You are on page 1of 7

Journal of Petroleum Science and Engineering 105 (2013) 100–106

Contents lists available at SciVerse ScienceDirect

Journal of Petroleum Science and Engineering


journal homepage: www.elsevier.com/locate/petrol

Rheology and mechanical degradation of high-molecular-weight


partially hydrolyzed polyacrylamide during flow through capillaries
A.R. Al Hashmi a,n, R.S. Al Maamari a, I.S. Al Shabibi a, A.M. Mansoor a, A. Zaitoun b, H.H. Al Sharji c
a
Department of Petroleum and Chemical Engineering, Sultan Qaboos University, Oman
b
Poweltec, France
c
Petroleum Development Oman, Oman

art ic l e i nf o a b s t r a c t

Article history: High-molecular-weight partially hydrolyzed polyacrylamides are widely used in enhanced oil recovery.
Received 13 March 2012 Mechanical degradation of these polymers as a result of flow through pumps, chokes, valves and at the
Accepted 19 March 2013 sand face has been recently reported to have a negative impact on the application. This study presents
Available online 3 April 2013
capillary flow measurements of high-molecular-weight hydrolyzed polyacrylamide (degree of hydro-
Keywords: lysis ¼27.8%) in 2 wt% potassium chloride aqueous solvent. Flow was conducted at different flow rates
polyacrylamide through three different lengths of 125-μm stainless steel capillaries to investigate the apparent rheology
mechanical degradation and mechanical degradation. The apparent rheology was indicated by mobility reduction while the
entry point mechanical degradation was measured by the loss in viscosity of the solution effluent. The entry point
rheology
contribution in the overall mobility reduction and degradation was evaluated by extrapolation. In the
capillary
investigated range of shear rates, the polymer solution generally shows an initial apparent thickening
behaviour up to 15,000 s−1, above which the behaviour becomes thinning. After the evaluation of the
entry point contribution to the overall mobility reduction, the initial thickening behaviour is found to be
due to the coil–stretch transition at the entry point. The flow in the bulk of the capillaries is found purely
thinning in the whole shear rate range investigated. The total degradation is found to be almost constant,
below 20%, up to the shear rate of 15,000 s−1. The degradation starts to increase above 15,000 s−1 due to
the effect of shear in the bulk of the capillaries. The entry point degradation seem to contribute negligibly
(below 5%) to the overall degradation up to 100,000 s−1, above which it sharply increases its effect on
polymer degradation reaching 42% at 850,000 s−1. We believe that the results of this study will improve
the execution of the polymer enhanced oil recovery by minimizing polymer mechanical degradation.
High shear devices such as pumps, chokes, valves can have a detrimental effect on the mechanical
stability of the polymer and hence should be utilized with caution. Also, more mechanically stable
polymers can be utilized. Moreover, in the light of the current study, it is expected that the elongational
flow of polymer at the sand face and the flow of polymer through perforations and/or fractures can also
degrade the polymer, which need to be evaluated.
& 2013 Elsevier B.V. All rights reserved.

1. Introduction can be irreversibly degraded through chain scission and/or struc-


ture reformation by mechanical, chemical and biological mechan-
Polyacrylamide-based polymers are widely used in the fields of isms (Sorbie, 1991; Caulfield et al., 2002). The overall result is a
water treatment (Bolto and Gregory, 2007), paper manufacturing loss in the viscosifying power of the polymer, thus, the efficiency
(Wong et al., 2006), drag reduction (Manfield et al., 1999; Al-Sarkhi of the application is adversely affected (Tolstikh et al., 1992).
and Hanratty, 2001; Liberatore et al., 2003) and enhanced oil In oilfield polymer flooding, the initial degradation of the polymer
recovery (Sorbie, 1991; Seright et al., 2009). High-molecular- is due to mechanical degradation as the result of flow of polymer
weight hydrolyzed polyacrylamides are used to increase the solution through pumps, flow lines, chokes, valves, and the flow
viscosity of water for a better oil displacement in polymer through rock formation at the sand face (Zaitoun et al., 2012). Such
enhanced oil recovery (Sorbie, 1991). Polymer chains in solution flow restrictions may induce very high stresses along the chain
backbone, which can ultimately result in chain breakage (Keller
n
et al., 1987; Zaitoun et al., 2012). It was found that polymer
Correspondence to: Department of Petroleum and Chemical Engineering,
Sultan Qaboos University, P.O. Box 33, 123 Muscat, Oman. Tel.: +968 241 425 44;
experienced more than 65% loss of their initial viscosity as they
fax: +968 241 413 54. flow from the injectors to the producers in some field applications
E-mail address: azizra@squ.edu.om (A.R. Al Hashmi). (Wang et al., 2006). Noik et al. (1995) conducted degradation

