Sunteți pe pagina 1din 16

J186103 DOI: 10.

2118/186103-PA Date: 12-January-18 Stage: Page: 18 Total Pages: 16

Polymer Stability After Successive


Mechanical-Degradation Events
S. Jouenne, Total E&P; H. Chakibi, IFP Energies Nouvelles; and D. Levitt, Total E&P

Summary
A key challenge in polymer-flood forecasting is the prediction of polymer stability far from the injector. Degradation may result from
various mechanical-degradation events in surface facilities and at the wellbore interface, as well as possible oxidative degradation
caused by the presence of oxygen and reduced transition metals. All these steps must be closely examined to minimize degradation and
ensure propagation of a viscous polymer solution.
In this paper, polymer solutions are pushed toward degradation rates that would be unacceptable for enhanced-oil-recovery applica-
tions to better understand the underlying physics. Multistep degradation events are induced in various geometries, such as capillaries,
blenders, and porous media.
For the geometries and range of polymer and salt concentrations investigated, degradation (as defined here) approaches an asymp-
totic value as the number of degrading events increases. An empirical normalization method is proposed, allowing superimposition of
curves of viscosity loss vs. time across multiple possible geometries. The normalization procedure is applied to predict the extent of
degradation during a field injection in which near-wellbore degradation occurs after degradation in surface facilities. We predict that
degradation in the porous medium reaches a stable value after passing through approximately 6 mm of rock.
Finally, degradation is proposed as a tool to probe the molecular-weight distribution and to narrow the polydispersity of polymers,
which can be used for maximizing both viscosifying power and injectivity simultaneously.

Introduction
Compared with biopolymers such as xanthan, hydrolyzed polyacrylamides (HPAMs) are more prone to mechanical degradation
(Seright et al. 2009; Zaitoun et al. 2011). This behavior is attributed to the conformation of the macromolecules. Xanthan is structured
in solution as a double-strand helix and can be considered to behave as a rigid rod in solution and will align in the direction of the flow.
HPAM, on the other hand, is a flexible chain with a coiled conformation that can be stretched by flow, leading to the macroscopic obser-
vation of viscoelasticity associated with chain extension and rupture (Chauveteau 1981; Magueur et al. 1985).
When injecting HPAM for mobility-control application, the level of degradation experienced in surface facilities, wellbore, and res-
ervoir must be known to accurately predict the increased oil recovery caused by improved mobility control. This degradation will result
from oxidative and mechanical degradation in surface facilities along with mechanical degradation near the wellbore where the strain
rates experienced are highest. Mechanical degradation of HPAM in the concentration range of polymer-flooding applications has
been extensively studied in porous consolidated (Maerker 1975; Seright 1983; Martin 1986) and unconsolidated media (Maerker 1976;
Ghoniem 1985; Southwick and Manke 1988; Noik et al. 1995). Different correlations were proposed to predict screen factor or viscosity
loss as a function of the permeability and the length of the porous medium. Seright (1983) proposed to neglect the length dependence
and to consider that mechanical degradation is confined to the sandface. In these studies, salinity, concentration, and molecular-weight
dependence were investigated. However, most studies examine polymers as received by manufacturers, rather than following the signif-
icant degradation representative of passage through surface facilities.
When passing through a degrading geometry, only a fraction of the polymer chains are degraded. Additional passes through similar
geometries (or degrading events) cause additional chain ruptures and viscosity loss. We refer to the dependence of cumulative viscosity
loss on the number of degrading events as “degradation kinetics” or “transient-degradation curve.”
The goal of this paper is to illustrate these kinetic aspects of polymer degradation, as well as to examine the implications when the
solution subsequently enters the reservoir. Degradation kinetics of HPAM solutions passing through different geometries (abrupt con-
traction, mixer, consolidated porous medium) are compared. An empirical method is then proposed to predict degradation resulting
from consecutive degrading events. The influence of physical-chemistry parameters such as viscosity, temperature, solvent quality, and
polymer concentration is not the focus of this paper, though it is well-known that these parameters can have a first-order effect on poly-
mer-solution viscosity. Because of various field-specific constraints, these parameters vary somewhat between experiments. Readers
are therefore encouraged to perform experiments by use of the approach described herein with parameters appropriate to their applica-
tion before extending this paper’s conclusions.

Brief Description of the Physics of Polymer Mechanical Degradation


Shear flow is the superposition of rotational and extensional flow. As a result, a flexible polymer coil will stretch in the flow direction,
rotate because of the velocity gradient in the perpendicular direction, and contract. In flows in which fluid is locally accelerated through
contractions, porous media, valves, or when turbulences are present, the extensional component of the velocity gradient can dominate
over the rotational component. As a result, coils are more extensively stretched. As predicted by De Gennes (1974), polymer coils in
dilute solution under extensional flow will experience a sudden coil-stretch transition if the strain rate (_e ) is higher than the inverse of
the longest relaxation time of the unperturbed random coil. The critical strain rate at which the coil-stretch transition occurs was found
to scale as e_ C / 1=Mw1:5 , which is consistent with Zimm relaxation time. This relaxation time is affected by the nature of the frictional
contact between molecule and solvent, which depends on molecular shape, solute/solvent interaction, and solvent viscosity (Keller and
Odell 1985).
As the chain starts to extend, hydrodynamic force exerted through the friction of the solvent on the stretched chain increases, and
can lead to chain rupture at a critical strain rate e_ R . Deformation and rupture depend upon the type of extensional flow (stagnation point

Copyright V
C 2018 Society of Petroleum Engineers

Original SPE manuscript received for review 5 September 2016. Revised manuscript received for review 30 March 2017. Paper (SPE 186103) peer approved 4 April 2017.

18 February 2018 SPE Journal

ID: jaganm Time: 18:15 I Path: S:/J###/Vol00000/170047/Comp/APPFile/SA-J###170047


