You are on page 1of 14

SPE Number 154045

Improving Heavy Oil Recovery Using An Enhanced Polymer


System
P. Fletcher, S. Cobos, C. Jaska, SPE, J. Forsyth, SPE, M. Crabtree, Oilflow Solutions Inc., Canada,
N. Gaillard, SPE, C. Favéro, SPE, SNF SAS, Andrézieux, France

Copyright 2012, Society of Petroleum Engineers

This paper was prepared for presentation at the Eighteenth SPE Improved Oil Recovery Symposium held in Tulsa, Oklahoma, USA, 14–18 April 2012.

This paper was selected for presentation by an SPE program committee following review of information contained in an abstract submitted by the author(s). Contents of the paper have not been
reviewed by the Society of Petroleum Engineers and are subject to correction by the author(s). The material does not necessarily reflect any position of the Society of Petroleum Engineers, its
officers, or members. Electronic reproduction, distribution, or storage of any part of this paper without the written consent of the Society of Petroleum Engineers is prohibited. Permission to
reproduce in print is restricted to an abstract of not more than 300 words; illustrations may not be copied. The abstract must contain conspicuous acknowledgment of SPE copyright.

Abstract

This paper introduces the benefits of a hybrid polymer system for enhanced heavy oil recovery (EOR) that contains
viscosifying anionic polyacrylamides (HPAM) and a recently developed, low toxicity, low molecular weight, dispersing
polymer (LMwDP).

Conventional EOR polyacrylamides increase sweep efficiency, but are not thought to mobilize residual oil since they have
limited effect on wettability and interfacial tension. To improve displacement efficiencies, costly concentrations of
surfactants are added to polymer solutions to increase capillary numbers. These surfactant concentrations are designed to
accommodate the high level of adsorption within the porous medium. Surfactant dosage can be reduced from 0.5% to 1%
down to 0.05% to 0.2% using alkalis, but this often requires expensive water softening systems and higher polyacrylamide
concentrations to reach the required viscosity.

The LMwDP in the hybrid polymer system mimics the surfactant role without some of their disadvantages, at concentrations
between 0.2% and 0.5%. Indeed, the LMwDP affects neither the viscosity of the HPAM solution, nor the injection and
production facilities. The system is particularly beneficial to the recovery of medium to heavy oil, i.e., above 200 cP.
Several HPAM chemistries have been tested at concentrations between 0.03% and 0.2%.

This paper presents experimental oil recovery factors, using HPAM and 20,000 g/mol LMwDP, in 9 to 12 darcy sandpacks
with porosities of approximately 40%. The oil viscosity was 3,000 cP at 50oC, 16,400 cP at 30oC and 40,700 cP at 20oC.

In these reported studies, oil recovery factors with the hybrid polymer system were 70% higher than a brine alone baseline
and 20% higher than brine plus HPAM. Trends in recovery factors indicated that each polymer in the formulation mutually
enhances the performance of the other, i.e., there is a synergy between the components. The sand pack results generated
essential data required for history matching, which then allowed up-scaling using an industry standard reservoir simulator.

Additional studies are in progress to optimize the dosages, and the cost of this hybrid polymer system, for a field
implementation of this EOR technology to the recovery of medium to heavy oil.

Introduction

In order to place the chemical formulations tested into perspective, we first review the characteristics of existing chemical
EOR systems applied for the recovery of medium to heavy oil. We then describe the respective chemistries of the dispersing
polymer (LMwDP) and of the displacing polymer (HPAM) composing the hybrid polymer system (HPS). This is followed
by the reporting of an experimental study into the rheological and interfacial properties of the HPS, as wells as its
performance in dynamic porous media experiments. Lastly, we present the results of a numerical model of displacement
from porous media developed to simulate the performance of the HPS technology in oil recovery.
2 SPE 154054

Recovering Oil from a Porous Medium

We may consider the distribution of oil left in reservoir pores as being distributed through three locations at the pore scale
(Figure 1). First, the largest proportion of oil resides in the wider ‘gaps’ in pores and is considered to be the most mobile oil
fraction. Secondly, a smaller proportion of oil is considered to be adsorbed strongly to the surfaces of pores through short
range interfacial forces. Thirdly, some oil is considered to be trapped mechanically, or bound by capillary forces, in
microcracks and in irregular cavities close to pore throats. Oil in the first location is considered to be removable partially
using water flooding and/or polymer flooding methods, whereas oil in the latter two categories is grouped together, termed
irreducible oil, and considered to be removable with systems containing surface active substances such as surfactants and
alkalis in addition to polymer. We will briefly describe the mechanism behind each system in order to compare their benefits
to the HPS technology that will be discussed in more depth later.

Irreducible Oil

Irreducible Oil

Mobile Oil
Pore Throat

Figure 1 - Mobile and irreducible oil at pore scale

Water Flooding

Waterflooding is usually the first option considered for the secondary recovery of mobile oils after the end of primary
recovery, when much of the natural reservoir drive has diminished. Oil candidates for waterflooding can have viscosities up
to several thousand centipoise at 20˚C. This is the case with the Brintnell Pelican Lake waterflooded pools in Canada, where
dead oil viscosities vary from 1,000 cP to 25,000 cP (ERCB 2011 a.). Other oils in different pools in the same formation
may be even more viscous (ERCB 2011 b.).

