Sie sind auf Seite 1von 8

International Journal of Engineering & Technology IJET-IJENS Vol:16 No:03

Rheological Approach of HPAM Solutions under


Harsh Conditions for EOR Applications
Bruno M. O. Silveira, Leandro F. Lopes, Rosngela B. Z. L. Moreno - University of Campinas
Abstract The hydrolyzed polyacrylamide (HPAM) is the
water-soluble polymer most often used in flooding applications.
However, the temperature, salinity and hardness of the
reservoirs affect its viscosifying properties reducing the sweep
efficiency of the polymer solution. This work aims to tailor
HPAM (Flopaam 5115SH) solutions prepared with synthetic
produced water (SPW) or with distilled water only, and evaluate
their rheological properties. The flow curves confirm the
detrimental effects of salinity and hardness. For the target
viscosity (10 mPa.s at ~7.8 s-1 and 23 C) chosen to perform
further polymer flooding tests, the required polymer
concentration is 1250 ppm. At this concentration, the solution
presents a thermothinning behavior as the temperature
increases. Regarding the viscoelastic properties, the viscous
modulus is predominant for fluids prepared with SPW, while the
elastic modulus is prevalent for fluids without salts. These results
contribute to the design of a polymer solution composition under
interest conditions.

Index Term Polymers, Viscosity, Rheology, Polymer flooding,


Enhanced Oil Recovery (EOR)

INTRODUCTION
Enhanced oil recovery (EOR) methods encompass advanced
techniques such as thermal, chemical, miscible and microbial.
They are used when the primary and secondary recoveries had
already been implemented, and a considerable portion of the
original oil in place (OOIP) remains in the reservoir [1-3]
The polymer flooding is a chemical method to enhance oil
recovery (CEOR). This method consists in changing the
rheological properties of the displacing fluid by the addition of
water-soluble polymers. The polymer addition increases the
viscosity and viscoelasticity of the injected/displacing fluid,
leading to a more favorable mobility. This mobility control
improves the sweep efficiency and also lowers the total
volume of water needed to achieve the residual oil saturation
[1], [2], [4].
Polyacrylamide-based polymers, such as partially hydrolyzed
polyacrylamide (HPAM), are widely used in CEOR, mainly
due to its availability and relatively low cost compared to
polysaccharides [5]-[8]. However, under high temperatures
and high salinity/hardness (HTHS) conditions, the
effectiveness of HPAM is severely affected [9], [10].
At low salinity solutions, the chains of HPAM are stretched
due to the charge repulsion of the carboxylic groups,
resulting in higher viscosities. When any cation is dissolved

in the solution, the negative charges are neutralized or


shielded, promoting a compression of the flexible chains,
resulting in a molecular shrinkage and thus a decrease in the
fluid viscosity [2], [9], [10].
Moreover, the viscosity reduction is more pronounced in the
presence of divalent cations such as calcium (Ca 2+) and
magnesium (Mg2+), than monovalent ones such as sodium
(Na+) and potassium (K+), when added in equivalent mass
percentage [9], [11]-[15].
This effect is explained by the ionic strength of the solution
(IS = 1/2mizi, where mi is the molar concentration of the
ion and zi is its charge). For example, the ionic strength of 1
molar solution of CaCl 2 is three times greater than that of
NaCl. Due to that, CaCl2 compresses the electrical double
layer of polymer chain more than NaCl [2].
Furthermore, the binding forces of the divalent ions are
stronger, owing to their higher charge and polarizability [2],
[10], [13]. Thus, the polymer chain can wraps in the
presence of low levels of Ca2+, due to its higher effectiveness
in shielding the negative charge of the polymer chains than
Na+ [9].
The detrimental effect on the viscosity of the HPAM
solution is also correlated to the temperature. At higher
temperatures, there is an increase in thermal motion of
molecules, leading to a decrease in the interaction time with
the neighboring polymer molecules. At this condition, the
intermolecular forces among polymer chains and the
viscosity of the polymer solution are decreased [16], [17].
Also, higher temperatures can increase the degree of
hydrolysis of the polymer, which converts amide groups
(CONH2) to carboxyl groups (COO). This effect introduces
more negatives charges to the polymer backbone, allowing
more interaction with divalent ions, which may cause their
precipitation, and thus promote a change in the rheological
properties of HPAM solutions [5], [10] ,[18], [19].
According to the literature, HPAM copolymerized with the
sulfonated
monomers,
such
as
a
2-acrylamidotertbutylsulfonic acid (ATBS group), is more resistant to
salinity and temperature [20], [21]-[24]. This resistance
contributes to minimize the losses on the rheological
properties, such as viscosity and viscoelasticity [21]. The
molecular structure of HPAM and HPAM with ATBS group
are illustrated in Fig. 1 (a) and (b), respectively.