J186103 DOI: 10.2118/186103-PA Date: 12-January-18 Stage: Page: 19 Total Pages: 16

or transient flow) and the chemistry of the polymer chain (Hunkeler et al. 1996a, b). By use of devices with a stagnation point such as
opposed jets, cross slots, or four-roll mills (Odell and Keller 1986; Odell et al. 1990), it was observed that chains may be completely
stretched before breaking at their center, and that the critical strain rate for scission scales as e_ R / 1=Mw2 . This scaling can be predicted
by applying Stokes law to an extended string of beads. By summing the friction force acting on each bead in a purely extensional flow,
the maximum tension T exerted at the center of the chain is predicted to scale as T / g_e N 2 , where g is the solvent viscosity and N is the
bead number. Temperature dependence of the scission process and central breaking are predicted if bond scission is considered as a
thermally activated process, the energy barrier of which is decreased by the tension exerted on the bond.
In a transient flow, such as flow through a contraction, chains are submitted to a distribution of strain rates and residence times
throughout the extensional-flow field, which begins several orifice diameters away from the contraction. Contrary to stagnation-point
flow, for which a very-small fraction of the chains are effectively passing into the stagnation point, most of the chains are experiencing
the extensional-flow field. As a result, the residence time in the degrading-flow field is larger and results in degradation higher by two
orders of magnitude at the same strain rate. In transient flow, critical strain rate for scission was found to scale as e_ R / 1=Mw (Nguyen
and Kausch 1988).
Whatever the type of flow (transient or stagnation point), there exists a strain rate below which chain scission does not occur, no
matter the residence time in the flow field. In a study on polymer degradation through orifices of different diameters, Nguyen and
Kausch (1991) observed that degradation strongly depends on the fluid velocity in the orifice but not on the magnitude of the strain rate.
As degradation resulting from passage through orifices is a unique function of the entrance-pressure drop, chain rupture is related to the
energy accumulated in the coil during the deformation process. A hybrid yo-yo model of chain extension with a surface-energy-accu-
mulation mechanism was proposed to reconcile the observed differences between transient- and stagnation-point flows (Hunkeler et al.
1996b). In this model, some parts of the chains remain in a coil state and are linked with multiple taut strands. Stress is transmitted
from the surface to the center of the coils, where energy is accumulated. Hence, friction and the resulting chain-scission process is no
longer related to the strain rate experienced by the polymer, but rather to the total accumulated strain, corresponding to the total elastic
energy stored. This model is compatible with the observed midchain scission [measured by gel-permeation chromatography by Nguyen
and Kausch (1988, 1991)] before full chain extension (Hunkeler et al. 1996a). This concept of degradation associated with the dissi-
pated energy was successfully applied to the degradation kinetics of drag-reducing agents by Henaut et al. (2012). Recently, Vanapalli
et al. (2006) proposed a model of polymer degradation in turbulent flow that unifies the molecular-weight dependency of the critical
strain rate for rupture in stagnation points and transient experiments. In their model, degradation results from the strain experienced in
turbulent eddies at the capillary entrance, rather than from the strain of the extensional flow resulting from the converging geometry.
Several studies agree on the impossibility to break macromolecules in simple laminar shear flow (Nguyen and Kausch 1986; Odell
et al. 1990). Because of the combination of the rotational and extensional components in laminar shear flow, chains are partially ori-
ented in the flow direction and rotate on themselves. Because of the limited residence time under extensional flow at a given position,
chain extension does not occur. Flow must be continuously extensional for the coil to be stretched. In capillary experiments, this type
of flow will occur if there is a contraction at the entrance or if the flow is turbulent.
Few studies address the degradation of semidilute solutions for which coils are entangled (Chow et al. 1988; Müller et al. 1992).
The critical strain rate e_ C of the coil-stretch transition in the semidilute region decreases with polymer concentration and degradation is
concentration dependent, as opposed to dilute solutions, where e_ C and degradation are concentration independent. In addition, it was
observed by gel-permeation chromatography that scission is randomized, contrary to the precise midchain scission observed in dilute
regime (Müller et al. 1992). Finally, it was found that monodisperse-polymer solutions are more shear resistant than polydisperse-poly-
mer solutions. Contrary to dilute solutions, for which the tension exerted on the chain is caused by the solvent/monomer friction, in
semidilute solutions, stress is transmitted by junction points of the network. As a consequence, stress per chain and entanglement den-
sity c/c* must be taken into account (Nghe et al. 2010; Dupas et al. 2012).

Experimental
HPAM in powder form with approximately 30% hydrolysis from the Flopaam series (SNF Floerger, Andrezieux, France) was used for
all experiments. Molecular weights given by the supplier were used as summarized in Table 1.

Mw (10 g/mol)
6
Polymer Grade
3130S 3.5
3230S 6.5
3330S 9.5
3530S 14.5
3630S 18.5

Table 1—Mw given by the supplier for Flopaam series HPAM.

Three brine compositions of 0.4, 6, and 35 g/L total dissolved solids were used. All the experiments were performed at ambient temperature.
A 10,000-ppm mother solution was first prepared by stirring the solution for 2 hours with a mechanical stirrer. This solution was
then diluted to the desired concentration in a beaker by gentle stirring with a magnetic stirrer. In this study, the concentration range
investigated is 100 to 1,400 ppm, although most of the experiments were performed between 800 and 1,400 ppm, which corresponds to
the semidilute regime. For the experiments by use of a blender, solutions were filtered through a 100-mm nylon screen. For the experi-
ments in which solutions were injected through a capillary or a porous medium, solutions were filtered in series through three consecu-
tive nitrocellulose Millipore filters with 8-, 5-, and 1.2-mm average pore size (Dfilter ¼ 142 mm) at a constant flow rate of 6 cm3/h to
avoid a plugging issue (Chauveteau and Kohler 1984).
Degradation was evaluated by viscosity measurement. Depending on the type of experiment, either a Brookfield DV-I viscometer
with ULA adapter or a ProRheo low shear LS300 rheometer was used. With the Brookfield viscometer, viscosity was measured in the
shear-rate range 7.3–122 s1. Because of the poor sensitivity of this viscometer at low shear rates, viscosity at 73 s1 was chosen as a
reference. With the low shear rheometer, viscosity was measured in the shear-rate range of 0.1 to 50 s1. With this very-sensitive rhe-
ometer, it was possible to accurately determine the low shear Newtonian viscosity, and viscosity at 0.1 or 7 s1 was used as a reference.

February 2018 SPE Journal 19

ID: jaganm Time: 18:15 I Path: S:/J###/Vol00000/170047/Comp/APPFile/SA-J###170047


J186103 DOI: 10.2118/186103-PA Date: 12-January-18 Stage: Page: 20 Total Pages: 16

Different devices were used to mechanically degrade polymer solutions: capillaries, laboratory blender, and porous media. The ex-
perimental protocol of each technique is briefly described below.
In single-pass capillary or porous-medium experiments, the polymer solution was injected at different flow rates with an Isco 1000D
pump through:
• An abrupt contraction 2/0.5 mm created by the junction of a PEEK tube with an inner diameter (ID) Dt ¼ 2 mm, in which was
inserted a stainless-steel capillary with ID DC ¼ 0.5 mm. Pressure drop through the capillary was measured with a Rosemount
3051S pressure transducer.
• A consolidated porous medium (ceramic frit from OFITE, Dcore ¼ 25.4 mm, Lcore ¼ 6.3 mm, kw ¼ 8,740 md, / ¼ 0.43).
Flow rate was increased step by step. For each step, samples were collected at the exit of the capillary or the porous medium for vis-
cosity measurement. Degradation was evaluated as a function of the volumetric-flow rate Q, the DP through the capillary, the velocity
at the capillary entrance VC , or the shear rate at the capillary wall c_ c .
In multiple-passes capillary or porous-medium experiments, the polymer solution was injected many times at constant flow
rate through:
• An abrupt contraction 1/0.17 mm created by the junction of two PEEK capillaries with IDs of 1 and 0.17 mm.
• A porous medium (PEEK frit IDEX OC820, Dcore ¼ 10 mm, Lcore ¼ 1.6 mm, kw ¼ 115 md, / ¼ 0.28).
A Nemesys syringe pump was used. A special setup was built to measure inline the viscosity of the polymer solution after each pas-
sage through the degrading geometry. Degradation was measured as a function of the number of passes for different flow rates.
In degradation in a blender, the polymer solution was sheared at constant rotational speed in a laboratory blender. The solution was
sampled at different time intervals for viscosity measurement. In this experiment, the shearing time and the rotational speed were the
two parameters controlling the degradation.
The degradation of polymer solution corresponding to the extent of viscosity loss, gloss was calculated by use of
g0  gdeg
gloss ¼  100; . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð1Þ
g0  gH2 O