In view of the high viscosity contrast between heavy oil and water, heavy oil water flooding is expected to be influenced
substantially by viscous fingering. In such conditions, water “fingers” will penetrate the oil contained in the rock and create a
form of unstable displacement (Chuoke et al. 1959; Peters and Flock 1981), leaving much of the oil remaining in the
reservoir, the recovery factor is in general in the region of 5% to 15%. Even when viscous fingering occurs additional oil
may be recovered after water breakthrough (Mai and Kantzas 2007), but this is at the expense of very high water cuts and low
oil rates, which is uneconomical and impractical.

Polymer Flooding

To further improve recovery factors, at reasonable water cuts, tertiary oil recovery methods have been developed. For heavy
oil, CO2/solvent injection, thermal methods and chemical methods have been widely studied. The latter method frequently
involves chemicals to increase the viscosity of water. Heavy oil water flooding can be improved using the current
understanding of the parameters controlling the efficiency of the process (Vittoratos 2010), but improvements will always be
always limited by the high viscosity contrast of water and heavy oil. Given this viscosity contrast, one common way to
improve the efficiency of the water flooding is through the addition of polymers to increase the viscosity of the displacing
fluid (Manrique et al. 2007; Martin et al. 1983). This tertiary recovery method leads to a decrease in the mobility contrast
between the displacing aqueous phase and the oil, an increase in the volumetric sweep efficiency and a corresponding
increase in the recovery factor.

It is believed that typical polymer injection does not reduce the residual oil saturation per se; rather, it allows the undisplaced
oil to approach to this low level faster, while producing less water in the process. This makes the process more practical and
economical than handling the high water cuts that would be obtained with extended waterflooding. Also, upgrading from
SPE 154045 3

waterflooding to a polymer flood involves a relatively modest increase in surface equipment.

The molecular weights of polymers used in tertiary recovery normally fall in the range 2 million to 22 million g/mol.
Polymers of these dimensions, when in solution at concentrations above 200 mg/l, can spontaneously undergo a process of
entanglement to generate solution viscosity.

Polymer flooding is designed to decrease the mobility ratio (M), which is simply the ratio of the mobility of the displacing
polymer solution (λing) to that of the displaced oil (λed) (Equation 1). These mobilities are themselves regulated by other
ratios concerning properties of the displacing solution and the displaced oil and the porous medium, i.e., the mobility of any
fluid in a porous medium (λ, in darcy/cP), is the ratio of effective permeability of the fluid (k in darcy) to the viscosity of the
fluid (μ in cP) (Craig 1971; Craig 1993; Green and Willhite 1998).

k ing
λing μ ing
M= =
λed k ed
μ ed .……………………………………………………………………………………………. (1)

Equation 1 shows that the mobility ratio can be reduced by, a. decreasing the effective permeability of the displacing fluid
(king), b. increasing the viscosity of the displacing fluid (μing), c. increasing the effective permeability of the displaced fluid
(ked) and d. decreasing the viscosity of the displaced fluid (μed). Despite this apparent simplicity, managing a polymer flood
in a reservoir is a challenging operation since polymers are chemicals that can experience mechanical, chemical and thermal
degradation, all of which can affect the viscosity of the injected fluid (Gaillard et al. 2010).

Mobilizing Additional Oil

Another parameter improving recovery factor concerns the lowering of interfacial tension to increase the displacement
efficiency. This improves the ability of a displacing fluid to release additional oil from microscopic regions of pore space
where oil is trapped by interfacial forces. Oil is released from these regions by increasing the dimensionless capillary number
(Νc) for the system. Factors that increase the capillary number include reducing the interfacial tension between the displaced
fluid and the displacing fluid, increasing the velocity of the fluids flowing through the pores, increasing the viscosity of the
displacing fluid and increasing the water wettability of the natural rock (Adamson and Gast 1997).

The dimensionless capillary number is the ratio of the viscous forces to capillary forces in flow through a capillary. It is
defined in equation 2.

  Fv ν ⋅ μw
Nc = =
Fc σ ow ⋅ cos θ ….…..………………………………………………………………………………………… (2)

where Fv = viscous forces, Fc = capillary pressure forces, ν = displacing flow rate (velocity), σow is the interfacial tension
between the displaced and displacing phases and µw is the viscosity of the displacing fluid. The cos θ term is the cosine of
the reservoir contact angle (Green and Willhite 1998), which is normally determined experimentally for specific rock/oil
combinations. Empirical modifications have been made to equation 2 in order to improve correlations for heavier oils
(Abrams 1975).

Surfactant Polymer Systems (SP’s) and Alkali Surfactant Polymer Systems (APS’s)

Surfactants can be added to polymer solutions to increase the capillary number (Schramm 2000). Surfactant concentrations
fall in the range 0.5% to 1%, but may be reduced to the range 0.05% to 0.2% through the addition of alkalis (McInnes 2008).
Alkaline substances react with hydrocarbons to form in situ surfactants (soaps) by the process of saponification. These soaps,
which are an additional source of surfactancy, modify interfacial tensions at oil/rock interfaces leading to the release of
irreducible oil and the improvement of oil displacement. Another effect of alkali is that it reduces the adsorption of the
surfactant by altering the wettability of the rock. The alkalis can be alkali metal salts of carbonates, silicates or hydroxides.
The use of surfactants is very efficient in recovering more oil but is generally considered as a costly method since it either
requires a high dosage of surfactant or specific pre-conditioning of the injection water to eliminate the hardness that would
react with the alkali. In addition, each reservoir necessitates a full study to design the surfactant slug to be injected properly
and to ensure that a mobile microemulsion is formed and allowed to propagate up to the producer. On the production side,
4 SPE 154054

the surfactant slug will be partially back-produced leading to challenging oil and water separation and reclamation, adding
some costs and complexity to the implementation of these technologies.

