).
There are 3 PDE 's (water, sodium, and acid balances) which are solved by using the
coherence technique which transforms PDE's to algebraic equations
Suggested reading on this topic are the papers by Helfferich, SPEJ, February, 1981 and
by Hirasaki, SPEJ, April, 1981
Figure 42, and Figure 43 show the saturation and chemical profiles for an alkaline flood
at two different times.

M
pH

0.8
1-Sor(CA-)
SR

= 0.25 PV
= 2.0
= 12.75
= 1.5

0.6
Waterflood

Sw
0.4

CHA o = 0. 05mol/dm (2.8 )


3

0.2

0.0

CA = 5 .5 10 4 mol/dm 3
5.0
~

1.0

13
12
11 pH
10
9
8

Figure 42 - Saturation and chemical profiles for a secondary alkaline flood at = 0.25
PV of displacement.

Class Notes for PETE 609 Module 6


Author: Dr. Maria Antonieta Barrufet Fall, 2001

Page 86/91

PETE 609 - Module 6


Improved Water Flooding Processes

= 5.0 PV
M = 2.0
pH = 12.75
= 1.5

0.75

0.55

Sw

Waterflood
OH-

0.35

13
12
11 pH
10
9
8

CHAo = 0.05 mol/dm (2.8)


CA- = 5.5(10-4) mol/dm 3

0.15
0.0

0.2

0.4

0.6

0.8

1.0

Figure 43 - Saturation and chemical profiles for a second alkaline flood at = 5 PV.

Additional References for Module 6


Akstinat, M.H., "Surfactants for EOR Process in High-Salinity Systems, Product
Selection and Evaluation, "Enhanced Oil Recovery, Elsevier Scientific, New York, New
York, 1981.

Bragg, J.R., Gale, W.W., McElhannon, W.A., Davenport, O.W., "Loudon Surfactant
Flood Pilot Test," SPE 10862, presented at the Third Society of Petroleum Engineers
Symposium on EOR, Tulsa, Oklahoma, 1982.

Class Notes for PETE 609 Module 6


Author: Dr. Maria Antonieta Barrufet Fall, 2001

Page 87/91

PETE 609 - Module 6


Improved Water Flooding Processes

Bunge, A.L., and Radke, C.J. "Migartion of Alkali ne Pulses in Reservoir Sands," SPEJ,
Vol 22, (December 1982), 998-1012.

Clampitt, R.L., and Reid, T.B., "An Economic Polymerfood in the North Burbank Unit,
Osage County, Oklahoma," SPE 5552,presented at the 50th Annual Fall Technical
Conference and Exhibition of the SPE, Dallas, Texas, Sept. 28, Oct.1, 1975.

DeZabala, E.F, Vislocky, J.M, Rubin, E., Radke, C.J., "A Chemical Theory for Linear
Alkaline Flooding" SPEJ, April 1982 pp 245-258.

GIinsmann, G.R., "Surfactant Flooding with Microemulsions Formed In-situ - Effect of


Oil Characteristics," SPE 8326, presented at the Society of Petroleum Engineers 5th
Annual Technical Conference and Exposition, Las Vegas, Nevada, 1975.

Glover, C.J., Puerto, M.C., Maerker, J.M., and Sandvik, E.I., "Surfactant Phase
Behavior and Retention in Porous Media," Society of Petroleum Engineers Journal, 19,
(1979) 183-193.

Gogarty, W.B., Meabon, H.P., and Milton, H.W., Jr., "Mobility Control Design for
Miscible-type Waterfloods Using Micellar Solutions," Journal of Petroleum
Technology, 22, (1970), pp 141-147

Graue, D.J. and Johnson, C.E., "Field Trial of Caustic Flooding Process," JPT,
(December 1974). Pp 1353-1358.

Gupta, S.P., "Compositional effects on displacement mechanisms of the micellar fluid


injected in the Sloss Field Test," SPE 8827, presented at the First Joint Society of
Petroleum Engineers/Department of Energy Symposium on Enhanced Oil Recovery,
Tulsa, Oklahoma, 1980.