161403-2828-IJET-IJENS June 2016 IJENS

IJENS

International Journal of Engineering & Technology IJET-IJENS Vol:16 No:03

(a)

(b)
Fig. 1. Molecular structure of (a) HPAM and (b) HPAM with ATBS group. Adapted [22], [25].

A rheological study allows setting/design the solution


composition under interest conditions [21], [24]. Focused on
this purpose, this work aims to evaluate some rheological
properties, such as viscosity, shear stress, overlap concentration
and viscoelastic characteristics of the polymer solutions tailored
with Flopaam 5115SH for a given synthetic produced water
(SPW) composition (high salinity and hardness conditions).
Additionally, the procedure was used to determine the fluid
composition with a target viscosity for further core flooding
through specific conditions. The fluid viscosity was set as ~10
mPa.s at 23 C for ~7.8 s-1 of shear rate, which can be
correlated to the shear rate that the polymer solution faces at
low flow velocity into the porous medium [26].
EXPERIMENTAL SECTION
Materials
The polymer fluids were prepared with an SPW composition.
On Table I is shown the type of salt, its concentration, and its
ionic strength.
The polymer used was Flopaam 5115SH (14-16 x10 6
g/mol of molecular weight (Mw) and 15% of hydrolysis
degree) from SNF. According to the supplier, this is an
ATBS-based polymer less sensitive to temperature and
salinity, recommended for reservoir temperatures up to

95C. Furthermore, this polymer has 15% of ATBS (2acrylamido-tertbutylsulfonic acid), and total anionicity of
25% [23].
Fluid Handling
The formulation of the polymeric fluids was based on APIRP-63 [27]. According to this standard, a stock solution with
5000 ppm of polymer concentration should be first prepared.
After that, the stock solution is diluted to obtain the desired
polymer concentration. It is also required that the solution
exhibits a homogeneous aspect, i.e., no insoluble particle
(fisheye) should be present.
To prepare the stock solution, the distilled water or the SPW
solution was vigorously stirred in a beaker within 30
seconds. Meanwhile, the polymer in powder form was
uniformly sprinkled into the vortex of the solution. Then, the
stirring speed was reduced in order to avoid mechanical
degradation of the polymer. The homogenization was kept
for three hours. After that, the stock solution was stored
overnight.
The stock solution was diluted with distilled water or SPW
until the desired polymer concentration was reached. The
solutions were put into a beaker and homogenized by a
magnetic stirrer at low speed (120 rpm) for 10 minutes.

Table I
Synthetic produced water (SPW) composition

Salt
CaCl2.2 H2O
MgCl2.6 H2O
KCl
NaCl
Na2SO4

Molecular weight (Mw)


(g/mol)

Concentration
(g/L)

147.01
203.3
74.55
58.44
142.04
TOTAL

10
6.3
0.6
86.6
1.3

Rheological apparatus and procedure


The rheological tests were performed using a HAAKE MARS
III rheometer, which is a high precision instrument. The

Ionic strength (mol/L)

(%)

0.204
0.093
0.008
1.482
0.027
1.814

11.25
5.12
0.44
81.67
1.51
100

sensor used was the double gap (DG41) because this


cylindrical sensor is preferable for low-viscous fluids. The
temperature control used was the THERMO HAAKE C25P

161403-2828-IJET-IJENS June 2016 IJENS

IJENS

International Journal of Engineering & Technology IJET-IJENS Vol:16 No:03

refrigerated bath with a Phoenix II Controller. A new sample


was applied at each test, and every data analyzed were within
the measuring range of the sensor.

The rheological behavior of the stock and diluted solutions,


prepared with distilled water or SPW, are presented in Fig. 2
and Fig. 3 respectively.

The analysis of viscosity, shear stress, temperature and


overlap concentration was made by flow curve tests, and the
last one was performed at the shear rate of interest. The
viscoelasticity analysis was performed by frequency sweep
tests. These tests require stress values within the linear
viscoelastic response (LVR), wherein the total resistance of
the material, at a given strain, is independent of the stress
imposed [28]-[30]. Thus, amplitude sweep tests were run for
each fluid, and the shear stress was chosen to perform the
frequency sweep.