General Characteristics of the Hybrid Polymer System (HPS)

The distribution between mobile oil and irreducible oil is less clearly delineated than in the preliminary description in Figures
2a and 2b, which are used to illustrate the principles underpinning existing and the new technology.

a. Irreducible Oil ‘Pulled’


b.
and ‘Stripped ‘ Away by
Combined Effect of
Irreducible Oil Dispersing Polymer and
Displacing Polymer
Irreducible Oil

Mobile Oil
Pore Throat

Highly Mobile Oil Dispersion.


Created by Dispersing
Polymer and Pushed by
Displacing Polymer

Before Treatment During Treatment

Figure 2 - The distribution of oil before and during treatment

To achieve the displacement benefits, a chemical EOR solution containing two polymers is deployed. The term displacing
polymer or HPAM is used to define the component that provides viscous, Non-Newtonian, characteristics to a water based
formulation for the purpose of hydraulic displacement. The term dispersing polymer or LMwDP refers to a water soluble
component that allows the formulation to disperse oil and simultaneously modify interfacial characteristics. In essence, the
proposed Hybrid Polymer System or HPS provides the combined properties of both polymers in order to remove oil from
all three locations (Figure 2). Each polymer enhances the performance of the other and we consider the formulation to be
synergistic in that it provides an improvement in the macroscopic sweep efficiency of a treatment and simultaneously
improves the microscopic displacement efficiency. Both polymers are soluble in fresh waters and brines with total salinities
of up to 100,000 mg/l. The formulation enhances oil production by modifying the following properties:

1. The sweep efficiency will be increased, by decreasing the mobility ratio (M) through the appropriate modification of
points a to d below (Equation 1).
a. The viscosity of the displacing phase (μing) will be increased by the presence of the HPAM.
b. The chemical formulations will reduce the effective permeability of the displacing fluid (king). This effect is
difficult to quantify.
c. The viscosity of the displaced fluid  (μed)  will be decreased by the formation of a low viscosity oil dispersion in
the pore space, through the action of the dispersing polymer.
d. The effective permeability of the displaced fluid (ked) will be increased by the water wetting of the pore surfaces
and the reduction on oil/water IFT, through the action of the dispersing polymer.
2. In addition to improvements in the sweep efficiency through the mobility ratio, the formation of a low viscosity
dispersion within a pore will minimize the viscosity difference between the displacing fluid and the displaced fluid, thus
minimizing viscous fingering.
3. The displacement efficiency will be increased as a consequence of the increase in the capillary number. This effect is
directly related to the increase in the viscosity of the displacing phase and inversely related to the reduction in the
interfacial tension.
SPE 154045 5

4. The residual resistance factor is lower than that achievable with polymer solutions alone, preferably less than 1.3. We
believe that this is achieved as a consequence of the reduced adsorption, and reduced mechanical entrapment, of the
displacing polymer in the presence of the water wetting dispersing polymer.

Figures 3 and 4 show generic molecular formulae for the two polymers.

CH2 CH CH2 CH CH2 CH


x y z
C=O C=O C=O

NH2 O- Na+ NH

CH3 C CH3

CH2

SO3- Na+

Figure 3 - The displacing polymer (HPAM)

CH2 CH CH2 CH CH2 CH


n m t
OH O R

C=O

CH3

Figure 4 - The dispersing polymer (LMwDP)

The displacing polymer (Figure 3) can be linear and of high enough molecular weight to allow the generation of viscosity at
concentrations between 200 mg/l and 4000 mg/l. Depending on reservoir conditions, the displacing polymer can be
structured to be associative, star or comb. It can be designed to provide salt, shear, temperature resistance, to increase the
actual viscosity in the porous medium for a same polymer concentration as a linear one or be viscoelastic in solution. The
symbols n, m, t, x, y and z are integers representing the numbers of each functional group on the respective polymer
backbone. For this displacing polymer (Figure 3), x is in the range of 50% to 100 mol%, y and z are in the range of 0% to 50
mol%. The sum of x, y and z corresponds to 100 mol%. The molecular weight for the displacing polymer may vary between
2 and 22 million g/mol.

For the dispersing polymer (Figure 4), n and m are in the range of 50% to 100 mol% and t is in the range of 0% to 10 mol%.
The sum of n, m and corresponds to 100 mol%. The degree of hydrolysis, (100n/(n+m)) is less than 90%. The molecular
weight for the dispersing polymer may vary between 5,000 and 25,000 g/mol. The functional group R is a neutral species.

