Você está na página 1de 21

SPE-172850-MS

Polymer Flooding for Extra-Heavy Oil: New Insights on the Key Polymer
Transport Properties in Porous Media
F. Rodriguez, PDVSA, Paris Diderot University; D. Rousseau, and S. Bekri, IFP Energies nouvelles;
M. Djabourov, ESPCI Paris Tech; C. Bejarano, PDVSA

Copyright 2014, Society of Petroleum Engineers


This paper was prepared for presentation at the SPE International Heavy Oil Conference and Exhibition held in Mangaf, Kuwait, 8 10 December 2014.
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
Primary cold production in extra-heavy oil reservoirs (7-12 API) is currently a low percent of the OOIP,
as for some oil fields in La Faja Petrolfera del Orinoco (FPO), Venezuela. EOR studies are being
conducted in order to increase recovery factors primarily in those thin bedded reservoirs which host up
to 35% of the OOIP. Thermal EOR is usually the first option implemented to increase recovery. However,
thermal methods are not suitable for thin pay zones and involve high water consumption which can create
environmental issues. For these reasons, chemical EOR is becoming a feasible tool for mobility control
and mobilization of residual oil.
Polymer flooding applied to extra-heavy oil may require rather high concentration polymer solutions.
The objective of this study is to contribute to a better understanding of the specific mechanisms involved
in this context. Transport properties of polymer in porous media are investigated through a set of
corefloods tests performed on model sandpacks, the high permeability (4 Darcy) and the temperature
(50C) being representative of the conditions of the FPO. Polymer transport is analyzed versus injection
concentration in terms of the key petrophysical parameters controlling the performance of a polymer
flooding operation, namely: retention and inaccessible pore volume, in-depth propagation and injectivity.
Results show that all parameters strongly depend on polymer concentration. Specifically, the concentration dependence of polymer retention, inaccessible pore volume and in-depth propagation behavior
cannot be interpreted by conventional models. Injectivity is also strongly affected by concentration: both
rheo-thinning and rheo-thickening effects become more marked when concentration increases. This set of
experimental data are the key for modeling the transport of high concentration polymer solutions in porous
media. They also provide direct information needed, through reservoir simulation, to sanction the
technical and economical feasibility of polymer flooding for extra-heavy oil.

Intoduction
Heavy and extra-heavy oils are crude oils characterized by a dense, viscous and asphaltic nature (high
heteroatoms, heavy metals, asphaltene and/or resin content). Most of the heavy and extra-heavy oils are
product of biological alteration of conventional crude oils mediated by in-reservoir bacterial colonies (Huc
et al. 2011).

SPE-172850-MS

Figure 1Location of La Faja Petrolifera del Orinoco (FPO) in the Southern Margin of the Eastern Venezuelan Basin (Rojas et al. 2009).
Table 1Comparison of Venezuelan extra-heavy crude oil and Canadian/Chinese/Oman oils.
Heavy oil reservoir
location
Canada
China
Oman
Venezuela

API

Acid number
(mgKOH/g-oil)

Oil viscosity
(cps)

Reservoir Temperature
(C)

11.5-16.5
15-20
20
7-12

1-1.5
0.8
4.3

600-80000
500-2000
250-500
1000

12-17
55
50
48-60

Primary recovery for heavy and extra-heavy oil reservoirs is limited to few percentage of the OOIP.
This is the case of some extra-heavy oil reservoirs in the FPO in Venezuela: many of them have recovery
factors lower than 5 %. Figure 1 shows a location map of the FPO, which bears the largest reserves of
extra-heavy oil in the world with about 1.2 trillon barrels across a surface of 55000 square kilometers.
These deposits lie parallel to the northern bank of the Orinoco River whose main characteristics are the
following: length from east to west is 500-600 km, depth between 350 and 1000 m (1150-3280 ft), average
porosity 30%-35% and permeability higher than 2 D. Venezuelas extra heavy oil is categorized as a
foamy oil (initial GOR between 60 and 75 scf/stb), distinguished by its API gravity (7-12API), its acidity
and its very high viscosity at reservoir conditions (1000 cP). The temperature of the reservoir in the FPO
is high enough to keep the oil mobile (Banerjee et al. 2012), in the range of 48-60C. Table 1 shows the
comparison of Venezuelan extra-heavy crude oil and Canadian/Chinese/Oman oils.
According to the literature, both lab and field experiences have demonstrated that water flooding of
heavy oil reservoirs achieves very low oil recovery factor due to the adverse mobility ratio (poor sweep
efficiency), causing water to finger through the oil and to bypass large portions of reservoirs, leaving
behind substantial amounts of oil (Dong et al. 2011).
Using Enhanced Oil Recovery (EOR) processes is technically and economically attractive for production of heavy and extra-heavy oils. Thermal methods are more commonly applied at field scale (cyclic
steam stimulation, continuous steam injection and SAGD). Nonetheless, many thin-pay-thickness heavy
and extra-heavy oil reservoirs are not suitable for thermal methods (thin formations, less than 30 ft; Farouq
Ali et al. 1976); because the heat losses are significant and there is not benefit of the gravity effect when
steam is injected into these thin formations. This is the case of many reservoirs in Venezuela and Western
Canada (Farouq Ali et al. 1979 and Wang et al. 2010). It is also important to mention that thermal methods
involve high water consumption which can create environmental issues.

SPE-172850-MS

Chemical EOR (CEOR) shows big potential for enhancing viscous oils recovery in reservoirs where
thermal processes are not feasible. CEOR processes are based on the addition of chemical additives
(polymer and potentially surfactant and/or alkali agents) into the injection water. These methods represent
a challenge for enhancing heavy and extra-heavy oil production and require significant efforts of research
and field testing to reduce the cost-efficiency ratio (Baviere et al. 1991). Polymer flooding is usually the
first CEOR method considered for an heavy oil field. It consists of adding polymer to the water of a
waterflood to increase its viscosity which translates into decreasing its mobility and improving the sweep
efficiency through a more favorable oil/water mobility ratio (Larry, L et al. 1989; Sorbie et al. 1991;
Wassmuth et al. 2007 and Sheng et al. 2011). Nowadays, elaborated research works on polymer flooding
for heavy oil are aimed at establishing empirical models to predict the recovery improvement that polymer
floods can entail (Mai and Kantzas, 2009), better understanding the sweep enhancement mechanisms
(Seright et al. 2010) in particular under adverse mobility ratio conditions (Levitt et al. 2013), better
predicting the adsorption/retention properties of the polymers (Zhang and Seright, 2013; Manichand and
Seright, 2014) and investigating the performances of new polymer molecular architectures (Wassmuth et
al. 2012).
Polymer flooding has recently been successfully applied at field scale in the Pelican Lake heavy oil
field, Canada (Tabary et al. 2013). Regarding the Venezuelans conditions for extra-heavy oil production,
nowadays, there is not any implementation of CEOR methods at field scale. An early attempt of using
alkaline for enhanced recovery of heavy oils was reported in 1979 (Farouq Ali et al. 1979) for a Canadian
and a Venezuelan cases. Although the conclusions of this laboratory investigation were positive (recovery
factors between 59 and 65%, by injecting caustic solutions into the heavy oil) trials have not been
performed on the basis of this work. Implementation of polymer flooding technology is now also
considered for Venezuelan extra-heavy oil in the FPO (Romero et al. 2013; Fabbri C. et al. 2013 and
Fabbri C. et al. 2014); lab and simulation studies have been performed for Petrocedeo area (joint venture
between PDVSA and TOTAL) looking for near future field application.
The viscosity required for a polymer flooding operation needs to be predicted though reservoir
simulation assays which has to include both technical and economical aspects. For heavy oil, the field
experience on Pelican Lake (Tabary et al. 2013) has shown that lower than potentially expected polymer
solution viscosity, and hence rather adverse mobility ratio, could lead to positive results. However, due
to the particularly high viscosity of extra-heavy oils, injecting polymer solutions with high viscosities
(typically around 100 cP) might be envisioned. The transport in porous media of polymer solutions of this
viscosity range are rarely studied in the literature and many of its aspects are not known. The purpose of
this work is to identify and quantify some of the specific transport properties of high viscosity polymer
solutions in porous media, with respect to conventional viscosities, in terms of adsorption/retention
behavior, in-depth propagation and injectivity.