where g0 is the viscosity of the nondegraded solution, gdeg is the viscosity of the degraded solution, and gH2 O is the water viscosity
(taken as 0.9 cp at 20 C).
Polymer solutions being non-Newtonian, the degradation of a polymer solution calculated with Eq. 1 depends on the shear rate at
which polymer viscosity is measured. In this paper, several series of experiments are presented, corresponding to different project con-
ditions over the years. Depending on the series, viscosity at 0.1, 7, or 73 s1 is used.

Results and Discussion


Identifying the Physical Origin of Polymer Degradation. Results of single-pass-capillary experiments are presented in Fig. 1. For
this experiment, polymer solutions of 3630S at 800 ppm in a 35 g/L brine were injected at constant flow rate in capillaries of different
lengths with a constant internal diameter (DC ¼ 0.5 mm) connected to tubes of internal diameter (Dt ¼ 2 mm).

80 80
Degradation at 73 s–1 (%)

Degradation at 73 s–1 (%)

70 70
60 60
50 50
40 40
30 30
Dt = 2 mm; Dc = 0.5 mm; L = 5 cm Dt = 2 mm; Dc = 0.5 mm; L = 5 cm
20 Dt = 2 mm; Dc = 0.5 mm; L = 10 cm 20 Dt = 2 mm; Dc = 0.5 mm; L = 10 cm
Dt = 2 mm; Dc = 0.5 mm; L = 20 cm Dt = 2 mm; Dc = 0.5 mm; L = 20 cm
10 10 Dt = 2 mm; Dc = 0.5 mm; L = 30 cm
Dt = 2 mm; Dc = 0.5 mm; L = 30 cm
0 0
0 10 20 30 0 200 400 600
ΔP (bar) Flow Rate (cm3/min)

Fig. 1—Degradation of an 800-ppm 3630S polymer solution in 35-g/L brine through capillaries (DC 5 0.5 mm) of varying lengths
connected to an upstream tube (Dt 5 2 mm). (Left) Degradation vs. pressure drop through the capillary; (right) degradation vs. flow
rate. Degradation is calculated from viscosity measurements at 73 s21.

As seen on the left of Fig. 1, degradation increases with the pressure drop. For a given pressure drop, degradation increases when
the length of the capillary is decreased. When plotting degradation as a function of flow rate, all the curves are superimposed. At con-
stant flow rate, degradation is thus independent of the capillary length, as already observed by Culter et al. (1975), Ghoniem et al.
(1981), and Stavland et al. (2016), whereas Al Hashmi et al. (2013) observed a small dependence on length. This independence demon-
strates that degradation occurs at the entrance of the capillary, which corresponds to an abrupt contraction. Fluid is accelerated from Vt
to VC when passing from the tube with Dt ¼ 2 mm into the capillary with DC ¼ 0.5 mm. The ratio of velocities is 16 (according to Eq.
4, VC =Vt ¼ D2t =D2C ¼ 22/0.52). Polymer coils are stretched and break because of the strain and the velocity differential. As seen in Ap-
pendix A, the inlet-tube diameter, and thus the inlet velocity, have negligible effects on polymer degradation.
When degradation is plotted as a function of the pressure drop, degradation is the highest with the shortest capillary. The measured
pressure drop is indeed the sum of the entrance-pressure drop, the exit-pressure drop, and the frictional-pressure drop with the capillary
wall. For capillaries of different lengths, the entrance and exit terms are approximately equal, whereas the frictional-pressure drop is
proportional to the length (as a first approximation because entrance and exit effects can propagate at some distance into the capillary).
To obtain the same pressure drop between two capillaries of different lengths, it is thus necessary to increase the flow rate in the shortest
capillary. As a consequence, for a fixed pressure drop, degradation is higher in the shortest capillary.

20 February 2018 SPE Journal

ID: jaganm Time: 18:15 I Path: S:/J###/Vol00000/170047/Comp/APPFile/SA-J###170047


J186103 DOI: 10.2118/186103-PA Date: 12-January-18 Stage: Page: 21 Total Pages: 16

Degradation through a capillary is often incorrectly assumed to be a function of the shear rate in the capillary. Stavland et al. (2016)
found that degradation scales with shear rate in laminar flow and with a modified shear rate in turbulent flow. Comparing orifices of dif-
ferent diameters, Nguyen and Kausch (1991) found that degradation is a function of the velocity in the capillary or the entrance-pres-
sure drop. An in-depth investigation of the influence of tube and capillary diameter on polymer degradation is presented in Appendix A.
Comparing capillaries of different diameters, we confirm that degradation depends on the velocity at the capillary entrance VC rather
than on the shear rate at the capillary wall.

Kinetics of the Mechanical Degradation. Normalization of Degradation Experiments. Typical results of degradation experiments
with a laboratory blender are presented in Fig. 2. For this experiment, identical volumes of polymer solutions of 3630S at 1,400 ppm in
6 g/L brine were sheared at constant rotational speed in a laboratory blender. Samples were collected at different time intervals for vis-
cosity measurement. The influence of the rotational speed V of the blender was investigated.

10 100
V1 = 2,000 rev/min

Degradation at 73 s–1 (%)


Viscosity at 73 s–1 (cp)

V2 = 2,850 rev/min 80
V3 = 5,750 rev/min
60

40
V3 = 5,750 rev/min
y= 3.504x –0.21 20 V2 = 2,850 rev/min
V1 = 2,000 rev/min
1 0
1 10 100 0 20 40 60 80
Shearing Time (minutes) Shearing Time (minutes)

Fig. 2—Evolution of (left) viscosity at 73 s21 and (right) degradation vs. time of a 1,400-ppm 3630S polymer solution in 6-g/L brine
with a laboratory blender at different rotational speeds (2,000, 2,850, and 5,750 rev/min). Degradation is calculated from viscosity
measurements at 73 s21.