Specific Characteristics of the New Polymer Formulation (HPS)

Dispersion Formation (Using Dispersing Polymer LMwDP Alone)

A solution containing 0.5% of the dispersing polymer (LMwDP) has a Newtonian viscosity of ~1.2cP at 20oC. When such a
solution of dispersing polymer is blended into a heavy oil, at an oil to solution ratio of 70:30, the polymer interacts with oil
surfaces spontaneously to break-up the oil into mobile deformable droplets. These separated oil droplets are encapsulated by
a molecular layer of the dispersing polymer. The outcome is that the oil droplets can slip and roll over each other, and also
slip over surfaces to which they would ordinarily be bound. Oil droplets are therefore mobilized in environments where they
would otherwise be immobile. Minimal agitation is required to achieve this effect, and the properties of the oil in the
droplets remain unchanged by the dispersing polymer. Given time, and zero agitation, these droplets coalesce since they are
not stabilized through tight emulsion formation. The LMwDP differs from conventional surfactants in a number of ways.
First, surfactants stabilize oil dispersions via the formation of tight emulsions rather than through the formation of unstable
deformable droplets. In contrast, oil droplets in LMwDP dispersions can be an order of magnitude greater than those found
in surfactant stabilized emulsions. This is demonstrated in Figure 5, which shows droplet size distributions for dispersions
made with a heavy oil (1,500 cP at 25oC) using either LMwDP at 0.5% or the surfactant Triton X-100 at 5000 mg/l. All
dispersions had the Oil:Solution ratio of 70:30 and were made by mixing the oil and the water phases by hand and shaking 50
6 SPE 154054

times (low shear mixing). In these experiments, high shear blends were made by shearing the low shear dispersions further
using an Ultra Turrax mixer, with a 25 mm head, at 6,000 RPM for 15 minutes (high shear mixing). Droplet size
distributions were measured using Laser Light Scattering (Malvern Mastersizer 2000). Figure 5 also demonstrates that a
dispersion made with LMwDP can be made using minimal shear and that further shearing leads to a reduction in mean oil
droplet size. In the case of the Triton X-100 dispersion, compared with LMwDP, the droplets size is reduced significantly
using the same amount of shear energy leading to stable emulsions (Sjoblom 2001). Secondly, air/water and water/oil
interfacial tensions are reduced less by LMwDP than by surfactants (Table 1), which we interpret as being critical part to the
formation of mobile deformable droplets as opposed to the formation of tight stable emulsions (Winsor 1968).

LMwDP (High Shear)


12
LMwDP (Low Shear)
% Population 10 Triton X-100 (High Shear)

0
1 10 100 1000 10000

Size μm

Figure 5 - Droplet size distributions for oil

For Newtonian heavy oils, the rheology of the dispersions are largely independent of the oil type. Equivalent Newtonian
viscosities for dispersions, in cylindrical pipeline geometries, are commonly in the range 100 cP to 300 cP. In this sense, and
to a first approximation, the polymer solution standardizes the mobility of heavy oils (Figure 6). Typically, oils that have
viscosities of less than 100,000 cP at 25oC conform to this rule.

100,000

90,000

80,000
Viscosity cP

70,000

60,000

50,000

40,000

30,000

20,000

10,000

0
100k cP Oil 30k cP Oil 300 cP Dispersion

Figure 6 - Apparent viscosity reduction

At a more detailed level, dispersions exhibit pseudo-plastic rheology that is best described using temperature dependent
consistency indices and power law coefficients. For example, Figure 7 shows flow curves for low shear 70:30 dispersions, at
25oC, made using a selection of heavy oils. All dispersions were made using tap water and 0.5% dispersing polymer. The
shear thinning characteristics are clear, as is the relative independence of the flow curves on the original oil viscosity.
SPE 154045 7

10000
Original Oil Viscosity

1,016cP

Apparent Viscosity  cP
1000 6,739cP
6,912cP
17,212cP
100
30,241cP
37,573cP
10
47,145cP
71,869cP

1 86,314cP
0.1 1 10 100
109,143cP
Shear Rate 1/s

Figure 7 - Flow curves for multiple low shear dispersions

All flow curves were determined using an Anton Paar MCR 501 rheometer and experimental uncertainty was estimated to be
± 25%. Dispersion properties were determined using parallel plate geometries, whereas oil properties were determined using
cone and plate geometries.

Wettability Modification Using Dispersing Polymer LMwDP Alone

These tests, with conventional Amott cells, concerned the semi-quantitative comparison of the tendency of the dispersing
polymer solution to spontaneously water-wet a silica surface, and displace any associated oil. The tests involved soaking
identical slurries composed of sand, oil and brine, with either the LMwDP solution or a representative brine, for sixteen days,
and monitoring the proportion of detached oil as a function of time. The sand:oil:brine ratio in the slurries was 60:34:6,
which represents an oil saturation of 85%, a corresponding brine saturation of 15%, and is consistent with selected field
conditions. After the sixteen day soak period, the dispersing polymer solution was observed to have released the highest
proportion of oil (44.3% original oil), whilst the brine solution released only 16.9%. The brine solution was found to reach
its maximum oil recovery after only two days, but the dispersing polymer solution was found to detach oil more rapidly and
continue to detach oil for the entire 16 day period (Figure 8). The increased mobilization of oil under these static conditions
indicates that the wettability of the sand surface is changed in the presence of the LMwDP polymer towards a more water-wet
character.

50
45
40
35
Oil Recovery %

Brine Alone
30
25 Dispersing Polymer Solution LMwDP
20
15
10
5
0
0 5 10 15 20
Elapsed Time (Days)

Figure 8 - Oil recovery in Amott cell experiments.