In Fig. 2 (a) and Fig. 3 (a) it is shown that the apparent viscosity
of the fluids increases as the polymer concentration increase.
This behavior is related to the rise in the intermolecular
entanglement [2], [9], [31]. Also, the shear thinning behavior is
observed as the shear rate increases. This effect is due to the
uncoiling and the aligning of the polymer chains when they are
subjected to shearing [9], [31]. It can also be observed that the
salt content negatively affects the viscosity of the polymer
solutions, as described in the literature [2], [9], [11]-[15].

RESULTS AND DISCUSSIONS


Flow curves
Viscosity and shear stress versus shear rate of stock and dilute
solutions

From the shear stress versus shear rate curves (Fig. 2 (b) and
Fig. 3 (b)), the pseudoplastic behavior can be fitted by
Ostwald de Waele model (
. The rheological
parameters of each fluid according to this model, formulated
with distilled water or SPW, are summarized in Table II.
30

100000

Viscosity (mPa s)

10000

500 ppm
750 ppm
1000 ppm
1250 ppm
1500 ppm
3000 ppm
5000 ppm

10 ppm
20 ppm
30 ppm
50 ppm
75 ppm
100 ppm
300 ppm

25

Shear Stress (Pa)

10 ppm
20 ppm
30 ppm
50 ppm
75 ppm
100 ppm
300 ppm

1000

100

10

20

500 ppm
750 ppm
1000 ppm
1250 ppm
1500 ppm
3000 ppm
5000 ppm

15
10
5
0

1
0.1

10

100

Shear Rate (s-1)

(a)

1000

200

400

600

800

1000

800

1000

Shear Rate (s-1)

(b)

Fig. 2. (a) Viscosity and (b) shear stress vs. shear rate of dilutions with distilled water (at 23 C).

10 ppm
20 ppm
30 ppm
50 ppm
75 ppm
100 ppm
300 ppm
500 ppm
750 ppm

Viscosity (mPa s)

1000

30

1000 ppm
1250 ppm
1500 ppm
2000 ppm
2500 ppm
3000 ppm
3500 ppm
5000 ppm
SPW

100

10

20

1000 ppm
1250 ppm
1500 ppm
2000 ppm
2500 ppm
3000 ppm
3500 ppm
5000 ppm
SPW

15
10

5
1

0.1

(a)

10 ppm
20 ppm
30 ppm
50 ppm
75 ppm
100 ppm
300 ppm
500 ppm
750 ppm

25

Shear Stress (Pa)

10000

10

Shear Rate (s-1)

100

1000

(b)

200

400

600

Shear Rate (s-1)

Fig. 3. (a) Viscosity and (b) shear stress vs. shear rate of dilutions with SPW (at 23 C).

161403-2828-IJET-IJENS June 2016 IJENS

IJENS

International Journal of Engineering & Technology IJET-IJENS Vol:16 No:03


It can be observed that the higher the polymer concentration
is, the higher the parameter K will be, which will provide a
higher resistance for the fluid to flow. Furthermore, it implies
that the rheological behavior of the polymer solution will drift
away from the Newtonian behavior (n 1) [24], [30]. For
polymer solutions prepared with SPW, these effects are less
pronounced.
From Fig. 3 (a), one can see that 1250 ppm of polymer
concentration provides the target viscosity (~10 mPa.s at ~7.8
s-1) for flooding. This value is within the range of polymer
concentrations for field applications that usually ranges from
500 to 2500 ppm [32], [33].
An analysis of the influence of each salt on its given
concentration (Table I) is presented in Fig. 4 (a). The thermal
evaluation of the bulk composition is presented in Fig. 4 (b).
Both analyses were performed for solutions containing 1250
ppm of polymer concentration.

On Fig. 4 (a) is shown that the addition of any salt negatively


affects the apparent viscosity when compared with the fluid
without salt. Furthermore, it is possible to see the higher
detrimental effect caused by divalent ions. The fluids prepared
with MgCl2.6H2O or CaCl2.2H2O, even having lower ionic
strength values than that prepared with NaCl only, promoted a

similar reduction in the viscosity of the polymer solution. In


another work, similar results were obtained with Flopaam
AN110 SH (HPAM) from SNF, which is also an ATBS-based
polymer [24].
From Fig. 4 (b) the thermothinning behavior is observed with
the increase in the temperature [34]. This behavior is caused
by the reduction of the intermolecular forces among the
polymer chain [16], [17]. Taking the results at 23 C and 7.8 s-1
as a reference, the increase in the temperature to 30, 40, 50, 60
and 70 C reduced the solution viscosity in 20%, 34%, 43%,
47% and 49% respectively.
Overlap concentration (c*)
Plotting the viscosity versus polymer concentration for a
specific shear rate, two concentration regions can be
highlighted, a dilute and a semi-dilute.
In the dilute regime, the macromolecules behave
independently because they are separated from each other. In
the semi-dilute regime, frictions are imposed on the
macromolecules due to the polymeric entanglement. The
transition between both regimes is called overlap
concentration (c*) and its is characterized by a change in the
slope of the viscosity versus polymer concentration curve
[35].