0920-4105/$ - see front matter & 2013 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.petrol.2013.03.021
A.R. Al Hashmi et al. / Journal of Petroleum Science and Engineering 105 (2013) 100–106 101

experiments by flowing high-molecular-weight hydrolyzed poly- chloride (KCl). Sodium azide (NaN3) with a concentration of
acrylamide through glass bead pack investigating the effect of 400 ppm was added to the polymer solution as a stabilizer.
polymer molecular weight, concentration, and concentration and Three 125-μm, stainless steel capillaries having three different
nature of electrolytes. They found that degradation is more severe lengths of 10, 20, and 30 cm were used in capillary flow experiments.
for higher molecular weight, and polymer concentration. The capillaries were used as received without prior treatment. More-
The flow in abrupt flow field changes such as the converging over, there is no information regarding their roughness, hence, they
flow through capillaries and porous media induces elongation of were assumed to be smooth.
the flowing polymer coils (Boger, 1987; Sorbie, 1991). The elonga-
tion of the polymer chains in porous media depends on the 2.2. Methodology
polymer molecular weight, relaxation time, ratio of pore and
pore-throat radii (i.e. contraction ratio), and flow rate (Durst 2.2.1. Determination of overlapping concentration Cn and intrinsic
et al., 1982; Boger, 1987; Sorbie, 1991). In simple shear flow (such viscosity [μ]
as Poiseuille flow through a capillary or in Couette flow), full The overlapping concentration Cn measures the transition
extension and/or alignment of the polymer molecules cannot be between the dilute and the semi-dilute regimes. In the dilute
achieved (Georgelos and Torkelson, 1988). On the other hand, regime, the macromolecules are separated from each other and
flexible chains under the influence of such a flow field (i.e. simple behave independently. In the semi-dilute regime, the macromo-
shear flow) can be only slightly deformed. However, the coiled lecules are entangled and thus impose frictions on each other.
conformation of the polymer molecules in solution can be opened The transition between the dilute and the semi-dilute regime
up and partly extended at relatively high shear rates, which can is characterized by a change in the shape of the viscosity–
result in apparent shear thickening behaviour (Georgelos and concentration plot.
Torkelson, 1988). Elongational viscosity, which is defined as the A stock polymer solution at a concentration of 5000 ppm was
ratio of elongational stress to elongational strain rate, charac- prepared. The polymer solution was prepared by slowly adding the
terizes the resistance of a fluid to elongational deformation polymer granules powder to the shoulder of the vortex of the
(Sorbie, 1991).The elongational viscosity has been found to be water while maintaining vigorous stirring using a paddle mixer.
three times the simple shear viscosity (Boger, 1987). Elongational The paddle speed was decreased after 1 h of polymer addition and
flow through porous media can induce a shear thickening beha- the polymer solution was left under slow mixing overnight to
viour at Deborah number of 0.5 (Durst et al., 1982; Kulicke and ensure complete dissolution. Dilutions of the stock polymer
Hass, 1984; Sorbie, 1991). solution at different concentrations were conducted and the zero
The application of partially hydrolyzed polyacrylamide in EOR shear viscosity was measured for each concentration at 30 1C using
subjects the polymer to different shear rate values. The shear the Low-Shear 30 viscometer of Contraves.
′′
rate through a conventional pipe ð278 Þ can be in the range of The overlap concentration Cn was obtained by plotting the zero
−1
100–500 s . On the other hand, shear rates in the range of 5000– shear viscosity with respect to concentration of the polymer
15,000 s−1 may exist in coiled tubing and at the sand face of solution as shown in Fig. 2. The plot is log–log with two different
vertical wells. The highest shear conditions are met in chokes or
valves under high differential pressures, which can be in the range
of 50,000–100,000 s−1.
Recently, mechanical degradation of polymers used in
enhanced oil recovery was found to be partly responsible for
lower than anticipated efficiency in field applications (Wang et al.,
2006). Degradation was found to be greatest at the entry plane of
the injection sand face (Seright, 1983; Wang et al., 2006). More-
over, the studies of Ghoniem and co-workers have concluded that
polymer degradation occurs primarily at the entrance point where
stretching of the macromolecules is maximum and extensional
forces are sufficiently strong to cause bond scission (Ghoniem
et al., 1981; Ghoniem, 1988). The current study was conducted to
investigate the apparent rheological behaviour and mechanical Fig. 1. Chemical structure of the monomers composing the polymer.
degradation during the flow of high-molecular-weight, partially
hydrolyzed polyacrylamide through stainless capillaries with dif-
ferent lengths. Moreover, the contribution of the entry point to the
overall apparent thickening/thinning behaviour and mechanical
degradation was evaluated using an extrapolation technique.