Compatibilities

The objective of the compatibility tests was to demonstrate that the preferred LMwDP did not significantly change the
rheological properties of the displacing polymer in solution. The tests consisted of creating solutions containing the LMwDP,
in brine of 36,000 mg/l total dissolved solids, and then determining their steady shear rheograms at 20oC, in the presence or
absence of the preferred displacing polymer. Duplicate rheological measurements were made using a Bohlin Gemini
Rheometer from Malvern with a cone and plate geometry (2°, 6 cm). Two displacing polymers (HPAM) were selected for
8 SPE 154054

testing: FP3630S (18 million g/mol, 30% copolymer anionic polyacrylamide) and FP3330S (9 million g/mol, 30% copolymer
anionic polyacrylamide). Data in Figure 9 shows that there are only minor reductions in the solution viscosities of the
potential displacing polymer when the dispersing polymer is added. FP3630S was taken for sandpack displacement studies.

Figure 9 - Rheograms for a selection of HPAM solutions with and without LMwDP

Dispersion Formation in Low Salinity Tap Water (HPS Solution Containing Both HPAM and LMwDP)

The objective of these tests was to demonstrate the ability of the HPS, containing both HPAM and LMwDP, to create low
viscosity dispersions using oils from Northern Alberta. Three crude oils were selected for testing, one of which was taken
forward to sandpack displacement testing. The treatment fluid was composed of the LMwDP at 5000 mg/l plus the
displacing polymer 300 mg/l (FP3630S). Figures 10a. and 10b. show the flow curves for the three oils and their
corresponding low shear, 70:30, dispersions at two temperatures.

a. Data at 20C b. Data at 25C
100000
100000

Oils
Apparent Viscosity  cP

Oil 1
Apparent Viscosity  cP

Oil 2 Oils Oil 1


10000
Oil 3 10000 Oil 2
Dispersion with Oil 1 Oil 3
Dispersion with Oil 2 Dispersion with Oil 1 
Dispersion with Oil 3 Dispersion with Oil 2
Dispersion with Oil 3
1000
1000

Dispersions
Dispersions

100
100
0.1 1 10 100
0.1 1 10 100

Shear Rate 1/s Shear Rate 1/s

Figure 10 - Rheograms for neat oil and low shear dispersions made with the tap water based treatment fluids containing dispersing
and displacing polymers:70:30 Oil:Water

Given the experimental uncertainty in dispersion characterization, Figure 10 must be interpreted as clearly demonstrating the
ability of the full treatment formulation (HPS) to increase the mobility of the respective oils.

Dispersion Formation in High Salinity Brine (HPS Solution Containing Both HPAM and LMwDP)

The above tap water experiments are repeated using a brine carrier fluid with a total dissolved solids content of 36,000 mg/l,
a calcium content of 600 mg/l and a magnesium content 226 mg/l. The dispersion rheograms made with Oil 2, at 25oC and
35oC, in the presence and absence of the brine components, are shown on Figure 11. The anticipated effect of the brine
SPE 154045 9

components is to cause the hydrodynamic volume of the displacing polymer chains to shrink, leading to a reduction in
viscosity at all shear rates. This is seen in at both temperatures in Figure 11.

100000

Oil 2 at 25C

Apparent Viscosity cP
10000
Dispersion in Tap Water at 25C
Dispersion  in Brine at 25C
Dispersion in Tap Water at 35C
Dispersion in Brine  at 35C
1000

100
0.1 1 10 100

Shear Rate 1/s

o
Figure 11 - Rheograms for neat oil and oil dispersions at 25 C (HPS in the presence and absence of brine:70:30)

Interfacial Tension Reduction (HPS Solutions or HPAM and LMwDP Independently)


Table 1 shows IFT measurements performed on Oil 2 (from Northern Alberta) in the presence of a number of test solutions.
IFT measurements were performed using the pendant drop technique and a Kruss DSA 100. Oil 2 had a viscosity of 40,700
cP at 25oC and a BS&W of <0.5%.

Table 1- IFT Measurements


Test Solution IFT Solution-Air (mN/m) IFT Solution-Oil (mN/m)
Waterflood Brine (36000 mg/l TDS) 73.43 24.41
HPAM Alone (300mg/l in brine) 71.78 23.17
LMwDP Alone (5000mg/l in brine) 50.33 10.43
HPS 48.41 9.29
Triton X-100 Surfactant (5000mg/l) 32.15 0.82

Table 1 shows that the displacing polymer (HPAM, FP3630S, 300 mg/l) has little effect on the interfacial properties of the
brine or the oil, but the presence of the dispersing polymer (LMwDP, 5000 mg/l) reduces interfacial tensions by up to 50%.

Sand Pack Displacement Characteristics

General Methodology

Experiments discussed in this section were designed to confirm the ability of a preferred treatment fluid to disperse oil from
porous media and demonstrate enhanced behaviour. These involved conventional one dimensional sandpack displacement
experiments, conducted in order to demonstrate the effects and also to extract the petro-physical parameters needed to
construct a three dimensional model for full field implementation. An overview of experimental methodologies is given
elsewhere (Craig 1993). Oil 2 in Figure 10 (3,000 cP at 50oC, 16,400 cP at 30oC, 40,700 cP at 20oC) was selected for
sandpack testing. Tests were conducted using one or other of the following comparative solutions.

1. A: Brine solution (no polymers) with total a dissolved solids content of 36,000 mg/l, a calcium content of 600 mg/l and a
magnesium content 226 mg/l.
2. B: Brine A plus dispersing polymer at 5000 mg/l.
3. C: Brine A plus displacing polymer (HPAM, FP3630S) at 300 mg/l, with a viscosity of ~3cP at a shear rate of 1s-1 at
25oC.
10 SPE 154054

4. D: Brine A plus dispersing polymer LMwDP at 5000 mg/l, and HPAM at 300 mg/l, with a viscosity of ~3cP at a shear
rate of 1 s-1 at 25oC.