Table II
Rheological parameters of the polymer solutions

Polymer concentration

K
(Flow consistency index)

n
(Flow behavior exponent)

r
(Correlation coefficient)

ppm

Distilled water

SPW

Distilled water

SPW

Distilled water

SPW

10

0.0016

0.0013

0.9551

1.0020

0.9999

1.0000

20

0.0020

0.0012

0.9364

1.0050

0.9998

1.0000

30

0.0031

0.0013

0.9011

1.0010

0.9992

1.0000

50

0.0055

0.0013

0.8249

1.0010

0.9989

1.0000

75

0.0057

0.0015

0.8627

0.9865

0.9976

1.0000

100

0.0084

0.0015

0.8200

0.9960

0.9969

1.0000

300

0.0757

0.0027

0.5520

0.9530

0.9971

1.0000

500

0.1528

0.0042

0.4912

0.9193

0.9966

1.0000

750

0.3204

0.0075

0.4207

0.8683

0.9954

1.0000

1000

0.5270

0.0100

0.3648

0.8525

0.9960

1.0000

1250

0.6906

0.0150

0.3487

0.8194

0.9957

1.0000

1500

0.8925

0.0197

0.3280

0.7959

0.9960

1.0000

2000

0.0312

0.7650

1.0000

2500

0.0468

0.7348

1.0000

3000

2.0630

0.0685

0.2972

0.7035

0.9925

1.0000

3500

0.0996

0.6731

1.0000

5000

3.6800

0.1948

0.2656

0.6350

0.9921

0.9996

161403-2828-IJET-IJENS June 2016 IJENS

IJENS

International Journal of Engineering & Technology IJET-IJENS Vol:16 No:03

10000

10000
HPAM - No Salt
HPAM - All Salts
HPAM + CaCl2
HPAM + MgCl2
HPAM + KCl
HPAM + NaCl
HPAM + Na2SO4

23 C
30 C

Viscosity (mPa s)

1000

Viscosity (mPa s)

100

10

1000

40 C
50 C
60 C

70 C

100

80 C

10

0.1

10

Shear Rate

(a).

100

1000

(s-1)

0.1

10

100

1000

Shear Rate (s-1)

(b)

Fig. 4. Viscosity vs. shear rate. (a) Influence of each salt on solutions with1250 ppm of HPAM at 23 C (a) (10g/L CaCl2.2H2O; 6.3 g/L MgCl2.6H2O; 0.6 g/L KCl; 86.6 g/L NaCl
and 1.3 g/L Na2SO4). (b) Temperature effect on solutions formulated with 1250 ppm of HPAM and SPW.

In Fig. 5 it is possible to see both regions for polymer


solutions prepared with distilled water or SPW. Dashed lines
represent the dilute region, and continuous lines represent the
semi-dilute one.
The c* of the polymer solutions prepared with distilled water
(Fig. 5(a)) or SPW (Fig. 5 (b)) were evaluated for fixed shear
rates. The c* values were defined in the intersections between
the dashed and the continuous lines. The parameters of each
trend line and the resulting c* are shown in Table III.
The solutions with SPW presented similar values for c*
(average value was 337.7 ppm at 23 C) regardless of the
shear rate, i.e., little change was observed within the shear rate
range under study.

This value of c* (337.7 ppm) is close to values reported by


some authors. For a solution prepared with HPAM (27.8%
acrylate, 72.2% acrylamide monomers and 18.5x106 g/mol of
Mw, supplied by SNF) and 2% of KCl at 30 C, 295 ppm was
obtained [35]. In another work, the solution with HPAM (SNF
Flopaam 3230S, 8x106 g/mol of Mw and 30% degree of
hydrolysis, approximately) and 2% NaCl at 25 C, 300 ppm
was reported for c* [36].
Concerning the Flopaam 5115SH solutions with distilled
water, one can see that the higher the shear rate, the higher
was the c* value. Similar results were reported for
polystyrene solutions, attributing this behavior to the
deformations of the chains induced by the flow of the
solutions [37].