As seen in Fig. 2, the viscosity decreases, according to a power-law function of shearing time. Black curves are power-law functions
that fit the experimental data. The equation listed within the graph corresponds to the fit of the experiment at 5,750 rev/min.
When expressed in terms of degradation, it tends asymptotically toward a final value depending upon rotational speed. In this type
of experiment, the solution experiences successive degradation events at constant strain rate. For this strain rate, there exists a critical
molecular weight Mwc , below which polymer chains are no longer degraded. For each rotational speed, degradation thus evolves toward
a steady-state plateau value, corresponding to Mwc . Before reaching this plateau value, in the transient region, it can be anticipated that
the molecular-weight distribution is different from the steady-state distribution, where no chain longer than Mwc remains.
According to Hunkeler et al. (1996b) and Henaut et al. (2012), degradation results from energy accumulation during the stretching
process, which is correlated with the energy dissipated by the blender. For a given set of physical-chemistry parameters (such as tem-
perature, polymer concentration, viscosity, and salinity), the number of degradation events necessary to reach the steady-state degrada-
tion would thus be a function of the intensity of the extensional-flow field at each degradation event (the energy rate) integrated over
the total residence time in the extensional field. Total residence time can be further decomposed into the product of total shearing time
multiplied by the frequency of degradation events.
An experiment was performed by varying the fluid mass introduced into a blender at constant rotational speed to test the previously
discussed hypothesis. Results are presented in Fig. 3. At early times, degradation is higher with lower fluid mass, as would be expected
because of the increased frequency with which degrading events occur. If the number of degrading events is constant, it should be possi-
ble to normalize the time (t0 ) of the experiments by polymer mass as
t0 ¼ t1  mref =m1 ¼ t2  mref =m2 ¼ t3  mref =m3 ; . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð2Þ

where m1 , m2 , and m3 are the masses of solution of each experiment, mref is a reference mass, and t is the time of each experiment
before normalization. As seen in Fig. 3, after time normalization (m3 ¼ 1200 g, is taken as mref ), curves are quite well-superimposed at
low and high shearing times.

100 100
Degradation at 0.1 s–1 (%)

Degradation at 0.1 s–1 (%)

80 80

60 60

40 m1 = 400 g 40 m1 = 400 g
m2 = 800 g m2 = 800 g
20 m3 = 1,200 g 20
m3 = 1,200 g

0 0
0.01 0.1 1 10 100 0.01 1 100
Shearing Time (minutes) Normalized Time t ′ (minutes)

Fig. 3—Degradation in a laboratory blender vs. time of a 1,100-ppm 3630S polymer solution in 6-g/L brine. Variation of the mass of
solution introduced in the blender. Degradation is calculated from low-shear Newtonian plateau viscosity measurements at 0.1 s21.

February 2018 SPE Journal 21

ID: jaganm Time: 18:15 I Path: S:/J###/Vol00000/170047/Comp/APPFile/SA-J###170047


J186103 DOI: 10.2118/186103-PA Date: 12-January-18 Stage: Page: 22 Total Pages: 16

We hypothesize that a similar normalization dependent on the frequency of degrading events can be performed on the experiments
presented in Fig. 2, for which rotational speed is varied. If we assume that degradation depends on the total energy accumulated and not
on the intensity of the degrading events (the energy-rate function of the rotational speed), shearing at high speed for short time or low
speed for long time should lead to the same degradation product. In other words, high speed would result in a higher number of breaking
events in a given time interval. The energy rate would increase the frequency of the degrading events, but not their nature.
Fig. 4 presents the results for 2,000, 2,850, and 5,750 rev/min, with the latter two normalized as

t0 ¼ t2;850rev=min  a2;850rev=min ¼ t5;750rev=min  a5;750rev=min ; . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð3Þ

where a2;850rev=min and a5;750rev=min are empirical shifting factors adjusted to overlay the curves with the reference experiment at 2,000
rev/min (a2;000rev=min ¼ 1). These shifting factors can be viewed as the ratio of frequencies at which degrading events occur at each rota-
tional speed divided by the frequency of degradation events at 2,000 rev/min. The superposition of the curves at different rotational
speeds is very good. It means that in the present conditions, degradation is additive and is independent from the nature and the history
of the degrading events. It would be very valuable to investigate such degradation kinetics by following the evolution of the molecular-
weight distribution by size-exclusion chromatography. Values found for a2;850rev=min and a5;750rev=min are 4 and 30, respectively, which
indicates that shift factor is not linearly correlated with the rotational speed. In the blender, only one part of the dissipated energy is
involved in the degradation process. If we think in terms of total energy accumulated for the degradation process (stretching and break-
ing of the polymer chains), shift factor should correlate with the energy dissipated in zones where the extensional component of the
flow field dominates.

100 100
Degradation at 73 s–1 (%)

Degradation at 73 s–1 (%)


80 80

60 60

40 40 V3 = 5,750 rev/min
V3 = 5,750 rev/min
V2 = 2,850 rev/min
20 V2 = 2,850 rev/min 20
V1 = 2,000 rev/min
V1 = 2,000 rev/min
0 0
0.1 1 10 100 0.1 10 1,000
Shearing Time (minutes) Normalized Time t ′ (minutes)

Fig. 4—Evolution of degradation vs. time (left) or normalized time (right) of a 1,400-ppm 3630S polymer solution in 6-g/L brine with
a laboratory blender at different rotating speeds (2,000, 2,850, and 5,750 rev/min). Degradations are calculated from viscosity
measurements at 73 s21.

Degradation Kinetics in Various Degrading Geometries. A similar normalization was attempted with experiments performed by
use of different degrading geometries:
• Shearing in a blender at different rotational speeds (3630S at 1,400 ppm in 6-g/L brine, V ¼ 2,000, 2,850 and 5,750 rev/min).
• Multiple passes in an abrupt contraction (Dt ¼ 1 mm/DC ¼ 0.17 mm) at different flow rates (3630S at 800 ppm in 35-g/L brine,
Q ¼ 2, 5, and 20 cm3/min).
• Multiple passes in a porous medium (PEEK frit IDEX OC820, Dcore ¼ 10 mm, Lcore ¼ 1.6 mm, kw ¼ 115 md, / ¼ 0.28) at different
flow rates (3630S at 800 ppm in 6 g/L brine, Q0 ¼ 1, 2, and 5 cm3/min).
The evolution of the degradation vs. the shearing time or the number of passes is presented in Fig. 5 for each geometry.