Solution D is the new treatment fluid (HPS) and solution A is a brine representing a reference waterflood. Individual
sandpack experiments involved the following generic steps:

1. Construct a sandpack.
2. Saturate the pack with the selected brine and determine initial brine permeability.
3. Displace the brine to residual brine saturation with the oil and then determine the total oil volume in the pack and the
permeability to oil at the irreducible brine saturation.
4. Age at 50oC for a period of 10 days.
5. Displace the oil with 4 PV’s of Fluid A (brine).
a. Collect all volumetric fractions of displaced fluid. Each fraction being ~10cm3
b. Record differential pressure across the pack continuously.
c. Measure oil and water content of all collected fractions
6. Repeat step e. using 4 PV’s of fluid B.
7. Calculate recovery factors, mobility ratios and residual resistance factors (Green and Willhite 1998).
8. Construct a plot of recovery factor vs total pore volumes of fluid injected (PVinj), for the entire two-fluid displacement
sequence.
9. Repeat steps 1. to 8. two further times, using fluid A in all cases but replacing fluid B with Fluid C and then D.

All experiments were performed using a sandpack (66 cm length x 4 cm internal diameter) constructed from 316 stainless
steel. The sandpacks were packed with silica sand using a slurry packing method, and the porosity was determined via
calculation from the weight of sand used and the sandpack volume. The selected silica sand had a standard mesh size of 100
to 140 (149 microns to 105 microns), a spherical grain shape, a specific gravity of 2.65 g/cm3, and a chemical composition
of 98.2 % SiO2 where the major impurities were Al2O3 (0.49 %) and Fe2O3 (0.14 %). The process lead to the sand packs
having a mean permeability of 10.5 ± 1.5 darcy, and a mean porosity of 37.8 ± 0.2%. After packing, 4 pore volumes of brine
were injected into the sandpack held in a vertical orientation in order to flush any residual air from the system and ensure
complete brine saturation. A Rosemount differential pressure transducer was connected across the injection and production
ports in order to record continually the differential pressure across the sandpack for each flooding process. This allowed the
determination of the brine permeability of the sandpack prior to any flooding experiments. The pressure transducer was used
to measure and record data using LabView software. There was no confining pressure. A dual ISCO injection pump was used
to inject fluids at the appropriate rate, and the pump injection pressure was continuously recorded using LabView software.
The input line from the injection pump to the oven was kept at a constant 25oC using a re-circulating water bath, while the
production line from the oven to the effluent collector was kept at a constant 50oC using a cable heater. The sandpack and
associated transfer vessels were kept at a constant 50oC for the duration of the experiments. Figure 12 shows a general
schematic of the experimental set-up.

Oven
Pump
Sand Pack
Fluid Collector
Transfer
Vessel

Pump
Pressure
Transducer
Transfer
Vessel
Test Tubes

Injection Fluid
Container Data
Acquisition

Information Transfer Lines

Figure 12 - Experimental configuration


SPE 154045 11

Recovery Factors and Chemical Retention


The recovery factor curves for the waterflood, the dispersing polymer alone, the displacing polymer alone and the HPS, are
compared in Figure 13. To eliminate inconsistencies between experimental trials, the recovery factor results after the initial 4
PV of water flood were normalized to a value of 42.1%, corresponding to the recovery factor at the end of the initial
waterflood phase determined by averaging all readings. Figure 13 shows that displacement with the various formulations
resulted in four observations:

1. With the two component blend (HPS), the recovery factor had risen to 71.0 %, which represents an incremental recovery
after water flood of 28.9 % (Table 2). With the HPS, the incremental recovery is much greater than for the dispersing
polymer alone (10.3 %) or the displacing polymer alone (15.8 %).
2. The rate of recovery with the HPS, immediately after waterflood, was significantly higher than with either the dispersing
polymer alone or the displacing polymer alone.
3. Evidence emerged that the formulation has synergistic elements. This can be seen by comparing the recovery curve for
the HPS with a hypothetical curve generated from the simple additive contribution from HPAM and LMwDP (Figure
13). The performance of the two component fluid is more than the sum of its parts.
4. The improved recovery coming from the HPS is not due to an increase of the viscosity of the water phase as there are no
major interactions between HPAM and LMwDP (Figure 9). If anything, we see a small decrease in the HPS viscosity
compared to HPAM alone.

80.00

70.00

60.00
Recovery Factor %

50.00

40.00

30.00 Averaged Brine (Waterflood)
LMwDP Solution Alone
20.00
HPAM Solution Alone
HPS
10.00
Additive Contribution  from HPAM and LMwDP
0.00
0.00 2.00 4.00 6.00 8.00

Injected Pore Volumes (PVinj)

Figure 13 - Recovery factors vs injected pore volumes (PVinj)

Recovery factors are summarized in Table 2.