Class Notes for PETE 609 Module 6


Author: Dr. Maria Antonieta Barrufet Fall, 2001

Page 88/91

PETE 609 - Module 6


Improved Water Flooding Processes

Healy, R.N., and Reed, R.L., "Physiochemical Aspects of Microeumulsion Flooding,"


Society of Petroleum Engineers Journal, 14, (1974) pp 491-501.

Healy, R.N., and Reed, R.L., and Stenmark, D.G., "Multiphase Microemulsion
Systems," Society of Petroleum Engineers Journal, 16, (1976) 147-160.

Hirasaki, G.J., and Pope, G.A., "Analysis of Factors Influencing Mobility and Adsorption
in Flow of Polymer Solution through Porous Media," Society of Petroleum
Engineers Journal, 14, (1974) pp.337-346.

Holm, L.W., "Design, Performance and Evaluation of the Uniflood Micellar-Polymer


Process- Bell Creek Field," SPE 11196, presented at the 57th Annual Fall Technical
Conference and Exhibition of the Society of Petroleum Engineers, New Orleans,
Louisiana, 1982.

Jennings, R.R., Rogers, J.H., and West, T.J., "Factors Influencing Mobility Control by
Polymer Solutions," Journal of Petroleum Technology, 23, (1971) 391-401.

Maerker, J.M., "Mechanical Degradation of Partially Hydrolyzed Polyacrylamide


Solutions in Unconsolidated Porous Media," Society of Petroleum E ngineers Journal,
16, (1976) 172-174.

Lake L.W, Enhanced Oil Recovery, (1989). Chapters 8 and 9. Edited by Prentice Hall.

Lake, L.W., Stock, L.G., and Lawson, J.B., "Screening Estimation of Recovery
Efficiency and Chemical Requirements for Chemical Flooding," SPE 7069, presented at
the Society of Petroleum Engineers Fifth Symposium on Improved Methods for Oil
Recovery, Tulsa, Oklahoma, 1978.

Class Notes for PETE 609 Module 6


Author: Dr. Maria Antonieta Barrufet Fall, 2001

Page 89/91

PETE 609 - Module 6


Improved Water Flooding Processes

Lake, L.W., and Helfferich, F., "Cation Exchange in Chemical Flooding, Part II- The
Effect of Dispersion, Cation Exchange, and Polymer/Surfactant Adsorption on Chemical
Flood Environment," Society of Petroleum Engineers Journal, 18, (1978) 435-441.

Lake, L.W., and Pope, G.A., "Status of Micellar-Polymer Field Tests," Petroleum
Engineers International, 51, (1979) 38-69.

Nelson, R.C., and Pope, G.A., "Phase Relationships in Chemical Flooding,:Society of


Petroleum Engineers Journal, 18, (1978) pp 325-338.

Nelson, R.C., "Effect of Live Crude on Phase Behavior and Oil-Recovery Efficiency of
Surfactant Flooding Systems," Society of Petroleum Engineers Journal, 23, No. 3,
(1983) pp. 501-520.
Winsor, P.A., Solvent Properties of Amphiphilic Compounds. Edited by Butterworths,
London, 1954.

Savins, J.G., "Non-Newtonian Flow Through Porous Materials," I & EC, 61 (10), (1969)
19.

Seright, R.S., "The Effects of Mechanical Degradation and Viscoelastic Behavior


on Injectivity of Polyacrylamide Solutions," Society of Petroleum Engineers Journal,
vol. 23 (3), (June 1983) pp. 475-485

Wellington, S.L., 1980, "Biopolymer Solution Viscosity Stabilization Polymer


Degradation and Antioxidant Use," SPE 9296, presented at the 55th Annual Fall
Technical Conference and Exhibition of the SPE, Dallas, Texas, Sept. 21-24, 1980.

Willhite, G.P. 1986, Waterflooding Monograph. SPE Textbook series Vol. 3.

Class Notes for PETE 609 Module 6


Author: Dr. Maria Antonieta Barrufet Fall, 2001

Page 90/91

PETE 609 - Module 6


Improved Water Flooding Processes

Class Notes for PETE 609 Module 6


Author: Dr. Maria Antonieta Barrufet Fall, 2001

Page 91/91

You might also like