100 80 60
90
Degradation at 73 s–1 (%)

Degradation at 73 s–1 (%)

Degradation at 73 s–1 (%)

70
50
80 Capillary Q3′ = 20 cm3/min
60
70 Capillary Q2′ = 5 cm3/min
40
50 Capillary Q1′ = 2 cm3/min
60
50 40 30
40 30
Blender V3 = 5,750 rev/min 20
30 PM Q3 = 5 cm3/min
Blender V2 = 2,850 rev/min 20
20 PM Q2 = 2 cm3/min
Blender V1 = 2,000 rev/min 10
10 10 PM Q1 = 1 cm3/min
0 0 0
0 20 40 60 80 0 50 100 150 0 5 10 15 20
Shearing Time (minutes) Pass Number Pass Number

Fig. 5—Degradation (left) in a laboratory blender vs. shearing time for three different rotational speeds, through a capillary
(DC 5 0.17 mm) vs. the number of passes for three different flow rates (center), through a porous medium (PM) vs. the number of
passes for three different flow rates (right).

22 February 2018 SPE Journal

ID: jaganm Time: 18:15 I Path: S:/J###/Vol00000/170047/Comp/APPFile/SA-J###170047


J186103 DOI: 10.2118/186103-PA Date: 12-January-18 Stage: Page: 23 Total Pages: 16

As discussed previously, a normalization factor can be found for each experiment that allows the pass number or shearing time to be
shifted along the abscissa such that a master curve is obtained, as presented in Fig. 6.

100

Degradation at 73 s–1 (%)


80

Blender V3
60 Blender V2
Blender V1
Contraction Q3′
40
Contraction Q2′
Contraction Q1′
20 Porous medium Q3
Porous medium Q2
Porous medium Q1
0
0.01 0.1 1 10 100 1,000 10,000
Normalized Time or Pass Number

Fig. 6—Degradation experiments in a blender at varying rotational speeds V1, V2, and V3, in an abrupt contraction at flow rates
Q10 ; Q20 ; Q30 in a porous medium at flow rates Q1, Q2, and Q3. Shearing time in the blender and pass number are normalized. 3630S
polymer is used in all experiments.

The influence of physical-chemistry parameters (polymer concentration and salinity) in these experiments is discussed in
Appendix B.
By adjusting the shift factors, transient-degradation curves obtained at different salinities or polymer concentrations can be overlaid.
Although the shape of the master curve and the values of the shift factors depend on the shear rate at which degradation is calculated, a
unique curve shape can be obtained for any measurement shear rate. The existence and the shape of the master curves have to be con-
fronted with predictive models of viscosity and mechanical degradation, such as that recently developed by Brakstad and Rosenkilde
(2016), associated with Mw -distribution measurements by size-exclusion chromatography. Moreover, the existence of a master curve
for the degradation of the elastic properties has to be investigated.
Our understanding of the physical significance of the superposition obtained in Fig. 6 is that polymer mechanical degradation in the
transient region is path independent. In other words, chains are broken in the same way, whatever the nature and the history of degrada-
tion. The scenario of bond rupture behind this hypothesis has to be investigated through Mw -distribution measurements by size-exclu-
sion chromatography. We may be in a specific case in our experiments because degradation levels are quite high. The construction of
the master curve in more real situations with lower degrees of degradation has to be investigated. This finding may appear trivial at first
glance, but it has important implications for the prediction of degradation following multiple degrading events in different geometries
(pumps, valves, porous medium). This kind of approach has already been performed by Henaut et al. (2012), who studied mechanical-
degradation kinetics in rheometer and small-scale loop through the loss of drag-reduction efficiency. They have found that degradation
kinetics is correlated with dissipated energy (energy-dissipation rate per unit volume integrated over time). Hence, our shifting factor
can be considered as the energy-dissipation rate contributing to the degradation process (chain stretching and breaking) specific to
each geometry.
Additivity of Mechanical Degradation. The degradation-time-normalization method described previously can be applied to predict
the degradation of a solution that undergoes multiple degradation events of different natures.
As a first example, we may consider a solution that is sheared in a blender for a time interval t1 at a rotational speed V1 and then
sheared again for a time interval t2 at a rotational speed V2 > V1 . Degradation curves are presented in Fig. 7.

100 ηloss1+2 100


ηloss1+2
Degradation at 0.1 s–1 (%)

Degradation at 0.1 s–1 (%)

80 80
ηloss2 ηloss1
60 60

40 V2 after V1 40 V2 after V1

V2 V2
20 20
V1 V1
0 0
0.1 1 10 100 0.01 0.1 1 10 100
Shearing Time (minutes) Normalized Time (minutes)

Fig. 7—(Left) Degradation vs. shearing time in a laboratory blender for two rotational speeds V1 and V2 and for a shearing at V2 af-
ter 10 minutes shearing at V1. (Right) The same curves with the time normalized. 3630S polymer solution at 1,100 ppm in 6-g/L
brine is used. Degradation is calculated from low-shear Newtonian plateau viscosity measurements at 0.1 s21.

For the time interval t1 (10 minutes) at a rotational speed V1 , degradation is 75% (gloss1 on curve V1 ). If solution was sheared at a
rotational speed V2 for a time interval t2 (0.5 minutes), degradation would be 68% (gloss2 on curve V2 ). When the solution is sheared
consecutively 10 minutes at the rotational speed V1 and 0.5 minute at a rotational speed V2 , the measured degradation is 79% (gloss1þ2

February 2018 SPE Journal 23

ID: jaganm Time: 18:15 I Path: S:/J###/Vol00000/170047/Comp/APPFile/SA-J###170047


J186103 DOI: 10.2118/186103-PA Date: 12-January-18 Stage: Page: 24 Total Pages: 16