2. Experimental

2.1. Materials

Hydrolyzed polyacrylamide, HPAM, which is a random copoly-


mer composed of 27.8% acrylate and 72.2% acrylamide monomers,
is used in this study with a nominal molecular weight of
18.5  106 g/mol. This polymer is supplied by SNF as one of its
commercial polymers widely used in enhanced oil recovery. The
chemical structure of the building monomers is shown in Fig. 1.
The polymer is dissolved in deionized water with 2 wt% potassium Fig. 2. Determination of the overlapping concentration.
102 A.R. Al Hashmi et al. / Journal of Petroleum Science and Engineering 105 (2013) 100–106

slopes at low and high concentrations. The overlap concentration Once overlap concentration and intrinsic viscosity have been
is the concentration at the intersection of these two linear slopes, obtained, the hydrodynamic radius (Rg) of the polymer coils in
which is 295 ppm for the polymer/solvent system used in solution can be calculated using the following equation (Buchholz
this study. et al., 2004):
Intrinsic viscosity [η] is a measurement of the viscosifying Mw
power of the polymer in a certain solvent. It is the reduced (ηR) cn ¼
ð4=3 Þπ N A R3g
and inherent viscosity (ηI) when concentration tends to zero
(Sorbie, 1991): where NA is Avogadro's number, and Mw is the molecular weight.
The characteristic relaxation time for non-dilute solutions (τc)
½η ¼ lim ηR
c -0 can be obtained using the following equation (Chauveteau et al.,
1984):
or
6 η −1 M w
½η ¼ lim ηI τc ¼ η r0
c -0 π2 s c RT

where where ηr0 is the zero shear relative viscosity.

η− ηs
ηR ¼ ; 2.2.2. Capillary flow
c ηs
For the capillary flow measurements, 850 ppm HPAM solution
ln ηr in 2 wt% KCl was prepared following the preparation procedure
ηI ¼ ; described above. Afterwards, the viscosity of the polymer solution
c
was measured at 30 1C using Bohlin Gemini Rheometer (MAL-
η VERN) with a double gap geometry (DG 24/27) in the shear rate
ηr ¼ ;
ηs range of 1–100 s−1 in 5 min. Then, the brine was injected at 30 oC
using a syringe pump (ISCO, Model 100 DX) in which the
η is the polymer solution viscosity, ηs is the solvent viscosity, ηr is
temperature of the injected liquid can be controlled. Once in the
the relative viscosity, and c is the polymer concentration.
piston of the pump at 30 1C, the brine was injected at different
Hence, the intrinsic viscosity can be obtained by plotting the
flow rates through the capillary while measuring the pressure at
reduced and inherent viscosities with respect to polymer concen-
times of stabilization. Afterwards, the polymer solution was
tration and extrapolating the linear curves obtained to the zero
injected at 30 1C at different flow rates. 10 ml of the effluent was
concentration as shown in Fig. 3.
collected at each flow rate and subsequently sealed. The viscosity
of this effluent was measured after one day to erase any flow
memory of the polymer chains due to the applied stresses in the
capillary. The viscosity of the effluents at the shear rate of 10 s−1
was compared with the initial viscosity. Mechanical degradation in
per cent (DR %) was calculated using the following equation:
 
μr0 − μrf
DR ð%Þ ¼  100%
μr0 − 1
where μr0 is the initial relative viscosity and μrf is the relative
viscosity of the effluent. This procedure was repeated for three
capillaries.
The pressure during the flow of solvent and polymer solution
was measured using low pressure transducer (LPT) (Druck, 0–
100 psi) for pressure values less than 100 psi. High pressure
transducer (HPT) (up to 7000 psi) was used for pressure values
higher than 100 psi. These pressure transducers were calibrated
before conducting the measurements. The overall experimental
Fig. 3. Determination of the intrinsic viscosity. set-up is shown in the schematic in Fig. 4.