Experimental Methods and Materials


Mostly, two types of polymers are extensively used in CEOR processes: polyacrylamides and polysaccharides (Green D.W. and Whillhite G.P. et al. 1998). Experimental studies have demonstrated the notable
viscosifying power of the partially hydrolyzed polyacrylamides (HPAM). The addition as little as few
hundreds parts per million of HPAM in brines can increase the solutions viscosity substantially (Carreau,
Ait-Kadi and Chauveteau et al. 1987; Green D.W. and Whillhite G.P. et al. 1998). This viscosifying
property of HPAM polymers represents a key parameter for mobility control in CEOR processes,
particularly for heavy and extra-heavy oils at moderate temperatures (less than 80C or 175F). Polyacrylamides hydrolyze when temperature exceeds 80C. The hydrolysis of HPAM results in its sensitivity
to calcium, as chemical degradation (Levitt, D., Pope G. and Jouenne, S. et al. 2011).
HPAM are long-linear and flexible chains, whose conformation are coils in water with a salinity high
enough to neutralize electrostatic repulsions between carboxylate groups (Chauveteau, G. et al. 1981). For

SPE-172850-MS

a model salinity brine equal to 20 g/L NaCl, the


macromolecular conformation is slightly expanded
coil (Chauveteau, G. et al. 1981). Polymer solubility
in water is due to the chemical nature of monomers
and it increases when there are charged groups on
the chain (polyelectrolytes), (Baviere et al. 1991).
Water-soluble HPAM polymers are obtained by hydrolysis of polyacrylamide (PAM) with sodium or
Figure 2Molecular structure of polyacrylamide and partially hydropotassium hydroxide Figure 2.
lyzed polyacrylamide.
A classical HPAM polymer, Flopaam 3630S
from SNF FLOERGER, was used for the laboratory
experiments of this study. Flopaam 3630S polymers
have typical molecular weight in the order of 20
million g/mol and 30% hydrolysis (Kamal et al.
2013). It was decided to work with polymer solutions concentrations ranging from 0.8 to 2.5 g/L, to
explore the range of viscosities that could be used
for CEOR operations for FPOs extra-heavy oil. So,
different HPAM solutions were prepared. The polymer (powder) was dispersed and dissolved slowly in
a standard brine (brine made of 20 g/L NaCl and
0.4 g/L NaN3 in Milli-Q grade water) using the
magnetic stirrer vortex method. Sodium azide
(NaN3) was added to prevent bacterial degradation
during the total duration of the experiments. The Figure 3Sandpack core holder and intermediate pressure taps layout.
standard brine was used thorough the study. The
HPAM solutions were filtered through Millipore
filters (20 m) in order to remove any debris or microgels (Kohler, N. and Chauveteau, G. et al. 1981)
which might provoke plugging of the porous medium. Afterwards, these HPAM solutions were degassed.
The concentration of each HPAM solution was determined with a carbon analyzer TOC-V CSH from
Shimadzu, used in Total Carbon (TC) mode and equipped with a Total Nitrogen (TN) analyzer. Coupling
TC and TN analysis (TC/TN method) allowed determining the degree of hydrolysis of the polymer and
hence to determine the polymer concentration from the TC or TN data.
Different laboratory experiments were performed to study the transport of polymer solutions in porous
media at different concentrations: polymer viscometry and coreflood tests involving the study of polymer
irreversible retention, inaccessible pore volume, in-depth transport and injectivity. For the present study,
the temperature has been set to 50C: this temperature is representative of the conditions encountered in
the FPO.
Viscosities of HPAM solutions have been measured with an imposed strain Low Shear rheometer
(LS30) equipped with a coaxial cylinder geometry.
Coreflood experiments have been performed on model sandpacks of high permeability. Such sandpacks
allowed obtaining repeatable results in terms of porosity and permeability from one experiment to another.
Monodispersed Fontainebleau sand grains (siliceous sandstones: 99 % of SiO2, with density of 2.65 g/cm3
and typical grain size of 91 m) were used to make each sandpack. A specific preparation process was
applied to treat these sand grains for removing fines as well as traces of metal oxides (Dupas et al. 2013).
The sand was then dry-packed in a core-holder with rigid walls and vacuum saturated with the standard
brine. As it can be observed in Figure 3, the core holder was equipped with intermediate pressure taps
allowing the measure the pressure drop on 3 internal sections of the porous medium: 1-5 cm, 5-9 cm and

SPE-172850-MS

Table 2Petrophysical properties of the sandpacks for the 4 coreflood experiments performed.
Permeability, k (D)
Experiment
CF1
CF2
CF3
CF4

Pore volume, PV (mL)

Porosity, (%)

Section 1-5 cm

Section 5-9 cm

Section 5-9 cm

Temperature (C)