Table 2 - Recovery Factors 


Fluid Recovery Factor (%) Final Incremental Recovery Factor
(% over Reference Waterflood)
At 4 PVinj At 8 PVinj
Waterflood Brine (36,000 mg/l TDS) 42.1 42.1 N/A
LMwDP Alone (5000mg/l in brine) 42.1 52.4 10.3
HPAM FP3630S Alone (300mg/l in brine) 42.1 57.9 15.8
HPS 42.1 71.0 28.9

We interpret the results (Figure 13 and Table 2) as indicating that a combination of the two polymer components, each with
different chemical functions, serves to maximize oil recovery by affecting the oil phase and aqueous phase viscosities as well
as the interfacial tension and wettability. This translates into positive changes in the oil phase and aqueous phase mobilities
as well as capillary number that help to increase oil recovery efficiency. In addition to these changes, a reduction in the
displaced phase viscosity would increase the sweep efficiency by reducing small-scale viscous instabilities, which are
commonly encountered when there is a large viscosity contrast between the displaced fluid and the displacing fluid. Table 3
shows calculated mobility ratios and residual resistance factors for the comparative displacement fluids, supporting the above
12 SPE 154054

argument. More detailed analysis of the data indicated a reduction in the residual oil saturation, which is consistent with the
proposition that the HPS acts effectively without the need for a substantial reduction in IFT as is normally required with
surfactants (Table 1).

Table 3 - Calculated Mobility Ratios and Residual Resistance Factors


Fluid Mobility Ratios Residual Resistance Factors
Waterflood Brine (36,000mg/l TDS) 8.46 N/A
LMwDP Alone (5000mg/l in brine) 2.92 1.28
HPAM FP3630S Alone (300mg/l in brine) 0.7 0.86
HPS 0.27 1.08

In a tertiary treatment it is important that the retention of any chemical is minimal, in order that loss of chemical to the
formation is minimized and that the performance of the system is not impaired. This is particularly important with the hybrid
formulations, where we anticipate that synergy will be maximize when both polymers travel through the formation together.
Therefore, there should be no significant chromatographic separation of the polymers. Chemical analysis of the multiple
fractions collected from the sandpack test, using customized spectroscopic methods, allows the relative retention of the
polymers to be calculated. Figure 14 shows a plot of the two polymer concentrations (expressed as a fraction of their
injected concentrations) as a function of injected pore volumes. It is clear that the two polymers travel through the sandpack
at almost identical rates.

0.45 0.045

HPAM Polymer Concentration %


0.40 0.040
LMwDP  Polymer Concentration  %

0.35 0.035 LMwDP

0.30 0.030

0.25 0.025 HPAM

0.20 0.020

0.15 0.015

0.10 0.010

0.05 0.005

0.00 0.000
0.00 1.00 2.00 3.00 4.00 5.00 6.00
Pore Volumes Injected

Figure 14 - Polymer retention

Computer Simulations of Treatment Performance

A basic numerical model to simulate the performance of the synergistic EOR process at field scales was constructed. The use
of simulators to ‘up-scale’ reliably is a process where both laboratory data and field data is history matched iteratively, with
the final outcome being the ability to predict three dimensional (3D) performance from one dimensional (1D) laboratory
measurements. If laboratory data cannot be history matched within experimental error, there is little chance that full 3D
simulations will be reliable. However, once validated, simulations provide useful guidelines for full field simulation and
optimization. Results in this section should therefore be interpreted as an early stage tentative simulation, with a view to
assessing the internal consistency of the proposition, rather than as a final conclusion.

We elected to use a commercial simulator package supplied by CMG (Computer Modeling Group). The model was built,
using symmetry elements, for a vertical well configuration with the geometry of an inverted 5-well pattern. The distance
between the injectors was 141 m, and the distance between the injector and producers was 100 m. Pay zone thickness was
15 m and the injectors were ‘¼ injectors’. Geometry considerations required that the flow rate per injector was 1/4 of the
total injection rate in order to keep the injector to producer ratio equal to 1. The total injection rate was set to 100m3/day.
The permeability, porosity, reservoir geometry and injector/producer pattern, used in these simulations, were selected to
represent generic conditions found in Northern Alberta.
SPE 154045 13

Data obtained during the laboratory displacement tests were first history matched to obtain relative permeability curves for
the different fluids. Additional relevant data, such as polymer viscosity as function of concentration, temperature and shear
rates, polymer adsorption and oil viscosity were also required input data for the history matching. Once the history matching
of laboratory data was completed the model was reconfigured to the represent the 3D reservoir geometry and then used to
predict displacement characteristics. During the 3D prediction, the sets of relative permeability curves derived from the
history matching of laboratory data were interpolated depending on the concentration of the dispersing polymer. The basic
numerical modeling presented here does not consider different permeability layers or heterogeneities that would impact a real
reservoir case. This initial ‘up-scaling’ procedure basically consisted of increasing simulation block dimensions by a length
factor and scaling up injection rates and times using the same factor. Other important parameter such as injection pressures
that are constrained to a maximum allowable level were studied in more advanced model not presented here.

A schematic of the reservoir and injection pattern is given in Figure 15.

Figure 15 - The 3 dimensional reservoir Model

Figure 16 shows the predicted recovery factors for this injection scenario. It can be seen that the same trends as displayed in
the experimental sandpack tests (Figure 13) were reproduced. The simulated recovery factors for the two component
formulation (HPS) are higher than for either the displacing polymer alone (HPAM) or the dispersing polymer alone
(LMwDP). The waterflood achieved the lowest oil recovery. These recovery factor predictions were qualitatively consistent
with the sand pack data, which we take to support the basic proposition of synergy within the HPS.