Fig. 4. Schematic diagram of the experimental set-up.


A.R. Al Hashmi et al. / Journal of Petroleum Science and Engineering 105 (2013) 100–106 103

Table 2
Characterization of the HPAM in 2 wt% KCl.

Mw 18.5  106 g/mol


Hydrolysis 27.8%
[η]0 8173 cm3/g
Cn 295 ppm
Rg 292 nm
τc 0.094 s

Fig. 5. Schematic diagram of the contraction flow into the capillary.

Table 1
Shear rate values for the different flow rates used in
this study through the 125-μm capillary.

Flow rate Shear rate


(cc/h) (s−1)

1 1449
5 7247
10 14,490
50 72,470
100 144,900
300 434,800
500 724,700 Fig. 6. The evolution of mobility reduction for different capillary lengths with
700 1,015,000 respect to shear rate inside the capillary. Contribution of the entry point to the
800 1,160,000 mobility reduction is shown as well.
900 1,304,000
1000 1,449,000

Mobility reduction (Rm) is the ratio between pressure drop 3.1. Apparent rheology
during polymer solution flow (ΔPp) and solvent flow (ΔPs) at a
certain flow rate. Apparent viscosity during polymer solution flow Fig. 6 gives the mobility reduction obtained using the three
(ηp) can be calculated by multiplying the mobility reduction by the capillaries as a function of shear rate. Generally, the mobility
viscosity of the solvent (ηs) at a constant temperature: reduction initially increases with similar slopes for the three
capillaries reaching a critical shear rate (γmc) of 15,000 s−1, above
ΔP p ηp
Rm ¼ ¼ which it stabilizes for the 10-cm capillary and slightly decreases
ΔP s ηs
for the other two capillaries. In a third trend, the mobility
The flow through the contraction into the capillary is depicted reduction decreases steeply above 100,000 s−1. The three mobility
in Fig. 5. The entry region has an ID (Dent) of 1.753 mm, which curves converge to the same mobility reduction at around
makes the contraction ratio Dent/Dcap ¼14. Shear rate (γ) for a 400,000 s−1. Beyond this shear rate, the three curves diverge again
circular tube of radius (R) can be calculated using the following with different trends.
equation (Georgelos and Torkelson, 1988): The initial thickening behaviour has been attributed to differ-
ent mechanisms such as coil–stretch transition (Chauveteau et al.,
4U 4Q 1984; Georgelos and Torkelson, 1988), reversible, gel-like polymer
γ¼ ¼
R πR3 aggregates (Briscoe et al., 1999), concentration fluctuation
(Moldenaers et al., 1993), and/or various macromolecular associa-
where U is the average velocity, m/s, and Q is the volumetric flow
tions (Kishbaugh McHugh, 1993a, 1993b; Dupuis et al., 1994;
rate, m3/s.
Hu et al., 1995; Liberatore et al., 2003). In order to investigate
The shear rate upstream the entry section is negligible com-
about the most probable mechanism, the mobility reduction at the
pared to the shear rate inside the capillary, hence, shear rate in this
entry point was evaluated by extrapolating the mobility reduction
study will always indicate the one inside the bulk of the capillary.
curves with respect to capillary length at different flow rates
The shear rate values corresponding to the flow rates used in this
through capillaries as shown in Fig. 7.
study are listed in Table 1.
The evaluated entry-point mobility reduction is shown along
with the total mobility reduction for the three capillaries in Fig. 6.
The mobility reduction curves for the three capillaries due to shear
3. Results and discussion inside the capillaries as shown in Fig. 8 are evaluated by subtract-
ing the entry-point mobility reduction from the overall one.
Polymer intrinsic viscosity and overlapping concentration of Apparently, the behaviour is solely thinning in the bulk of the
the polymer solution at 30 1C are listed in Table 2 along with the capillaries. We can thus conclude that the apparent shear thicken-
calculated hydrodynamic radius and relaxations time for 850 ppm ing behaviour in the total mobility reduction up to 15,000 s−1 is
polymer concentration (non-dilute regime). due to the entry-point effect only.
104 A.R. Al Hashmi et al. / Journal of Petroleum Science and Engineering 105 (2013) 100–106

Fig. 7. Mobility reduction as a function of capillary length at different flow rates.