1000
@ 7.848 s-1
@ 33.60 s-1
@ 88.59 s-1
@ 233.6 s-1
@ 615.8 s-1

Viscosity (mPa s)

Viscosity (mPa s)

1000

100

10

100

10

1
1

(a)

@ 7.848 s-1
@ 33.60 s-1
@ 88.59 s-1
@ 233.6 s-1
@ 615.8 s-1

10

100

1000

Polymer concentration (ppm)

10000

(b)

10

100

1000

10000

Polymer concentration (ppm)

Fig. 5. Critical overlap concentration (c*) for fluids prepared with: (a) distilled water and (b) SPW (at 23 C).

161403-2828-IJET-IJENS June 2016 IJENS

IJENS

International Journal of Engineering & Technology IJET-IJENS Vol:16 No:03

Table III
Parameters of the dilute and the semi-dilute concentration regions

Trend line 1*
(Dashed lines)
a
b

Shear Rate

7.848 s-1
33.60 s-1
88.59 s-1
233.6 s-1
615.8 s-1

0.317
0.410
0.509
0.591
0.641

7.848 s-1
0.932
33.60 s-1
1.002
88.59 s-1
1.025
233.6 s-1
1.060
615.8 s-1
1.061
* adjusted by potential trend line:

Trend line 2*
(Continuous lines)
a
b
No salts
0.061
1.131
0.048
1.016
0.050
0.919
0.055
0.822
0.056
0.745
SPW
0.0003
1.498
0.0011
1.249
0.0024
1.118
0.0056
0.976
0.0106
0.861

0.750
0.540
0.412
0.320
0.274
0.123
0.091
0.079
0.065
0.066

c* (ppm)

Viscosity
(mPa.s)

74.43
89.67
95.50
111.3
175.4

8.03
4.64
3.33
2.67
2.64

345.3
358.0
341.3
316.7
327.3

1.907
1.707
1.626
1.544
1.551

Viscoelastic Measurements (G and G)


The viscoelastic properties, especially the elastic modulus
(G), plays a significant role in increasing oil recovery owing
to its microscopic features [38], [39].

On Table IV are presented the shear stress chosen to perform

The amplitude sweep tests were performed for both conditions,


with and without salts. Among several results, some were
chosen to illustrate the polymer concentration and the salinity
effects on the LVR. The results are presented in Fig. 6.

For both conditions, the magnitude of the moduli of G and G


increased as the polymer concentration increased, such that
the viscoelastic characteristics of the fluids can be guided by
the type and concentration of the polymer [30].

Comparing the results between the fluids prepared with


distilled water (Fig. 6 (a)) and SPW (Fig. 6 (b)), it is evident
that the salinity affects the LVR. For the solutions with SPW,
LVR was not identified for the solutions with polymer
concentrations lower than 1250 ppm. Considering solutions
without salts, the LVR was not observed only for polymer
concentrations lower than 50 ppm.

Also, if the polymer concentration in the solution is higher


than c*, the elastic modulus will be dominant over the viscous
modulus [40]. This condition was observed for fluids without
salts, where the values of G were higher than those of G for
polymer concentrations over 100 ppm (see Fig. 7 (a)).
However, for fluids containing salts, the viscous behavior was
predominant for all polymer concentrations evaluated.

the frequency sweep test for each fluid composition, and Fig. 7
presents the elastic (G) and viscous (G) moduli as a function
of the angular frequency.

100

10

10

G' G'' (Pa)

G' G'' (Pa)

100

0.1

0.01

0.001

G"
100 ppm

1250 ppm

1250 ppm

5000 ppm

5000 ppm

0.1

0.01
G'

G"

50 ppm

50 ppm

100 ppm

100 ppm

1250 ppm

1250 ppm

5000 ppm

5000 ppm

0.0001
0.0001

(a)

G'
100 ppm

0.01

0.001

Shear stress (Pa)

0.0001
0.0001

100

(b)

0.01

100

Shear stress (Pa)

Fig. 6. Amplitude sweep test for (a) fluids without salts and (b) fluids formulated with SPW, both at 23 C.