on curve V2 after V1 ). This result can be easily obtained by the time-renormalization method (Fig. 7). By use of the shift factors
obtained during the normalization procedure, 10 minutes of shearing at a rotational speed V1 is equivalent to 1.1 minutes (teq1 ) shearing
at a rotational speed V2 . When the solution is sheared for a time interval t2 at V2 after 10 minutes of shearing at V1 , degradation is equiv-
alent to a solution that would be sheared for a time interval t2þ1 ¼ t2 þ teq1 . For t2 ¼ 0.5 minutes, we have t2þ1 ¼ 1.6 minutes. Degrada-
tion can then be predicted from curve V2 . An interpolation between the points at 1 and 2 minutes gives 79%, which is equal to the
experimental value. This calculation was performed for time interval t2 at V2 , varying between 0.16 and 45 minutes after 10 minutes
shearing at V1 . Predictions were in total agreement with experimental values. As expected, after normalization, degradation curve V2 af-
ter V1 is completely superimposed with curve V2 (Fig. 7).
The same methodology was successfully applied for the prediction of total degradation after two consecutive degrading events in
geometries of different nature (shearing in a blender followed by injection in a capillary or a porous medium, or injection in a capillary
followed by injection in a porous medium). To summarize our approach, when a solution is degraded through two geometries, 1 and 2,
the total degradation gloss1þ2 is predicted from the knowledge of the degradation level gloss1 after the first geometry and the degradation
curve gloss2 ¼ f(t) of the second geometry.
The fact that degradation gloss1 is transient or steady state at the exit of the first geometry does not affect the degradation in the sec-
ond geometry. This result implies that both samples (transient and steady state) have similar Mw distributions. This interpretation drawn
from indirect observations should be compared with Mw -distribution measurements by size-exclusion chromatography.
This very-rough approach works well if the second geometry is causing additional degradation. In cases where the first degrading
event causes a degradation gloss1 that is higher than the steady-state degradation gloss2 of the second geometry (degradation that would
occur after a high pass number through the second geometry), the preceding procedure cannot be applied because no further degradation
can be achieved with the second geometry. A degradation gloss1 in the first geometry higher than gloss2 in the second geometry would
correspond to the formation of chains with molecular weight lower than Mwc2 . The only possibility is thus to consider that Event 1
occurred after Event 2 to use the degradation curve of the first geometry. A more-accurate methodology would consist in modeling the
evolution of the molecular-weight distribution of the degraded solution.
There is another limitation relative to this empirical approach: Prediction is performed by use of the transient-degradation curve of
the second geometry. In a case where the first geometry leads to different molecular-weight distributions for a given degradation
level, our procedure cannot perform any distinction and we will predict the same additional degradation in the second geometry for
both samples.
As a second example, we may consider the frontal injection of a polymer solution in a porous medium, where we want to determine
the depth of penetration at which degradation becomes steady state (no more evolution). A first indication can be obtained through the
multipass-capillary experiment already presented in Fig. 5. An 800-ppm 3630S polymer solution in 35 g/L brine is injected at constant
flow rate in a capillary DC ¼ 0.17 mm. This capillary is connected to a tube with ID Dt ¼ 1 mm. At each passage, the polymer passes
through a 1 mm to 0.17 mm abrupt contraction. Degradation is measured as a function of the pass number. The curves are similar to
those obtained with a blender (Fig. 5). Degradation increases rapidly on the 10 first passes and then evolves asymptotically toward a
steady-state value. This maximal degradation depends on the flow rate. At low flow rate (<0.5 cm3/h), no degradation is observed (not
shown in Fig. 5). This indicates that for a given polymer solution, there exists a critical strain rate at which degradation occurs. Under
this critical strain rate, degradation does not occur even after infinite solicitations because the coil-stretch transition has not occurred.
At higher flow rate, degradation increases with the number of passes and tends toward a steady-state value after 100 passes. If we esti-
mate the distance between pore throats to be approximately 50 mm and assume degradation in a porous media will also approach steady
state after passing through approximately 100 pore throats, this would represent a distance of approximately 5 mm.
Polymer injection (500 ppm Flopaam 3630S in a 0.4 g/L brine) was performed through a short ceramic core (ceramic frit from
OFITE, Dcore ¼ 25.4 mm, Lcore ¼ 6.3 mm, kw ¼ 8,740 md, / ¼ 0.43), and effluent viscosity was measured for different flow rates. Shear
rate in the porous medium was evaluated with the capillary-bundle model. As seen in Fig. 8, degradation of the initial solution (blue
curve) increases with the shear rate (Jouenne et al. 2016).

40

35 Initial
Degradation at 7 s–1 (%)

ηloss = 2%
30
ηloss = 15%
25 ηloss = 25%
20

15

10

0
10 100 1,000 10,000 100,000
Shearing Rate (s–1)

Fig. 8—Degradation of a polymer solution through an 8.7D ceramic frit (Dcore 525:4 m, Lcore 5 6.3 mm). The continuous blue line is
the degradation curve of the fresh polymer solution. Symbols represent the additional degradation of solutions reinjected on the
frit. Shear rates at which samples were collected for reinjection are indicated by the empty circles.

At different flow rates (indicated by the empty circles in Fig. 8), a large volume of solution was collected and reinjected. Additional
degradation was measured as a function of the flow rate during this second injection. Degradation curves of these predegraded samples
are plotted in Fig. 8. For all the predegraded solutions injected a second time in the porous medium, the shear rate at which starts an
additional degradation corresponds approximately to the shear rate at which solution was collected during the first injection, as reported
by Seright (1983). This suggests that degradation has reached a plateau or steady-state degradation during the first injection.

24 February 2018 SPE Journal

ID: jaganm Time: 18:15 I Path: S:/J###/Vol00000/170047/Comp/APPFile/SA-J###170047


J186103 DOI: 10.2118/186103-PA Date: 12-January-18 Stage: Page: 25 Total Pages: 16

It can be concluded that, in frontal injection, passing through 6 mm of porous medium is sufficient to reach the steady-state degrada-
tion. This result likely applies in the case of radial flow because shear rate varies little over the first 6 mm of the radial profile. However,
this result highlights the important role near-wellbore damage can have on polymer degradation.
Prediction of the Degradation of a Polymer Solution Injected in the Field. We may now consider the injection of a polymer solu-
tion, the anticipated level of degradation of which after passing through the near-wellbore region is gloss;wellbore (degradation is steady
state). If the solution is degraded in surface facilities at the level gloss;surface , total degradation gloss;tot can be predicted according to the
four possible situations sketched in Fig. 9.
• Scenario A: Surface degradation is steady state and lower than near-wellbore steady-state degradation. Total degradation will be
equal to the near-wellbore degradation.
• Scenario B: Surface degradation is transient with a steady-state plateau lower than near-wellbore steady-state degradation. Total
degradation will be equal to the near-wellbore degradation.
• Scenario C: Surface degradation is transient with a steady-state plateau higher than near-wellbore steady-state degradation. Total
degradation will be higher than surface degradation. For example, this case would correspond to a passage in a constriction that
would lead to higher degradation during a second passage. As previously explained, for the prediction of the total degradation, we
use the degradation curve of the first geometry and consider a first degrading event equal to gloss;wellbore .
• Scenario D: Surface degradation is steady state and higher than near-wellbore steady-state degradation. Total degradation will be
equal to the surface degradation.

ηloss,tot ηloss,tot
Degradation

Degradation
ηloss,wellbore ηloss,wellbore

ηloss,surface
ηloss,surface
A B

Number of Degrading Events Number of Degrading Events


(a) (b)
ηloss,tot
ηloss,tot
Degradation

Degradation

ηloss,surface ηloss,surface

ηloss,wellbore ηloss,wellbore
C D

Number of Degrading Events Number of Degrading Events

(c) (d)

Fig. 9—Schematic of the four possible situations encountered when injecting a polymer solution in the field. gloss;wellbore is the
anticipated level of degradation at the near-wellbore region. gloss;surface is the level of degradation in surface facilities (constriction,
pumps, valves). gloss;tot is the predicted level of total degradation.