39.15
38.87
40.18
38.80

39.7
39.2
40.3
39.5

4.7
4.5
4.9
4.2

4.3
4.5
5.4
4.2

4.4
5.8
-

50

9-13 cm; the length of the porous medium being approximately 14 cm (there are slights variation in the
porous medium length from one experiment to another due to changes of positions of the upstream and
downstream pistons). Table 2 shows the results of the petrophysical properties of the 4 sandpacks prepared
for the 4 coreflood tests reported in this paper. Porosity and permeability were approximately the same
for each sandpack (porosity39% and permeability4 D).
The brine and the HPAM polymer solutions were injected by using volumetric pumps. A backpressure
of 3 bar was applied during the injection of these fluids, in order to prevent the formation of bubbles. The
sandpacks were immersed into a thermostatic bath set at 50C. A capillary tube equipped with a pressure
drop sensor was set downstream the porous medium to determine the polymer breakthrough volume/time.
For each experiment, the standard injection procedure was as follows:
i. reference injection of the standard brine in both the core (measurement of the pressure drop PB)
and the capillary tube (PBcapi) at various flow rates and determination of the initial permeability;
ii. injection of the polymer solution at the reference flow rate Qref 3 mL/h (interstitial velocity,
0.3 m/day) in the capillary tube via the bypass line (PPcapi, enabling determination of the
reference maximum viscosity max PPcapi / PBcapi);
iii. first polymer slug: injection of the polymer solution at Qref in the core and the downstream
capillary tube (Pcapi, enabling determination of the effluent viscosity Pcapi / PBcapi) until
stabilization of all pressure drops, followed by chase reference brine;
iv. second polymer slug: injection of the polymer solution at Qref in the core and the downstream
capillary tube until stabilization of all pressure drops (this two slugs procedure allows determining
the irreversible retention of the polymer);
v. (optional) polymer injectivity study: increase of the injection flow rate up to a maximum value and
then stepwise decrease, injection of chase standard brine.
Mobility reduction (Rm - also called Resistance Factor, RF) and permeability reduction (Rk - also
called Residual Resistance Factor, RRF) for each flow rate are defined as the ratios of pressure drops
during (P) and after (PF) the polymer injection respectively, to pressure drops before polymer injection
(PB):
Eq 1
Eq 2
Rm represents the decrease in mobility of a polymer solution in comparison with the flow of the water
or brine in which it is prepared (Jennings, R. R. et al. 1971). Rm is a function of both the fluid and the
material rock: polymer molecular weight, water salinity, rock permeability and porosity, rock composition
and flow rate (Hirasaki and Pope et al. 1974). Rk reflects the permeability damage entailed by the
irreversible retention of the polymer. This is namely the polymer retained in the porous medium which
is not removed during a brine injection following the polymer. In this work, an effective relative viscosity

SPE-172850-MS

Table 3Concentration and Carreau-like fitting parameters of the flow curves for the 4 polymer solutions investigated.
Solution
Sol.
Sol.
Sol.
Sol.

1
2
3
4

Polymer concentration
Cpol (g/L)

r0

rinf

(s)

0.8
2.0
2.4
2.5

21.5
401
877
908

1.0
7.2
11.6
11.9

0.96
2.20
3.35
4.22

0.74
0.45
0.40
0.44

(reff) of the polymer solution flowing through the core is also defined, according to the Darcy law, as the
ratio between mobility and permeability reduction:
Eq 3
Polymer irreversible retention was computed from the shift between the two breakthrough curves
(effluent viscosity) during the injection of the two polymer slugs. Irreversible polymer retention is usually
attributed to irreversible adsorption of the polymer at the pore walls (Baviere et al. 1991). It is also worth
stressing that for all experiments, the total volume of polymer solutions injected represented more than ten
times the pore volume.

Results and discussion


Characterization of the polymer solutions
Powdered polymers are known to hold in water due to their hydrophilic nature (Zaitoun A. and Potie B.
et al. 1983). For this reason, the concentration of a polymer solution cannot be reliably determined by
weighting. In this work, it was determined by using the above described TC/TN method. This method led
to determining a degree of hydrolysis of the HPAM investigated of 26%. 4 polymer solutions were
prepared: their concentrations according to the TC/TN method are reported in Table 3. By comparing
estimated concentration from the mass of powder and TC/TN, estimated water contents between 12 and
13% were found.
Viscometry
The following viscosity was found for the standard brine at the temperature of the study (50C): brine
0.57 cP. The relative viscosity data as a function of the shear rate ( ) (flow curves) of the 4 polymer
solutions are presented in Figure 4 together, with the results of a fit with a Carreau-like law as follows:
Eq 4
were r0 and rinf are respectively the zero-shear and infinite-shear viscosities, a characteristic time
and n a characteristic exponent. The Carreau-like law fitting parameters are presented in Table 3. As
expected, the relative viscosity and the shear thinning behavior sharply increase when increasing
concentration. Sol. 2, 3 and 4 are representative of typical polymer solution viscosities that could be
applied for extra-heavy oil cases and Sol. 1 stands a reference conventional polymer solution. The
zero-shear rate plateau of the Carreau-like laws might not correctly account for the behavior of the
polymer solutions at very low deformation rate. More complex viscoelastic rheological behavior are
indeed likely to be observed, which are probably linked to the formation of specific microstructures in
semi-diluted solutions of very high molecular weight polymers. However, Carreau-like law fits such as
those presented in this work can be useful to model the flow curves for medium to high shear rates
(typically 1-100 s-1) which are typically representative of the average shear rates experienced by polymer
solutions flowing in oil reservoirs.

SPE-172850-MS

Figure 4 Relative viscosity vs. shear rate for polymer solutions together with the corresponding Carreau-like law fits.

Retention and inaccessible pore volume


Polymer retention corresponds to the amount of
polymer either trapped in the pore space or adsorbed
at the pore walls during the transport of polymer
solutions in porous media. Polymer molecules adsorb easily at solid interfaces such as petroleum
reservoirs rocks (Baviere 1991, Sorbie 1991). In
addition, different authors have proposed that the
polymer retention processes can be described in
analogy to the Langmuir adsorption by using Retention Isotherms (Ltsch et al. 1985). In this work,
the term retention is preferred to adsorption as it is
more general.
Physicochemical attraction between polymer and
pore walls may lead to significant retention. It could
involve two mechanisms (Baviere 1991, Sorbie Figure 5Ideal polymer breakthrough profiles for a polymer slug size
1.0 PV, taking into account both the polymer retention and IPV (after
1991): 1) irreversible adsorption of the injected par-
Dawson et al. 1972).
ticles to stationary surfaces in the porous media (in
this case, the grain surface of the sandpacks) and 2)
hydrodynamic trapping or reversible attachment of the injected particles to stationary surfaces (this type
of retention affects the speed of propagation of a polymer slug).
The determination of the polymer breakthrough profile permits the measurement of retention in porous
media, taking into account the portion of the pore volume that is not in contact with the polymer solution
(Dawson et al. 1972), which is called Inaccessible Pore Volume (IPV). The IPV is defined as the fraction
of the pore volume being inaccessible to large molecules but accessible to the small solvent and salt
molecules and ions. According to a conventional approximation, this inaccessible pore volume is assumed
to be occupied by water that contains no polymer. Figure 5 shows that polymer floods are affected by both
retention and IPV: The presence of IPV causes an early breakthrough of the polymer front (B), the

SPE-172850-MS

Figure 6 Polymer retention for polymer solutions at 2.4, 2 and 0.8 g/L and T50C.
Table 4 Irreversible retention and IPV results.
Experiment name

Polymer concentration (g/L)

Injected polymer solution

Pore volume, PV (mL)