161403-2828-IJET-IJENS June 2016 IJENS

IJENS

International Journal of Engineering & Technology IJET-IJENS Vol:16 No:03

Table IV
Shear stress choices to perform frequency sweep tests

Polymer concentration (ppm)


Fluids with distilled water
Fluids with SPW
< 50
< 1250
50 / 75 / 100
1250 / 1500 / 2000 / 2500 / 3000
300 / 500 / 750 / 1000 / 1250 / 1500
3500 / 5000
3000 / 5000
-

Shear Stress (Pa)


No LVR
0.01
0.1
1.0
100

100

10

10

G' G'' (Pa)

G' G'' (Pa)

0.1
0.01
0.001
G'
50 ppm
100 ppm
1250 ppm
5000 ppm

0.0001
0.00001
0.01

(a)

0.1

10

0.01
0.001

G"

G"

G'

50 ppm
100 ppm
1250 ppm
5000 ppm

100

0.1

0.0001
0.00001
0.01

1000

(Rad/s)

0.1

1250 ppm

1250 ppm

2500 ppm

2500 ppm

5000 ppm

5000 ppm

10

100

1000

(Rad/s)

(b)

Fig. 7. Frequency sweep test for (a) fluids without salts and (b) fluids formulated with SPW, both at 23 C.

CONCLUSIONS
The solutions prepared by mixing HPAM Flopaam 5115SH
with either SPW or distilled water presented a shear thinning
behavior (pseudoplastic behavior) that can be fitted by the
Ostwald de Waele model. It was observed that the salt content
in the polymer solutions reduces the polymer viscosity and the
shear stress. Furthermore, this detrimental effect was more
severe for salts with divalent cations (Ca2+ and Mg2+) in their
respective concentration.
It was found that the salt content directly influences the
determination of the overlap concentration (c*). The polymer
solutions prepared with distilled water indicate that the
overlap concentration increases as the shear rate increases,
unlike the results observed for fluids containing salts. For
these fluids, the overlap concentration was almost independent
of the imposed shear rate.
From the viscosity versus shear rate data, we determined that
the polymer concentration required to obtain the target
viscosity was 1250 ppm at 23 C (considering 10 mPa.s at 7.8
s-1 as shear rate).
The rheological evaluation also indicates that, at this polymer
concentration (1250 ppm), thermothinning behavior is
identified as the solution temperature increases. From room
temperature (23 C) to 70 C at 7.8 s-1, the decrease in the
viscosity of the polymer solution was around 49%.

Regarding the viscoelastic properties of the polymer solutions,


it was observed that the viscous behavior (G) was predominant
for the fluids prepared with SPW. For fluids without salt, the
elastic behavior (G) was more pronounced, except for fluids
with polymer concentrations lower than 100 ppm.

ACKNOWLEDGMENTS
The authors wish to thank STATOIL, ANP, CEPETRO and
the Division of Petroleum Engineering/DE-FEM-UNICAMP
for their support in this work.

REFERENCES
[1]

[2]
[3]

[4]

[5]

RAMIREZ, W. F. Application of Optimal Control Theory to Enhanced


Oil Recovery, Developments in Petroleum Science, v. 21, Elsevier. The
Netherlands, 1987.
SORBIE, K. S. Polymer-Improved Oil Recovery. 1st Ed., Boca Raton:
CRC Press, p. 37-79, 312-340, 1991.
VEERABHADRAPPA S. K., URBISSINOVA, T., KURU, E. Polymer
screening for EOR application - A rheological characterization
approach. Paper SPE 144570 presented at SPE Western North
American Regional Meeting held in Anchorage, Alaska, 7-11 May, 11
p., 2011.
CHEN Q, WANG Y, LU Z, FENG Y. Thermoviscosifying polymer
used for enhanced oil recovery: rheological behaviors and core flooding
test. Polymer Bulletin, v. 70, n. 2, p. 391-401, 2012.
LAKE L. W. Enhanced Oil Recovery. New Jersey, USA: Prentice
Hall. p. 1-12, 314-336, 1989.Osterloh, W. T., Law, E. J. SPE 39694,
1998.

161403-2828-IJET-IJENS June 2016 IJENS

IJENS

International Journal of Engineering & Technology IJET-IJENS Vol:16 No:03


[6]

[7]

[8]

[9]

[10]

[11]

[12]

[13]

[14]

[15]

[16]

[17]

[18]
[19]

[20]

[21]

[22]

[23]

[24]