Fig. 9. Viscosity curves of the polymer solution after different passages at


increasing flow rates through the 10-cm capillary.

Fig. 8. Shear-thinning behaviour in the bulk of the capillaries.

3.2. Mechanical degradation

The flow curves of the polymer samples collected after different


flow rates through the 10-cm capillary are shown in Fig. 9.
The flow curves shift to lower values indicating a progressive loss
in the viscosifying power of the polymer. The mechanical degrada-
tion in percentage calculated for the 10, 20 and 30-cm capillaries Fig. 10. Polymer degradation as a function of shear rate during flow through
at different shear rates is shown in Fig. 10. Generally, for a certain 125-μm capillaries with different lengths.
capillary length, the degradation is almost constant, below 10%,
until a critical shear rate of 15,000 s−1, after which it increases shear rates are low to induce elongational flow in the rock matrix.
more steeply to higher values. It should be noted that degradation On the other hand, degradation due to shear in the bulk of the
through the tubing before the capillary entrance was evaluated. capillaries is almost the same and negligible below 15,000 s−1,
There was no degradation in the tubing below the flow rate of above which shear degradation starts to increase with trends
100 cc/h (144,900 s−1). However, tubing degradation was 2.5%, proportional to capillary length. The three shear degradation
3.8%, 4.3% and 4.9% at 100, 300, 500, and 600 cc/h, respectively. curves seem to reach almost constant values above 100,000 s−1.
These values are almost 5% of the degradation obtained using the
capillary. 3.3. Molecular processes
The entry point and shear contributions to the overall degrada-
tion are shown Fig. 11. Degradation due to the entry point is In order to draw a picture about the molecular level at the
constant and insignificant below the shear rate of 100,000 s−1, different flow regimes, the thickening/thinning behaviour and
above which entry-point degradation increases steadily showing mechanical degradation are interrelated. The apparent shear
no trend for stabilization within the interval of shear rates thickening behaviour up to the shear rate of 15,000 s−1 is probably
investigated. As noted above, the entry point degradation may due to the coil–stretch transition at the entry point due to abrupt
occur through chokes and at the sand face at high differential change in the shear rate (Chauveteau et al., 1984; Georgelos and
pressures. Polymers in EOR applications are usually injected at Torkelson, 1988). In our system, the onset of coil–stretch transition
a flow rate corresponding to speed of around 1 ft/day, at which is expected at 10.6 s−1 (i.e. at Deborah number of one), above
A.R. Al Hashmi et al. / Journal of Petroleum Science and Engineering 105 (2013) 100–106 105

lengths up to this shear rate is negligible due to elastic dissipation


of the imposed straining stress. The polymer chains seem to reach
their fully extension at the entry point in the range of shear rate
from 15,000 to 100,000 s−1, in which entry-point contributes
negligibly to the overall degradation. However, the straining stress
at the entry point seems to reach a critical value at 100,000 s−1,
above which chains' scission at the entry point occurs at higher
intensity with the increase of shear rate thereafter. This is reflected
as a thinning behaviour in the overall mobility reduction above
this critical shear rate given that the rheology inside the bulk of
the capillaries is predominately shear thinning.
It is recommended in light of the current study to thoroughly
investigate the polymer degradation prior to any EOR application.
Design and modifications of the flow system and operational
conditions can also minimize the detrimental effect of polymer
flow through devices that can induce strong flow fields such as
pumps, chokes, and valves. Furthermore, the application of more
mechanically stable polymers can enhance the efficiency of the
EOR application. Moreover, it is expected that the high degrada-
tion values reported between injectors and producers is, among
other possible factors, due to highly elongational flow of at the
sand face, perforations, and/or fractures.
Fig. 11. Entry point and shear contributions to the overall mechanical degradation
through 10, 20 and 30-cm capillaries.