Two Component Polymer Blend


Displacing Polymer Alone (HPAM)
Dispersing Polymer Alone
Waterflood

Figure 16 - Predicted oil recovery for the 3D model

Conclusions

A new polymer system (HPS) has been developed and tested for its ability to enhance recovery factors. This system contains
two polymers, a displacing polymer based on HPAM plus a salt tolerant low molecular weight dispersing polymer
(LMwDP), that act synergistically to improve recovery factors. The dispersing polymer is designed to interact with oil
14 SPE 154054

surfaces and break-up the oil into mobile deformable droplets. In a sandpack study, measured oil recovery factors with the
HPS were 70% higher than a brine alone baseline and 20% higher than brine plus HPAM. Recovery factors for the HPS
were greater than would be anticipated from the simple addition of the experimental recovery factors for the two polymers
used separately, i.e., the performance of the two component fluid is more than the sum of its parts. Preliminary computer
simulations and history matching of laboratory sand pack experiments supported the basic proposition of synergy within the
HPS.

The improved recovery coming from the HPS is not due to an increase of the viscosity of the water phase since no major
interactions between HPAM and LMwDP were detected experimentally. Rather, it is proposed that synergy is derived from a
combination of factors that simultaneously enhance sweep efficiency and displacement efficiency, and also minimize viscous
fingering. The synergistic benefits are derived from the optimized rheological and interfacial properties of the HPS. In
particular, the HPS leads to an increase in the mobility of the resident oil phase, a reduction in brine/oil surface tension and a
tendency to water-wet silica surfaces leading to the release of bound oil.

References

Abrams, A. 1975. The influence of Fluid Viscosity, Interfacial Tension and Flow Velocity on Residual Oil Saturation Left by Waterflood.
SPE Journal, 437-447.
Adamson, A.W. and Gast, A.P. 1997. Physical Chemistry of Surfaces, 6th Edition, ISBN 0471148733.
Chuoke, R.L., van Meurs, C. and van der Poel. C. 1959. The instability of slow, immiscible, viscous liquid-liquid displacements in
permeable media. Trans. AIME 216, 188–194.
Craig, F.F. 1971. The Reservoir Engineering Aspects of Waterflooding, Monograph Series, SPE, Richardson, TX.
Craig, F.F. 1993. The Reservoir Engineering Aspects of Waterflooding, pp 134, ISBN 9780895202024
Eagland, D., Crowther, N.J., Murray J. and Tobin. A. 2003. Recovering Materials, International (PCT) No.WO03083259A2
Economides, M.J. and Nolte, K.G. 2000. Reservoir Stimulation, Third Edition, ISBN 0471491916.
ERCB, 2011a., Cenovus Report: Performance Review of In Situ Oil Sands Scheme Approval 9404k. Calgary, Pelican Wabiska Team.
ERCB, 2011b., CNRL Report: Annual ERCB Performance Presentation. In-Situ oil Sands Schemes 9673/ 10147/10423/10787.
Gaillard, N., Giovannetti, B. and Favero, C. 2010. Improved oil recovery using thermally and chemically protected compositions based on
co- and ter- polymers containing acrylamide. SPE 129756.
Green, D.E. and Willhite, G.P. 1998. Enhanced Oil Recovery, Textbook Series SPE, Richardson TX, Volume 6.
Huh, C.C. and Pope, G.A. 2008. Residual Oil Saturation From Polymer Floods: Laboratory Measurements and Theoretical Interpretation,
SPE 113417.
Mai, A. and Kantzas, A. 2007. Heavy Oil Waterflooding: Effects of Flow Rate and Oil Viscosity. Paper 144. Canadian International
Petroleum Conference.
Manrique, E.J., Muci, V.E. and Gurfinkel, M.E. 2007. EOR field experiences in carbonate reservoirs in the United States. SPEREE
(December).
Martin, F.D., Hatch, M.L., Sheptika, J.S. and Ward, J.S. 1983. Improved Water-Soluble Polymers for Enhanced Recovery of Oil, SPE
11786.
McInnis, L. 2008. Innovative Energy Technologies Program Application 01-023 Taber S Mannville B (Alkaline Surfactant Polymer Flood)
Peters, E.J. and Flock, D.L. 1981. The onset of instability during two-phase immiscible displacement in porous media. Society of
Petroleum Engineers Journal.
Schramm, L.L. 2000. Surfactants: Fundamentals and Applications in the Petroleum Industry, ISBN 0521640679.
Sheng, J.J. 2011. Modern Chemical Enhanced Oil Recovery, ISBN, 9781856177450.
Sjoblom, J. 2001. Encylopedic Handbook of Emulsion Technology, ISBN 0824704541
Vittoratos, E. 2010. Optimal Heavy Oil Waterflood Managment May Differ from that of Light Oils. SPE 129545.
Wang, D.M., Cheng, J.C., Yang, Q.Y., Gong, W.C., Li, Q. and Chen, F.M. 2000. Viscous-Elastic Polymer Can Increase Microscale
Displacement Efficiency in Cores, SPE 63227.
Winsor, P.A. 1968. Chem Rev, 68,1.
Xia, H.F., Wang, D.M., Wang, G., Ma, W.G. and Liu, J. 2008. Mechanism of the Effect of Microforces on Residual Oil in Chemical
Flooding. SPE 114335.
Yin, H.J., Wang, D.M. and Zhang, H. 2006. Study of Flow Behaviors of Viscoelastic Polymer Solution in Micropore with Dead End. SPE
101950.

You might also like