MOREL, D., LABASTIES, A., JOUENNE, S., NAHAS, E. Feasibility


Study for EOR by Polymer Injection in Deep Offshore Fields (IPTC11800). In Proceedings of International Petroleum Technology
Conference, Dubai, UAE, 4-6 December, 2007.
OSTERLOH, W. T., LAW, E. J. Polymer Transport and Rheological
Properties for Polymer Flooding in the North Sea Captain Field. Paper
SPE 39694 presented at Improved Oil Recovery Symposium held in
Tulsa, Oklahoma. 19-22 April, 12p. 1998.
ABIDIN, A. Z., PUSPASSARI, T., NUGROHO, W. A. Polymers for
Enhanced Oil Recovery Technology. Procedia Chemistry, v. 4, p. 1116, 2012.
NASR-EL-DIN, H. A., HAWKINS, B. F., GREEN, K. A. Viscosity
Behavior of Alkaline, Surfactant, Polyacrylamide Solutions Used for
Enhanced Oil Recovery, paper SPE 21028, SPE International
Symposium on Oilfield Chemistry, Anaheim, California, USA, 20-22
February, 1991.
SHENG, J., J. Modern Chemical Enhanced Oil Recovery - Theory and
Practice. Gulf Professional Publishing. Elsevier, p. 51-64; 101-117;
153-164, 2011
MUNGAN, N., SMITH, F.W., THOMPSON, J.L. Some Aspects on
Polymer Floods, paper SPE 1628, JPT, vol. 18, n. 9, pp. 1143-1150,
September, 1966.
MUNGAN, N. Shear Viscosities of Ionic Polyacrylamide Solutions,
paper SPE 3521, SPE Journal, vol. 12, n. 6, pp. 469-473, December,
1972.
CHANG, H. L. Polymer Flooding Technology - Yesterday, Today, and
Tomorrow, paper SPE 7043, Journal of Petroleum Technology, v. 30,
n. 8, p. 1113-1128, August, 1978.
MELO, M. A., SILVA, I. P. G., GODOY, G. M. R., SANMARTIM, A.
N. Polymer Injection Projects in Brazil: Dimensioning, Field
Application and Evaluation, paper SPE 75194 presented at SPE/DOE
13th Symposium on Improved Oil Recovery held in Tulsa, Oklahoma,
13-17 April, 2002.
MANDAL, A., OJHA, K. Optimum Formulation of AlkalineSurfactant-Polymer Systems for Enhanced Oil recovery. Paper SPE
114877 presented at Asia Pacific Oil & Gas Conference and Exhibition
held in Perth, Australia, 20-22 October, 12p. 2008.
SAMANTA, A., BERA, A., OJHA, K., MANDAL, A. J. Effects of
Alkali, Salts, and Surfactant on Rheological Behavior of Partially
Hydrolyzed Polyacrylamide Solutions. Journal of Chemical &
Engineering Data, v. 55, n. 10, p.4315-4322, 2010.
JUNG, J. C., ZHANG, K., CHON, B. H., CHOI, H. J. Rheology and
Polymer Flooding Characteristics of Partially Hydrolyzed
Polyacrylamide for Enhanced Heavy Oil Recovery. Journal of Applied
Polymer Science, v. 127, n.6, p. 4833-4839, 2013.
GREEN, D W; WILLHITE, G P. Enhanced Oil Recovery. Richardson,
Texas, USA: SPE Textbook Series, v. 6, p. 60-72, 100-168, 1998.
RASHIDI, M., BLOKHUS, M. A., SKAUGE A. Viscosity and
Retention of Sulfonated Polyacrylamide Polymers at High
Temperature. Journal of Applied Polymer Science, v. 119, n. 6, p.
3623-3629, 2011.
THOMAS, A., GAILLARD, N., FAVERO, C. Some Key Features to
Consider When Studying Acrylamide-Based Polymers for Chemical
Enhanced Oil Recovery. Oil & Gas Science and Technology - Rev. IFP
Energies nouvelles, v. 67, n. 6, p. 887-902, 2012.
KAMAL, M. S., HUSSIEN I.A., SULTAN A.S., HAN M. Rheological
Study on ATBS-AM Copolymer-Surfactant System in HighTemperature and High-Salinity Environment. Journal of Chemistry, v.
2013, p. 1 - 9, 2013.
TOVAR, F. D., BARRUFET, M. A., SCHECHTER, D. S. Long term
stability of acrylamide based polymer during chemically assisted CO2
WAG EOR. Paper SPE 169053 presented at Improved Oil Recovery
Symposium held in Tulsa, Oklahoma, USA, 12-16 April, 9p. 2014.
SERIGHT, R. S., SKJEVRAK, I. Effect of dissolved iron and oxygen
on stability of HPAM polymers. Paper SPE 169030 presented at
Improved Oil Recovery Symposium, held in Tulsa, Oklahoma, USA,
12-16 April, 12 p. 2014.
LOPES, L. F., SILVEIRA, B. M. O., MORENO, R. B. Z. L.
Rheological Evaluation of HPAM fluids for EOR Applications.
International Journal of Engineering & Technology IJET-IJENS, v. 14,
n. 3, p. 35-41, 2014.