which the polymer chains continue to be stretched at the entry Acknowledgement


point with the increase in shear rate, reaching the fully stretched
state at the plateau region above 15,000 s−1 (Keller et al., 1987). The authors would like to thank Petroleum Development Oman
Above 15,000 s−1, the chain backbone is able to withstand the (PDO) for providing the financial support to conduct this joint
associated straining stress up to 100,000 s−1, at which the staining work between Sultan Qaboos University and Poweltec.
stress at the entry point seem to be enough for chains' rupture
(Keller et al., 1987). Hence, the apparent thinning observed from
the overall mobility reduction above 100,000 s−1 is mainly due to
References
the scission of the polymer chains at the entry point into shorter
ones (i.e. lower molecular weight) (Noik et al., 1995; Buchholz
Al-Sarkhi, A., Hanratty, T.J., 2001. Effect of drag reducing polymer on annular gas–
et al., 2004; Stavland et al., 2010). liquid flow in a horizontal pipe, Int. J. Multiphase Flow 27, 1151–1162.
When the polymer chains are partly stretched at the entry Boger, D.V., 1987. Viscoelastic flows through contractions. Annu. Rev. Fluid Mech.
point at shear rates below 15,000 s−1, they resist degradation 19, 157–182.
Bolto, B., Gregory, J., 2007. Organic polyelectrolytes in water treatment. Water Res.
through their elasticity hence releasing any stress as extra mobility 41, 2301–2324.
reduction. This can explain the low entry and shear degradation Briscoe, B., Luckham, P., Zhu, S., 1999. Pressure influences upon shear thickening of
in this shear rate range. Above 15,000 s−1, more and more chains poly(acrylamide) solutions. Rheol. Acta 38, 224–234.
Buchholz, B.A., Zahn, J.M., Kenward, M., Slater, G.W., Barron, A.E., 2004. Flow-
seem to be degraded inside the bulk of the capillaries. This is induced chain scission as a physical route to narrowly distributed, high molar
because the partially elongated chains at the entry point, would mass polymers. Polymer 45, 1223–1234.
suffer greater and persistent stress with time inside the capillary Caulfield, M.J., Qiao, G.G., Solomon, D.H., 2002. Some aspects of the properties and
degradation of polyacrylamides. Chem. Rev. 102, 3067–3083.
leading to shear-induced degradation inside the bulk capillary
Chauveteau, G., Moan, M., Magueur, A., 1984. Thickening behaviour of dilute
in the shear rates between 15,000 and 100,000 s−1. Hence, polymer solutions in non-inertial elongational flows. J. Non-Newtonian Fluid
this shear-induced degradation increases with the length of Mech. 16, 315–327.
the capillary while degradation at the entry point is negligible Dupuis, D., Lewandowski, F.Y., Steiert, P., Wolff, C., 1994. Shear thickening and time-
dependent phenomena: the case of polyacrylamide solutions. J. Non-
up to 100,000 s−1. Above 100,000 s−1, the entry point starts to Newtonian Fluid Mech. 54, 11–32.
contribute in increasing proportions to the overall polymer degra- Durst, F., Haas, R., Interthal, W., 1982. Laminar and turbulent flows of dilute
dation reaching 42% at 850,000 s−1. polymer solutions: a physical model. Rheol. Acta 21, 572–577.
Georgelos, P., Torkelson, J., 1988. The role of solution structure in apparent
thickening behaviour of dilute PEO/water system. J Non-Newtonian Fluid Mech.
27, 191–204.
4. Conclusions Ghoniem, S., Chauveteau, G., Moan, M., Wolff, C., 1981. Mechanical degradation of
semi-dilute polymer solutions in laminar flows. Can. J. Chem. Eng. 59, 450–454.
Ghoniem, S.A.A., 1988. Effect of deformation sequence on rheological behaviour and
The current study has investigated the apparent rheology and mechanical degradation of polymer solutions. Chem. Eng. Commun. 63, 129.
degradation of high-molecular-weight partially hydrolyzed poly- Hu, Y., Wang, S., Jamieson, A., 1995. Rheological and rheooptical studies of shear-
thickening polyacrylamide solutions. Macromolecules 28, 1847–1853.
acrylamide using the stainless steel 125-μm capillaries. Semi- Keller, A., Muller, A., Odell, J., 1987. Entanglements in semi-dilute solutions as
dilute polymer solution of concentration 850 ppm at 30 1C was revealed by elongational flow studies. Prog. Colliod Polym. Sci. 75, 179–200.
injected through three capillaries with different lengths. Moreover, Kishbaugh, A.J., McHugh, A.J., 1993a. A rheo-optical study of shear-thickening and
structure formation in polymer solutions. Part I: experimental study. Rheol.
the contribution of the entry point to the apparent rheology and
Acta 32, 9–24.
degradation was evaluated by extrapolating the mobility reduction Kishbaugh, A.J., McHugh, A.J., 1993b. A rheo-optical study of shear-thickening and
and degradation to a zero capillary length. structure formation in polymer solutions. Part II: light scattering analysis.
Initial apparent shear thickening behaviour up to shear rate of Rheol. Acta 32, 115–131.
Kulicke, W.-M., Hass, R., 1984. Flow behaviour of dilute polyacrylamide solutions
15,000 s−1 is due to coil–stretch transition of the polymer coils at through porous media. 1. Influence of chain length, concentration, and
the entry point. The total mechanical degradation for all capillary thermodynamic quality of the solvent. Ind. Eng. Chem. Fundam. 23, 308–315.
106 A.R. Al Hashmi et al. / Journal of Petroleum Science and Engineering 105 (2013) 100–106