(g/g rock)

IPV (%PV)

CF1
CF2
CF3

0.8
2
2.4

Sol. 1
Sol. 2
Sol. 3

39.2
38.9
40.2

5.4
32.9
62.9

0
2
5

retention delays the polymer front (C) and, the combination of adsorption and IPV reduce and shift the
polymer bank (D).
Determining the retention and IPV of polymer always involves the injection of tracers (Dawson et al.
1972, Ltsch et al. 1985 and Pancharoen et al. 2010) or a fine determination of the dead volumes of the
experimental setup to set the reference breakthrough at 1 PV. The irreversible part of the retention (usually
attributed to irreversible polymer adsorption) can be determined by measuring the differences in volumes
at breakthrough for the injection of two successive polymer slugs, separated by an extensive brine wash.
In this work, several experiments have been performed to study how polymer concentration affects
retention and IPV in porous media. For determining IPV and retention, it was not necessary to inject any
tracer into the porous medium because the dead volumes (upstream and downstream of the sand pack)
were measured accurately. Retention and IPV measurements were made in fresh sand packs for each
polymer concentration with similar permeability and porosity (for experiments CF1, CF2 and CF3) by
injecting two successive polymer slugs in each porous medium at a flow rate corresponding to an
interstitial velocity of 0.3 m/day. Breakthrough profiles were obtained through the capillary viscometer set
downstream the core. The polymer slugs lasted for a few PV typically, namely until full stabilization of
the Rm over all sections of the core was reached. Between the two polymer slugs, an extensive brine slug
was injected at the same flow rate until all Rm reached a full stabilization. The polymer breakthrough
curves obtained for Cpol 2.4 g/L, 2.0 g/L and 0.8 g/L are shown on Figure 6. The blue curves

SPE-172850-MS

Figure 7Polymer Retentions vs. polymer concentration at 50C.

Figure 8 Polymer adsorption vs. polymer concentration (after Zhang and Seright, 2013).

corresponds to the first polymer slug and the red curves to the second polymer slug. The S shape of the
breakthrough profiles is due to the dispersion of the polymer front (Perkins et al. 1963, Ltsch et al. 1985).
The fact that front part of the blue curves do not show this shape is very likely due to fact that polymer
retention takes place during the first polymer slug. For each slug, the volume at breakthrough was
determined at 50% of the maximum P in the capillary. These two volumes are noted V1 and V2 for the
first and the second slug, respectively. Based on the mechanism presented by Dawson et al. 1972, IPV and
Vret, the volume of polymer solution lost by irreversible retention are linked to V1 and V2 and the pore
volume Vp by the following relationships:
Eq 5
Eq 6
Vp being known, Vret and IVP can hence easily be determined. Note that Vret can directly be determined
by calculating V1 V2 and is hence not subjected to the uncertainties in the determination of Vp and the
dead volumes. , The amount of polymer irreversibly retained by unit mass of the porous medium is then
given by:
Eq 7

10

SPE-172850-MS

Figure 9 Over-retention test, experiment (CF2): 1) Injection of polymer solution at 2 g/L and 2) Injection of polymer solution at 2.4 g/L (after having
absorbed polymer at 2 g/L).

Figure 10 Over-retention test, experiment (CF3): 1) Injection of polymer solution at 2.4 g/L and 2) Injection of polymer solution at 2 g/L (after
having absorbed polymer at 2.4 g/L).

mrock being the mass of the dry porous medium.


Irreversible retention and IPV results are gathered in Table 4. It can be observed that the IPV are very
low, namely 0, 2 and 5 percent of the total pore volume for concentration of 0.8, 2 and 2.4 g/L,
respectively: these results are consistent with the high permeability of the porous media.
The striking result is that polymer retention increases with concentration, being higher at higher
injection concentration. varies from 5.4 to 62.9 g/g of rock, for polymer concentrations between 0.8
and 2.4 g/L, respectively (Figure 7). For CEOR processes at field scale, high polymer retention may
represent a considerable loss of polymeric material. High polymer retention is economically detrimental
for CEOR processes because of the high polymer costs. These results are in line with some recent results
(Zhang and Seright, 2013) also showing an increase of retention increases with concentration (Figure 8).
The results presented do not allow by themselves assessing that the marked increase of irreversible
retention with concentration is due to an increase of polymer adsorption at the pore walls. When
concentration is high enough, complex mechanisms (such as the formation of weak physical gels) could
take place in those area of the pore space in which hydrodynamic forces and hence deformation rates are
low: this could translate into significant amounts of the polymer being retained, almost irreversibly (this
could understood if weak physical gels are formed due to their viscoelastic behavior), in the porous media.

SPE-172850-MS

11

Table 5Over-Retention and IPV results.


Experiment name
CF2
CF3

Polymer concentration (g/L)


2
2.4
2.4
2

Injected polymer solution


Sol.
Sol.
Sol.
Sol.

Figure 11Mobility reduction, Rm, versus the number of pore volumes


injected (PVI) of polymer Sol. 1 (Cpol 0.8 g/L) injected during step ii)
(first polymer slug) of experiment CF1.

Over-retention effects
To further retention study, additional polymer slugs
at different concentrations were injected during a
second step of both CF2 and CF3, which were
exposed to additional polymer slugs at Cpol 2.4
g/L and 2 g/L, respectively (Figure 9 and Figure
10). The results gathered in Table 5 show that
injecting 2.4 g/L polymer solution after 2 g/L leads
to a significant over-retention but that injecting at 2
g/L after 2.4 g/L leads to a much lower retention of
the 2 g/L slug. This could potentially be translated
into a field injection strategy aimed at reducing
polymer adsorption: a sacrificial high concentration
polymer slug could be injected first, before a main
mobility control slug. It is important to note, in view
of economics, that the concentration in the sacrificial slug does not seem to need to be significantly
higher than that in the mobility control slug.
This result is apparently not in line with that
obtained by Zhang and Seright (Zhang and Seright,
2013) as these authors report that first contacting a
porous medium with low-concentration diluted
HPAM prevents further adsorption, but the range of
concentration for the successive slugs explored in
the present work is not the same as for Zhang and
Seright.

2
3
3
2

Pore volume, PV (mL)

g/g rock)

IPV (%PV)

38.9

32.9
114.8
62.9
6.2

2
15.4
5
7

40.2

Figure 12Mobility reduction, Rm, versus the number of pore volumes


injected (PVI) of polymer Sol. 2 (Cpol 2.0 g/L) injected during step ii)
(first polymer slug) of experiment CF2.

Figure 13Mobility reduction, Rm, versus the number of pore volumes


injected (PVI) of polymer Sol. 3 (Cpol 2.4 g/L) injected during step ii)
(first polymer slug) of experiment CF3.

Figure 14 Mobility reduction, Rm, versus the number of pore volumes


injected (PVI) of polymer Sol. 4 (Cpol 2.5 g/L) injected during step ii)
(first polymer slug) of experiment CF4.