Predegradation of a Polymer Solution in the Laboratory. When a solution is degraded consecutively in two geometries, it was
shown that degradation in the second geometry can be expressed as a function of the degradation level in the first geometry. Hence, if
we have to prepare a predegraded solution in the laboratory that is representative of surface-degradation events, predegradation can be
performed with any kind of degrading geometry (capillary, porous medium, blender). Moreover, the transient or steady-state character
of the degradation is at first order not important (nevertheless, the existence of the degradation master curve for low levels of degrada-
tion has to be investigated).
Degradation as a Probe of Molecular-Weight Distribution. Our results imply that a polymer degraded to a certain viscosity has the
same molecular-weight distribution whatever the method used. Such a scenario is consistent with theories suggesting the preferential
breaking of high Mw chains even in semidilute regime for which scission was found to be more random (Müller 1992) compared with
the dilute regime. Statistical modeling of the Mw -distribution evolution following a random scission, near the midpoint, of polymer
chains exceeding a critical molecular weight leads to a decrease of the polydispersity index to values as low as 1.12. Multiple passes
through the transient extensional field, such as capillaries, were proposed as a route to generate monodisperse polymers (Buchholz
et al. 2004).
A mechanical-degradation experiment was performed in a blender to qualitatively measure polydispersity of industrial samples.
HPAM polymer solutions of each grade of the Flopaam series were prepared in 6 g/L brine at a concentration of 1,100 ppm. Viscosity
was measured as a function of the shearing time at fixed rotational speed. For each grade, evolution of the viscosity with shearing time
is plotted in Fig. 10. The degradation curve of the highest Mw polymer (3630S, Mw of approximately 18.5 MDa) is taken as the refer-
ence. For the other grades with lower Mw , shearing time is shifted so that their initial viscosity corresponds to the viscosity of degraded

February 2018 SPE Journal 25

ID: jaganm Time: 18:15 I Path: S:/J###/Vol00000/170047/Comp/APPFile/SA-J###170047


J186103 DOI: 10.2118/186103-PA Date: 12-January-18 Stage: Page: 26 Total Pages: 16

3630S. Whatever the grade (3530S, Mw ¼ 14.5 MDa; 3330S, Mw ¼ 9.5 MDa; 3230S, Mw ¼ 6.5 MDa; 3130S, Mw ¼ 3.5 MDa), viscosity
after an equivalent shearing time is lower than viscosity of a degraded 3630S. In the same way, viscosity of degraded 3530S remains
higher than viscosity of the lower Mw grades. These results demonstrate that at similar concentration, high Mw polymers will remain
more viscous than lower Mw polymers. This can be explained by the lower polydispersity of degraded high Mw polymers compared
with commercial low Mw polymers. Müller (1992) found that monodisperse-polymer solutions are more resistant to shear than polydis-
perse-polymer solutions. In our experiments, degradation is a way to narrow the polydispersity of the commercial polymer. Mechanical
degradation preferentially breaks the high Mw tail of the Mw distribution. Hence, comparing 3630S and 3530S, the same portion of high
Mw chains is broken but the fraction of low Mw chains is larger for 3530S.

25
Mw = 18.5 MDa

20 Mw = 14.5 MDa

Viscosity at 0.1 s–1 (cp)


Mw = 9.5 MDa

15 Mw = 6.5 MDa

Mw = 3.5 MDa

10

0
0.1 1 10 100 1,000

Time (minutes)

Fig. 10—Degradation vs. shearing times in a blender of industrial HPAM grades of different average molecular weights (3630S,
Mw 5 18.5 MDa; 3530S, Mw 5 14.5 MDa; 3330S, Mw 5 9.5 MDa; 3230S, Mw 5 6.5 MDa; 3130S, Mw 5 3.5 MDa). Polymer solutions are
1,100 ppm in 6-g/L brine. Degradation curve of the highest Mw polymer (3630S, Mw 5 approximately 18.5 MDa) is taken as the refer-
ence. For the other grades with lower Mw, shearing time is shifted so that their initial viscosity corresponds to the viscosity of
degraded 3630S.

One important extension of this finding is the beneficial effect of using a high-Mw polymer degraded to a steady-state level by multi-
ple degradation events, rather than a medium-Mw polymer as received, to eliminate the high-molecular-weight tail that contributes pri-
marily to elasticity, hence decoupling viscosity from injectivity characteristics. This result approach is particularly attractive for use in
low-permeability carbonate reservoirs for maximizing viscosifying power and injectivity (Levitt et al. 2012).
In the absence of analytical means such as size-exclusion chromatography, the preceding examples illustrate how mechanical-degra-
dation experiments may be used to qualitatively probe molecular-weight distribution. Indeed, molecular-weight distributions were
derived from extensional and degradation measurements in cross-slot or opposed jet devices for dilute (Keller and Odell 1985; Odell
and Keller 1986; Odell et al. 1990) or semidilute solutions (Müller et al. 1992).

Conclusions
Kinetic aspects of mechanical polymer degradation have been investigated by use of various geometries, including abrupt contractions,
blenders, and porous media.
It is demonstrated that pressure drop cannot be used as the controlling degrading parameter when dealing with degradation in capil-
laries of different lengths. Experiments in capillaries of different diameters revealed that degradation depends on the velocity rather
than the shear rate in the capillary.
Degradation achieved with the three geometries tested (multiple passages through an abrupt contraction or a porous medium at dif-
ferent flow rates, shearing in a blender at different rotational speeds) can be superposed by applying empirical shift factors to the shear-
ing time in a blender or the pass number in contractions or porous media. This superposition suggests that polymer mechanical
degradation is path independent. Hence, predegradation in the laboratory representating surface degradation experienced during a field
injection can be performed with any kind of degrading geometry. This result can be used to predict total degradation following several
degradation events by use of our empirical normalization procedure.
In a field application, degraded-polymer solutions may sometimes retain their original viscosity better and exhibit better injectivity
characteristics. At isoconcentration, degraded high-Mw polymers are found to be more shear tolerant than low-Mw commercial poly-
mers. By preferentially breaking the high-Mw tail of the Mw distribution, mechanical degradation narrows the polydispersity of commer-
cial polymers. This fact can also be used to probe the Mw distribution of a polymer solution. These results remain to be validated by
direct measurements of molecular-weight distribution by size-exclusion chromatography. Moreover, additional experiments at lower
levels of degradation associated with the measurement of the elastic properties are necessary.
Finally, by use of several degradation experiments, it was shown that, for the conditions examined, passing through 6 mm of porous
medium is sufficient to reach a steady state of degradation. This implies that mechanical degradation can be easily predicted in radial
flow. In a case where degradation occurs in surface facilities, total degradation can be predicted by the normalization method.