[25] LITTMAN, W. Polymer Flooding: Developments in Petroleum


Science. Elsevier, Amsterdam, v. 24, p. 14-18, 1988.
[26] MELO, M. A., HOLLOBEN, C. R. C., SILVA, I. P. G., BARROS
CORREIA, A., SILVA, G. A., ROSA, A. J., LINS, A. G., LIMA, J. C.
Evaluation of Polymer Injection Projects in Brazil. Paper SPE 94898
presented in Latin American and Caribbean Petroleum Engineering
Conference, Rio de Janeiro, Brazil, 20-23 June, 17p. 2005.
[27] American Petroleum Institute. Recommended practices for evaluation
of polymers used in enhanced oil recovery operations: API
recommended practices 63 (RP 63), Washington, D.C., 1990.
[28] BARNES, H. A.; HUTTON, J. E.; WALTERS, K. F. R. S. An
introduction to rheology. 3th ed. Amsterdam: Elsevier Science
Publishers B. V., p. 1-25, 1993
[29] XIN, X., XU, G., WU, D., LI, Y., CAO, X. The effect of CaCl2 on the
interaction between hydrolyzed polyacrylamide and sodium stearate:
Rheological property study. Colloids and Surfaces A: Physicochem.
Eng. Aspects, n. 305, p. 138-144, 2007.
[30] SILVEIRA, B. M. O., MORENO R. B. Z. L. Rheological Roles on the
Dynamic Behavior of Drill-in Fluid Invasion and Oil Permeability
Restoration of the Damage Zone. Paper OTC-24523-MS presented in
Offshore Technology Conference - Brazil held in Rio de Janeiro,
Brazil, 29-31 October, 10p. 2013.
[31] MISHRA, S., BERA, A., MANDAL, A. Effect of polymer adsorption
on permeability reduction in enhanced oil recovery. Journal of
Petroleum Engineering, v. 2014, p. 1-9, 2014.
[32] ALUHWAL, O. K. H. Simulation Study of Improving Oil Recovery by
Polymer Flooding in a Malaysian Reservoir. 2008. 185p. Master thesis
(Master of Engineering Petroleum) - Faculty of Chemical and Natural
Resources Engineering, Universiti Teknologi Malaysia, Malaysia,
2008.
[33] XIAODONG, K., JIAN, Z. Offshore Heavy Oil Polymer Flooding Test
in JZW Area. Paper SPE 165473 presented in Heavy Oil ConferenceCanada held in Calgary, Alberta, Canada, 11-13 June, 8p. 2013.
[34] LIU, X., WANG, Y., LU, Z., CHEN, Q., FENG, Y. Effect of inorganic
salts on viscosifying behavior of a thermoassociative water-soluble
terpolymer based on 2-acrylamido-methylpropane sulfonic acid.
Journal of Applied Polymer Science, v. 125, p.4041-4048, 2012.
[35] AL HASHMI, A. R., AL MAAMARI, R. S., AL SHABIBI, I. S.,
MANSOOR, A. M., ZAITOUNB, A., AL SHARJI, H. H. Rheology
and mechanical degradation of high-molecular-weight partially
hydrolyzed polyacrylamide during flow through capillaries. Journal of
Petroleum Science and Engineering, v. 105, p.100-106, 2013.
[36] ZHANG, G., SERIGHT, R. S. Effect of Concentration on HPAM
Retention in Porous Media. SPE Journal, SPE 166265, v. 19, n. 3, p. 18, 2014.
[37] PAPANAGOPAULOS, D., PIERRI, A., DONOS, A. Influence of the
shear rate, of the molecular architecture and of the molecular mass on
the critical overlapping concentration C*. Polymer, v. 39, n. 11, p.
2195-2199, Elsevier, 1998.
[38] ZHANG, L., YUE, X. Mechanism for viscoelastic polymer solution
percolating through porous media. Journal of Hydrodynamics, v.19,
n.2, p.241-248, 2007.
[39] ZHANG, Z., LI, J., ZHOU, J. Microscopic roles of "viscoelasticity" in
HPMA polymer flooding for EOR. Transp Porous Med, v. 86, p.199214, 2011.
[40] PERTTAMO, E. K. Characterization of Associating Polymer (AP)
Solutions - Influences on flow behavior by the degree of
hydrophobicity and salinity. Master thesis (Petroleum Technology Reservoir Chemistry), University of Bergen, Bergen, Norway, 2013.

161403-2828-IJET-IJENS June 2016 IJENS

IJENS

Das könnte Ihnen auch gefallen