12

SPE-172850-MS

Figure 15Comparison between equilibrium Rm values on section 5-9 cm of the cores, the bulk relative viscosity at the experiments equivalent shear
rate, rbulk, and the effective relative viscosity reff Rm/Rk for the 4 polymer concentrations injected as the first polymer slug of the 4 experiments.

Polymer in-depth transport


In-depth transport of a polymer solution reflects the ability for the polymer to be injected deep into a
porous media without creating filterability issues. It is a critical parameter for the success of a polymer
flooding operation as it is directly linked to capability to achieve high mobility reduction (Rm), and hence
to reduce the water/oil mobility ratio in-depth the reservoir. The quality of polymer in-depth transport can
be studied as the mobility reduction profile along a linear core. In this work, polymer in-depth transport
was assessed for different polymer concentrations during the injection of the first polymer slug (step ii of
0.3 m/day, which is typical of in-depth flow in
the experimental procedure), at the velocity
reservoirs. Results showing Rm on the internal sections of the 14-cm long cores versus the number of pore
volumes of polymer solution injected (PVI) for the 4 polymer concentrations investigated (corresponding
to 4 separated coreflood tests) are shown on Figure 11, Figure 12, Figure 13 and Figure 14, respectively
for polymer concentrations Cpol of 0.8 g/L, 2.0 g/L, 2.4 g/L and 2.5 g/L.
The first main observation to be made from these results is that, for all concentrations investigated, Rm
curves show a clear stabilization trend on all sections after a few PVI: this indicates that polymer in-depth
propagation is always satisfactory as no plugging due to in-depth filtration phenomena is observed. This
said, significant difference are observed for the various concentrations. A the lowest concentration of Cpol
0.8 g/L, the stabilized Rm values on all 3 internal sections are very close (Rm 17-18) whereas, when
polymer concentration is increased, discrepancies appear between the 3 sections. The following trend is
observed: Rm takes its lower value on the first, 1-5 cm section, reaches its maximum value on the
intermediate, 5-9 cm section and takes a lower value on the last, 9-13 cm section. Although these
observations do not question the quality of the in-depth propagation of the polymer, they show that
increasing the polymer concentration entails more complex transport mechanisms detailed study of these
mechanisms and their impact at reservoir scale being beyond the scope of this paper.
A critical concern for polymer flooding operations is the link between the bulk relative viscosity
(namely, the viscosity measured in a viscometer, which will be call rbulk in the following), and Rm which
corresponds directly to the reduction of the oil-water mobility ratio. Comparing rbulk and Rm requires at
first to estimate the equivalent shear rate corresponding to the low-velocity in-depth flow of polymer
solutions in reservoirs. This can easily be done by mean of a capillary-bundles representation of the porous
, writes (Chauveteau,
medium. According to this representation, the shear rate at each capillary wall,
1982):

SPE-172850-MS

13

Figure 16 Representation of the depletion layer in a capillary (corresponding to the porous mediums pore throats) of radius R, as a two-fluid model
with concentration Cdepl and viscosity depl in the depletion layer of thickness and bulk concentration Cbulk and viscosity bulk in the central zone.

Eq 8
where Q is the flow rate, S the section of the cylindrical porous medium, the porosity (the interstitial
velocity vi being vi Q/S), a corrective parameter reflecting the structure of the porous medium and
usually equals to 2.5 for granular media (Chauveteau, 1982) and rH the pores hydrodynamic radius given
by:
Eq 9
where k is the permeability and a corrective factor equals to 1.15 for granular media (Zaitoun et al.
1988, Chauveteau et al. 2002, Stalker et al. 2007). According to this representation, the equivalent wall
shear rate in the porous medium is found close to 4 s-1. Thanks to the Carreau law fitting of the flow
curves, rbulk corresponding to the polymer injection test can hence be computed and compared to the
stabilized Rm values. The corresponding results are shown on Figure 15 for the internal section 5-9 cm
of the cores (this section will be chosen as the studys section for the rest of the discussion).
It can be observed that Rm values are always higher that rbulk values. However, Rm does not reflect
only the viscosity of the polymer in the porous medium as it accounts also for the permeability impairment
entailed by the polymer. The most relevant estimation of the polymer effective viscosity in the core is
given by reff Rm/Rk, computed with Rk determined from the P during the brine injection following
the polymer injection. The plot of reff presented on Figure 15 shows that rbulk is always higher than reff
This effect, which appears only when polymer solutions are flowing in fine pores, has been widely
reported in the literature (Chauveteau et al. 1982, Uhl et al. 1986; Sorbie et al. 1991, Cannella et al. 1998,
Zamani et al. 2013), will be discussed and interpreted in the next section.
Depletion layer effects
Depending on the authors, the difference between the bulk and the effective viscosity is interpreted either
as a depletion layer effect or as a slipping effect. In this work, we investigate the case of a depletion layer
created by steric obstruction, or steric hindrance effects. In polymer solutions, steric obstructions reduce
the probability that macromolecular center of mass may be at a distance less than gyration radius from the
wall. Consequently, the segment density will increase from zero at the wall contact up to bulk density at
a distance related to macromolecular dimensions in the dilute and intermediate regimes (Chauveteau,
1982). In terms of hydrodynamics, this mechanism can be modeled by assuming that the polymer solution
is flowing in capillaries with radius corresponding to the porous mediums pore throat and that a coaxial
flow of two fluids with different viscosities takes place. This approach is illustrated in Figure 16 for a
capillary of radius R and two fluids with, repectively, a viscosity depl (corresponding to a polymer
concentration Cdepl and a relative viscosity rdepl) in the depletion layer of thickness and a viscosity bulk
(corresponding to a polymer concentration Cbulk and a relative viscosity rbulk in the central zone. The
integration along the capillary radius of the local viscous to pressure forces equilibriums, with a non-slip

14

SPE-172850-MS

Figure 17Principle of the numerical simulation performed to assess for a depletion layer caused by steric exclusion between hard discs.

or zero-velocity hypothesis at the capillary wall, leads to the following relationship between the
above-mentioned parameters and the effective relative viscosity reff namely the viscosity given by the
Hagen-Poiseuille law for the whole capillary:
Eq 10
In this work, it has been built a numerical approach to demonstrate the depletion layer effect, by
simulating the steric size-exclusion effect at the walls in a capillary tube with a 2D simulation for hard
round particles (discs) with repulsive short range interactions. The radial concentration variation inside the
tube (r/d) was calculated as follows:
Eq 11
Figure 18 shows the results of these simulations translated in terms of concentration (or volume
fraction, depl) of discs in the depleted layer versus bulk concentration or volume fraction bulk. It can
clearly be seen that depl is always lower than bulk: this proves the existence of a depleted layer due to
simple steric size exclusion. The linear adjustment presented, which is a priori the more relevant for the
concentration range of polymer solutions corresponds to depl 0.56 bulk.
For the following discussion, an upscaling hypothesis will be made. Namely, it will be assumed that
the relationship obtained for discs in 2D conditions (discs) is valid for 3D conditions (spheres). This said,
translating this relationship between concentrations in a relationship between viscosities in order to
achieve a prediction of reff as a function of the polymer concentration is rather easy as the viscosityconcentration relationship for the polymers investigated is already known. It is, however, more difficult
to propose an hypothesis on the thickness of the polymer depletion layer. In the case of hard discs, or
spheres according to the assumption made, the simulation results show that the typical thickness of the
depletion layer is the size the disc, or spheres. For the semi-diluted polymer solutions investigated, a
typical sizes of the polymers can obviously not be the hydrodynamic size of an individual macromolecule

SPE-172850-MS

15

Figure 18 Variation of the particle volume fraction depl (i.e. concentration) in the depleted layer as a function of bulk, the bulk (namely, far from
the depleted layer) particle volume fraction.