Liberatore, M.W., Pollauf, E.J., McHugh, A.J., 2003. Shear-induced structure forma- Stavland, A., Jonsbranten, H.C., Lohne, A., Moen, A., Giske, N.H., 2010. Polymer
tion in solutions of drag reducing polymers. J. Non-Newtonian Fluid Mech. 113, flooding-flow properties in porous media versus rheological parameters. In:
193–208. Proceedings of the EUROPEC/EAGE Annual Conference and Exhibition. Barce-
Manfield, C.J., Lawrence, C., Hewitt, G., 1999. Drag-reduction with additive in lona, Spain, SPE131103.
multiphase flow: a literature survey. Multiphase Sci. Technol. 11, 197–221. Tolstikh, L.I., Akimov, N.L., Golubeva, I.A., Shvetsov, I.A., 1992. Degradation and
Moldenaers, P., Yanase, H., Mewis, J., Fuller, G.G., Lee, C.-S., Magda, J.J., 1993. Flow- stabilization of polyacrylamide in polymer flooding conditions. Int. J. Polym.
induced concentration fluctuations in polymer solutions: structure/property Mater. 17, 177.
relationship. Rheol. Acta 32, 1–8. Wang, D., Han, P., Shao, Z., Chen, J., Seright, R.S., 2006. Sweep improvement options
Noik, C.H., Delaplace, P.H., Muller, G., 1995. Physico-chemical characteristics of for the Daqing oil field. In: The SPE/DOE Symposium on Improved Oil Recovery.
polyacrylamide solutions after mechanical degradation through a porous
Tulsa, OK, USA, SPE 99441.
medium. In: The International Symposium on Oilfield Chemistry. San Antonio,
Wong, S.S., Teng, T.T., Ahmad, A.L., Zuhairi, A., Najafpour, G., 2006. Treatment of
Texas, USA, SPE 28954.
pulp and paper mill wastewater by polyacrylamide (PAM) in polymer induced
Seright, R.S., 1983. The effects of mechanical degradation and viscoelastic behavior
flocculation. J. Hazard. Mater. 135, 378–388.
on injectivity of polacrylamide solutions. SPE J. 23, 475–485.
Zaitoun, A., Makakou, P., Blin, N., Al-Maamari, R.S., Al-Hashmi, A.R., Abdel-Goad, M.,
Seright, R.S., Seheult, M., Talashek, T., 2009. Injectivity characteristics of EOR
polymers. SPE Reservoir Eval. Eng. 12, 783–792. Al-Sharji, H.H., 2012. Shear stability of EOR polymers. SPE J. 17, 335–339,
Sorbie, K.S., 1991. Polymer Improved Oil Recovery. Blackie, Glasgow-London. SPE141113.

You might also like