Figure 19 Irreversible permeability reductions, Rk, entailed by polymer adsorption as a function of the injected polymer concentration Cpol and
corresponding interpretations in terms of hydrodynamic polymer adsorbed layer thickness, H.

in diluted solution nor its radius of gyration. In this work, we propose to assume that the typical
hard-sphere equivalent size in the polymer solutions investigated is the same as the hydrodynamic
thickness of the polymer adsorbed layer. This thickness, H, can be computed from the Rk values. Indeed,
according to the literature, there are at least three mechanisms for permeability reduction: polymer
adsorption, gel formation and plugging. In case of simple and well-filtered polymer flow in porous media,
the permeability reduction might be interpreted as an adsorbed layer of polymer molecular coils that
reduces the effective size of the pores. This reduction of permeability reduces the mobility both of the

16

SPE-172850-MS

Figure 20 Ratio of effective to bulk relative viscosities reff/rbulk versus injected polymer concentration Cpol comparison between experimental
results on section 5-9 cm of the porous media (blue) and calculation (green).

polymer solution and the brine following the polymer solution and of the brine following the polymer
(Hirasaki G.J. and Pope G. A. et al. 1974). In the frame of the capillary bundle model, H can be linked
to Rk as follows:
Eq 12
The measured values of Rk and the corresponding H are plotted in Figure 19. These results show that
H strongly depends on the polymer concentration.
The assumption H enables a direct determination of a prediction of reff and its comparison with
the experimental reff. The corresponding results are shown in Figure 20, in terms of the ratio between the
effective and the bulk viscosities reff/rbulk. A fairly good agreement is found between the measured data
and the calculations based on the depletion layer model. These results show that the polymer depleted
layer hypothesis can lead to a relevant interpretation of the effective viscosity of polymer solutions
flowing in porous media at different concentrations. This, naturally, does not exclude the presence of other
mechanisms. In addition, further investigations would be needed to fully understand why the thickness of
the adsorbed layer can be associated to the hard sphere equivalent size.
In terms of practical polymer flooding operations, this work shows that relevant predictions of the
mobility reduction ability of polymers in reservoirs require a fine modeling of the transport properties of
the polymer. Furthermore, this study shows that concentration has a very significant impact on the
effective viscosity as for the highest concentration, only 50% of the bulk viscosity in actually achieved
in porous media. The consequence of irreversible polymer retention in terms of high Rk could appears as
a positive factor, as this leads to increasing Rm, but the marked permeability reductions observed at high
concentration should be taken with care: they could be beneficial to achieve some degree of conformance
control in-depth the reservoir but could also potentially entail injectivity issues.
Polymer Injectivity
Polymer injectivity is a particularly important aspect to assess prior to polymer flooding operations as it
directly consists in determining the injection pressure needed to achieve a given flow rate at the
near-wellbore. EOR polymers are non-Newtonian fluids and typically their viscosity decreases with
increasing the flow rate (rheo-thinning effect). Also, HPAM polymers are known to be sensitive to

SPE-172850-MS

17

Figure 21Typical transitory behavior (0.8 g/L, 50C).

Figure 22Injectivity study, Rm on section 5-9 cm of the sandpack versus interstitial velocity, vi for experiments CF1 and CF4, corresponding
respectively to the injection of polymer solutions at concentration Cpol 0.8 g/L and Cpol 2.5 g/L.

extensional deformation, leading to rheo-thickening (Sorbie 1991; Chauveteau et al. 1997; Seright 2009;
Dupas et al. 2013) in porous media. As a result Rm can either decrease (rheo-thinning effect) or increase
(rheo-thickening effect) with velocity.
The objective of the limited injectivity study performed in this work was to provide a first insight on
how an increase of polymer concentration could impact the injectivity. Experiments consisted in injecting
polymer solutions at various flow rate, covering the range of typical near-wellbore velocities, and
determining for each flow rate the equilibrium Rm value. This has been done for experiments CF1 and
CF4, corresponding respectively to polymer concentrations Cpol of 0.8 g/L and 2.5 g/L.
A first important observation concerns the transitory behavior in the P signals. Example results for
Cpol 0.8 g/L are shown in Figure 21, the two curves corresponding to the sandpack sections 1-5 cm and
5-9 cm. At high injection rates stabilization takes place after a marked transitory regime showing a
decrease in P (A): this could correspond to expulsion of retained polymer from the porous medium due

18

SPE-172850-MS

to hydrodynamic forces. Conversely, at lower injection rates (B and C), P stabilization takes place after
an increasing step: this could be due to the rather slow introduction of additional polymeric material in
the adsorbed layer. These first observations are important as they indicate that: (i) polymer injectivity is
likely governed by different mechanisms, which are not the same for the different ranges of flow rates,
and (ii) permeability modification due to the polymer is likely to play a role in governing polymer
injectivity that cannot be neglected with respect to polymer rheology.
Rm data versus interstitial velocities for the two concentrations investigated in the injectivity study are
shown in Figure 22. For both concentration, the classical trends with, from low to high velocities, a
Newtonian plateau, a rheo-thinning domain and a rheo-thickening domain are observed. For Cpol 2.5
g/L, rheo-thinning appears to be more marked and rheo-thickening less marked than for Cpol 0.8 g/L.
This observation is clearly a positive insight for polymer injectivity at high concentration. This said, for
each polymer flooding operation, polymer injectivity studies remain to be performed on a case by case
basis under representative conditions (in particular with reservoir rocks).

Conclusions
The purpose of this study was to better understand the mechanisms involved during polymer flooding for
extra-heavy oils reservoirs approaching the FPOs conditions. For this particular case, the transport
properties of viscous polymer solutions (relative viscosities in the range of 100-200 at shear rates
corresponding to in-depth flow in reservoirs) are very different to that of more conventional polymer
solutions (relative viscosities in the range of 10-20) in terms of:
1. irreversible retention which can be in the order of ten times higher when viscosity is increased ten
times. Adsorption at the pore walls might not be the only phenomena responsible for this behavior
and complex rheological properties of viscous semi-diluted polymer solutions exposed to low
deformations could also play a role;
2. over retention effect: a viscous polymer slug injected after a less viscous one can lead to
significantly higher retention. Fortunately, the reverse do not appears to be true as no over
adsorption is observed when concentration is decreased. This could translate into field injection
strategies to mitigate polymer adsorption (a short, concentrated, sacrificial slug then a large, less
concentrated, main slug);
3. In-depth transport and apparent viscosity: apparent viscosity becomes markedly lower (up to 50%
less) than bulk viscosity for viscous polymers solutions. This phenomena can be modeled by
depletion layer effects. Fortunately, at high viscosity, polymer mobility reduction Rm (i.e.
resistance factor, RF) is also driven by permeability reduction;
4. Injectivity: rheo-thinning appears to be more marked and rheo-thickening less marked for high
viscosity than for low viscosity. This could be a positive insight for field applications.
These specific properties of viscous polymer solutions transport in porous media illustrate that polymer
flooding FPOs extra-heavy oil has significant chances of success but should be treated with extra
attention. Retention, in-depth transport and injectivity need to be carefully taken into account and
quantified in case feasibility studies.

Acknowledgement
We thank PDVSA and IPEN for supporting this research project. Thanks to Philippe POULAIN and
Nicolas ROUSSEAU at IFPEN for their advices regarding the procedures for the lab experiments.

References
Ait-Kadi, A., Carreau P.J., and Chauveteau G. 1987. Rheological Properties of Partially Hydrolyzed
Polyacrylamide solutions. Journal of Rheology, 31(7), 537561.

SPE-172850-MS

19

Banerjee, Dwijen K. 2012. Oil Sands, Heavy Oil & Bitumen. Penn Well Corporation, USA.
Baviere. 1991. Basic Concepts in Enhanced Oil Recovery Process. SCI, Elsevier applied science.
Chauveteau, G. 1981. Molecular Interpretation of Several Different Properties of Flow of Coiled
Polymer Solutions Through Porous Media in Oil Recovery Conditions. SPE 10060. Society of
Petroleum Engineers of AIME.
Chauveteau, G. 1982. Rodlike Polymer Solutions Flow through Fine Pores: Influences of Pore Size
on Rheological Behaviour. Society of Rheology, Inc. Published by John Wiley & Sons, Inc.
Journal of Rheology, 26(2), 111142.
Chauveteau, G. and Norbert K. 1984. Influence of Microgels in Polysaccharide Solutions on Their
Flow Behavior Through Prous Media. Society of Petroleum Engineers of AIME.
Chauveteau, G., Tabary R. and Kohler, N. 1997. Rheology of Polyacrilamide-Based Systems Under
Near-Well Bore Conditions. Paper SPE 37299 presented at the SPE International Symposium on
Oilfield Chemistry, Houston, Texas, 18-21 February.
Chauveteau, G., Denys, K., and Zaitoun, A. 2002. New Insight on Polymer Adsorption Under High
Flow Rates. Paper SPE 75183 presented at the SPE Improved Oil Recovery Symposium, Tulsa,
13-17 April.
Cohen, Y. and Christ, C. 1986. Polymer Retention and Adsorption in the Flow of Polymer Solutions
Through Porous Media. SPE Reservoir Engineering, March 1986.
Dominguez, J.G. and Willhite G.P. 1977. Retention and Flow Characteristics of Polymer Solutions in
Porous Media. American Institute of Mining, Metallurgical and Petroleum Engineers, Inc.
Dawson, R. and Lantz, R. 1972. Inaccessible Pore Volume in Polymer Flooding. Paper SPE 3522
presented at SPE 46th Annual Fall Meeting, New Orleans, 3-6 October.
Dong, M., Ma, S., Li, A. 2011. Sweep Efficiency Improvement by Alkaline Flooding for Pelican Lake
Heavy Oil. Paper CSUG/SPE148971 presented at the Canadian Unconventional Resources &
International Petroleum Conference, Calgary, Canada, 15-17 November.
Dupas A., Hnaut I., Rousseau D., Poulain P., Tabary R. and Argillier J.-F. 2013. Impact of polymer
mechanical degradation on shear and extensional viscosities: towards better injectivity forecasts in
polymer flooding operations. Paper SPE 164083 presented at the SPE International Symposium on
Oilfiled Chemistry, The Woodlands, 8-10 April.
Fabbri C., Romero C., Aubertin F., Nguyen M., Hourcq S. and Hamon G. 2013. Polymer Flooding in
Extra-Heavy Oil: Gaining Information on Polymer-Oil Relative Permeabilities. Paper SPE
165237, Kuala Lumpur, Malaysia, 2-4 July.
Fabbri, C. Cottin, C., Jimenez J., Nguyen M., Hourcq S., Bourgeois M. and Hamon G. 2014.
Secondary and Tertiary Polymer Flooding in Extra-Heavy Oil: Reservoir Conditions Measurements-Performance Comparison. Paper IPTC 17703, Doha, Qatar, 20-22 January.
Farouq Ali, S.M. 1976. Non-Thermal Heavy Oil Recovery Methods. Paper SPE 5893. American
Institute of Mining, Metallurgical, and Petroleum Engineers, Inc.
Farouq Ali, S.M., Figueroa, J.M., Azuaje, E.A. & Farquharson R.G. 1979. Recovery of Lloydminster
and Morichal crudes by caustic, acid and emulsion floods. Journal of Canadian Petroleum
Technology, (pp. 5359). Montreal, January-March, 1979.
Green, D.W. and Willhite G.P. 1998. Enhanced Oil Recovery. Society of Petroleum engineers,
Richardson, Texas.
Hirasaki, G.J and Pope, G.A. 1974. Analysis of Factors Influency Mobility and Adsorption in the Flow
of Polymer Solution Through Porous Media. Paper SPE 4026, August 1974.
Huc, A. 2011. Heavy Crude Oils. Editions Technip, France.
Jennings R.R., Rogers J.H. and West T.J. 1971. Factors Influency Mobility Control By Polymer
Solutions. Paper SPE 2867. Journal of Petroleum Technology.
Lake, L.W. 1989. Enhanced Oil Recovery. University of Texas at Austin, USA: Prentice Hall.

20

SPE-172850-MS

Levitt D., Pope G. and Jouenne, S. 2011. Chemical Degradation of Polyacrylamide Polymers Under
Alkaline Conditions. SPE Reservoir Evaluation and Engineering, 2011. Paper SPE 129879
presented at the SPE Improved Oil Recovery Symposium, Tulsa, 24-24 April.
Levitt D., Jouenne J., Bondino I., Santanach-Carreras E. and Bourrel M. 2013. Polymer Flooding of
Heavy Oil Under Adverse Mobility Conditions. Paper SPE 165267 presented at the SPE Enhanced
Oil Recovery Conference, Kuala Lumpur, Indonesia, 2-4 July.
Li, Z. And Delshad M. 2013. Development of an Analitycal Injectivity Model for Non-Newtonian
Polymer Solutions. Paper SPE 163672 presented at the SPE Reservoir Simulation Symposium in
The Woodlands, Texas USA, 18-20 February.
Ltsch, T., Mller, T. and Pusch, G. 1985. The Effect of Inaccessible Pore Volume on Polymer
Coreflood Experiments. Paper SPE 13590 presented at the International Symposium on Oilfield
and Geothermal Chemistry, Phoenix, Arizona, 9-12 April.
Mai, A. & Kantzas, A. 2009. Heavy Oil Waterflooding: Effects of Flow Rate and Oil Viscosity.
Journal of Canadian Petroleum Technology, 48(3): 4251.
Manichand R.N. and Seright R.S. 2014. Field vs. Laboratory Polymer-Retention Values for a Polymer
Flood in the Tambaredjo Field. Paper SPE169027-PA, SPE Reservoir Evaluation & Engineering,
August 2014.
Kamal, M., Ali Hussien, I., Sultan A. and Han M. 2013. Rheological Study on ATBS-AM CopolymerSurfactant System in High-Temperature and High-Salinity Environment. Journal of Chemistry.
Volume 2013, Article ID 801570.
Norbert K. and Chauveteau, G. 1981. Xanthan Polysaccharide Plugging Behavior in Porous MediaPreferential Use of Fermentation Broth. Society of Petroleum Engineers of AIME.
Perkins, T.K., Johnston, O.C. 1963. A Review of Diffusion and Dispersion in Porous Media. Society
of Petroleum Engineers Journal.
Pancharoen, M. Thiele, M.R. and Kovscek, A. 2010. Inaccessible Pore Volume of Associative
Polymer Floods. Paper SPE 129910 presented at the Improved Oil Recovery Symposium, USA,
24-28 April.
Rojas, G., Simon, C., Capoferri, M., Redaelli, E., Marcano, E. and Solorzano, E. 2009. 3D Static
Model of Oligocene and Lower Miocene Oil Reservoirs, Junin 5 Block, Orinoco Heavy Oil Belt,
Venezuela. Paper 2009-WHOC09-414 presented at World Heavy Oil Congress, Puerto La Cruz,
Venezuela.
Romero, C., Aubertin, F., Fabbri, C., Nguyen, M., Hourcq, S. and Hamon G. 2013. Secondary
Polymer Flooding in Extra-heavy Oil Core Experiments under Reservoir Conditions and Core
Scale Simulation. Paper B02 presented at the 17th EAGE Improved Oil Recovery Symposium,
Saint-Petersburg, Russia, 16-18 April.
Seright, R.S. 2010. Potential for Polymer Flooding Reservoirs with Viscous Oils. Paper SPE 129899
presented at the SPE Improved Oil Recovery Symposium, Tulsa, USA, 24-28 April.
Seright, R.S. 2014. Polymer Flooding versus Gel Treatment. http://baervan.nmt.edu/randy/, Tulsa,
Oklahoma.
Sheng, J. 2011. Modern Chemical Enhanced Oil Recovery: Theory and Practice. Elsevier.
Sorbie, K.S. 1991. Polymer-Improved Oil Recovery. Blackie, Glasgow and London.
Stalker R., Butler K. and Grahan, G. 2007. Evaluation of Polymer Gel Diverters for a HighTemperature Field With Special Focus on the Formation Shape Factor-An Important Parameter for
Enhancing Matrix Placement of Stimulation Chemicals. Paper SPE 107806 presented at the
European Formation Conference held in Schevening, The Netherlans, 30 May-1 June.
Tabary R., Zaitoun A., Renard G. and Delamaide E. 2013. Pelican Lake Field: First Successful
Application of Polymer Flooding in a Heavy Oil Reservoir. Paper SPE165234 presented at the
SPE EOR Conference, Kuala Lumpur, Malaysia, 2-4 July.

SPE-172850-MS

21

Verzaro, F., Bourrel, M., Garnier, O., Zhou, HG. and Argillier, J.F. 2002. Heavy Acidic Oil
Trasnportation by Emulsion in Water. Paper SPE 78959 presented at the SPE International
Thermal Operations and Heavy Oil Symposium and International Horizontal Well Technology
Conference, Canada, 4-7 November.
Wang, J. & Dong, M. 2010. Simulation of O/W Emulsion Flow in Alkaline/Surfactant Flood for
Heavy Oil Recovery. Journal of Canadian Petroleum Technology, 49(6): 46 52), June 2010.
Wang, J., Dong, M. and Arhuoma, M. 2010. Experimental and Numerical study and Improving Heavy
Oil Recovery by Alkaline Flooding in Sandpacks. JCPT, March 2010.
Wassmuth, F., Green, K., Hodgins L. and Turta, A. 2007. Polymer Flood Technology for Heavy Oil
Recovery. Petroleum Society. Canadian Institute of Mining, Metallurgy and Petroleum. Presented
at the Petroleum Societys 8th Canadian International Petroleum Conference, Calgary, Canada,
June 12-14.
Wassmuth, F., Green, K. & Bai, J. 2012. Associative Polymers Outperform Regular Polymers
Displacing Heavy Oil in Heterogeneous Systems. Paper SPE157916 presented at the SPE Heavy
Oil Conference, Calgary, Canada, 12-14 June.
Zaitoun A. and Potie, B. 1983. Limiting Conditions for the Use of Hydrolyzed Polycralylamides in
Brines Containing Divalent Ions. Paper SPE 11785 presented at the International Symposium on
Oilfield and Geothermal Chemistry, Denver, CO-USA, June 1983.
Zaitoun A. and Kohler N. 1988. Two-Phase Flow Through Porous Media: Effect of an Adsorbed
Polymer Layer. Paper SPE 18085 presented at the 63rd Annual Technical Conference and
Exhibition of the Society of Petroleum Engineers, Houston, TX, 2-5 October.
Zamani N., Kaufman R., Skauge T. and Skauge A. 2013. Pore-Scale Modeling of Polymer Flow.
Paper A18 presented at the 17th European Symposium on Improved Oil Recovery, St. Petersburg,
Russia, 16-18 April.
Zhang G. and Seright R.S. 2013. Effect of Concentration on HPAM Retention in Porous Media. Paper
SPE 166265 presented at the SPE Annual Technical Conference and Exhibition, New Orleans,
Louisiana, USA, 30 September-2 October.
Zhang T.; Murphy M.; Yu H.; Bagaria H.; Youl Yoon, K.; Nielson B.M.; Bielawski W., Jonhnston P.;
Huh Ch. And Bryant L. 2013. Investigation of Nanoparticle Adsorption during Transport in Porous
Media. Paper SPE 166346 presented at the SPE Annual Technical Conference and Exhibition,
New Orleans, Louisiana, USA, 30 September-2 October.

Você também pode gostar