Nomenclature
c/c* ¼ entanglement density
DC ¼ capillary internal diameter, mm, m
Dcore ¼ core diameter, mm, m
Dfilter ¼ filter diameter for the filtration of the polymer solution at constant flow rate, mm, m

26 February 2018 SPE Journal

ID: jaganm Time: 18:15 I Path: S:/J###/Vol00000/170047/Comp/APPFile/SA-J###170047


J186103 DOI: 10.2118/186103-PA Date: 12-January-18 Stage: Page: 27 Total Pages: 16

Dt ¼ inlet- and outlet-tube inside diameter in the degradation experiments through a capillary, mm, m
kw ¼ permeability, md, mm2
m1 ; m 2 ; m 3 ¼ mass of liquid in the blender, g
Mw ¼ average molecular weight, g/mol
Mwc ¼ for a given strain rate e_ , critical molecular weight below which polymer chains are no longer degraded
N ¼ number of beads when the polymer chain is considered as an extended string of beads
Q ¼ volumetric-flow rate in degradation experiments through capillary or porous medium, cm3/min, m3/s
Q0 ¼ volumetric-flow rate in degradation experiments through capillary or porous medium, cm3/min, m3/s
SC ¼ cross-sectional area of a capillary, m2
St ¼ cross-sectional area of a capillary, m2
t ¼ shearing time for the degradation experiments in a blender, minutes, seconds
teq1 ¼ equivalent time of shearing in a blender at a velocity V2 for a solution sheared for a time t1 at a velocity V1
t1 ¼ shearing time for the degradation experiment in a blender at rotational speed V1 , minutes, seconds
t2þ1 ¼ equivalent time of shearing in a blender at a velocity V2 for a solution sheared for a time t1 at a velocity V1 and then
sheared for a time t2 at a velocity V2
t0 ¼ normalized time or normalized pass number
T ¼ maximum tension exerted at the center of a stretched polymer chain in extensional flow
VC ¼ velocity at capillary entrance, m/s
Vt ¼ velocity in the inlet and outlet tubes for the degradation experiments in a capillary, m/s
V1 ¼ rotational speed of the blender, rev/min
ai ¼ empirical shifting factor adjusted to normalized the pass number or the shearing time and to overlay the degradation
curves obtained with different degrading geometries; a ¼ 1 for the reference experiment
c_ c¼ shear rate at the capillary wall
DP ¼ pressure drop through a capillary, bar
¼
e_ strain rate, s1
e_ C ¼ critical strain rate at which the coil-stretch transition occurs, s1
e_ R ¼ critical strain rate for chain scission
g ¼ viscosity, cp, Pas
gdeg ¼ viscosity of the degraded solution, cp, Pas
gH2 O ¼ water viscosity, cp, Pas
gloss ¼ degradation of the polymer-solution viscosity, value is specific to the shear rate at which viscosity is measured, %
gloss1 ¼ degradation through the first geometry, %
gloss1þ2 ¼ total degradation when a solution is consecutively degraded in two geometries or degrading conditions 1 and 2, %
gloss;surface ¼ degradation in surface facilities during a field injection, %
gloss;tot ¼ total degradation on a field injection following surface and near-wellbore degradation
gloss;wellbore ¼ degradation after passing through the near-wellbore region in an injection well, %
g0 ¼ viscosity of the nondegraded polymer solution, cp, Pas
/¼ porosity

Acknowledgments
We gratefully acknowledge Cyrille Hourcq, Guillaume Heurteux, and Michèle Joly from Total E&P and Sarah Bel Hadj, Lionel Longe,
and Faustin Ndong for performing laboratory experiments. We thank Guy Chauveteau for sharing his knowledge of extensional flows
and mechanical degradation. We thank Total E&P management for permission to publish this work.

References
Al Hashmi, A. R., Al Maamari, R. S., Al Shabibi, I. S. et al. 2013. Rheology and Mechanical Degradation of High-Molecular-Weight Partially Hydro-
lyzed Polyacrylamide During Flow Through Capillaries. J. Pet. Sci. Eng. 105 (May): 100–106. https://doi.org/10.1016/j.petrol.2013.03.021.
Brakstad, K. and Rosenkilde, C. 2016. Modelling Viscosity and Mechanical Degradation of Polyacrylamide Solutions in Porous Media. Presented at the
SPE Improved Oil Recovery Conference, Tulsa, 11–13 April. SPE-179593-MS. https://doi.org/10.2118/179593-MS.
Buchholz, B. A., Zahn, J. M., Kenward, M. et al. 2004. Flow-Induced Chain Scission as a Physical Route to Narrowly Distributed, High Molar Mass
Polymers. Polymer 45 (4): 1223–1234. https://doi.org/10.1016/j.polymer.2003.11.051.
Chauveteau, G. 1981. Molecular Interpretation of Several Different Properties of Flow of Coiled Polymer Solutions Through Porous Media in Oil Recov-
ery Conditions. Presented at the SPE Annual Fall Technical Conference and Exhibition, San Antonio, Texas, 5–7 October. SPE-10060-MS. https://
doi.org/10.2118/10060-MS.
Chauveteau, G. and Kohler, N. 1984. Influence of Microgels in Polysaccharide Solutions on Their Flow Behavior Through Porous Media. SPE J. 24 (3):
361–368. SPE-9295-PA. https://doi.org/10.2118/9295-PA.
Chow, A., Keller, A., Müller, A. J. et al. 1988. Entanglements in Polymer Solutions Under Elongational Flow: A Combined Study of Chain Stretching,
Flow Velocimetry and Elongational Viscosity. Macromolecules 21 (1): 250–256. https://doi.org/10.1021/ma00179a048.
Culter, J. D., Zakin, J. L., and Patterson, G. K. 1975. Mechanical Degradation of Dilute Solutions of High Polymers in Capillary Tube Flow. J. Appl.
Polym. Sci. 19 (12): 3235–3240. https://doi.org/10.1002/app.1975.070191210.
De Gennes, P. G. 1974. Coil-Stretch Transition of Dilute Flexible Polymers Under Ultrahigh Velocity Gradients. J. Chem. Phys. 60 (12): 5030–5042.
https://doi.org/10.1063/1.1681018.
Dupas, A., Henaut, I., Argillier, J.-F. et al. 2012. Mechanical Degradation Onset of Polyethylene Oxide Used as a Hydrosoluble Model Polymer for
Enhanced Oil Recovery. Oil Gas Sci. Tech. 67 (6): 931–940. https://doi.org/10.2516/ogst/2012028.
Ghoniem, S., Chauveteau, G., Moan, M. et al. 1981. Mechanical Degradation of Semi-Dilute Polymer Solutions in Laminar Flows. Can. J. Chem. Eng.
59 (4): 450–454. https://doi.org/10.1002/cjce.5450590406.
Ghoniem, S. A.-A. 1985. Extensional Flow of Polymer Solutions Through Porous Media. Rheol. Acta 24 (6): 588–595. https://doi.org/10.1007/
BF01332592.
Henaut, I., Glenat, P., Cassar, C. et al. 2012. Mechanical Degradation Kinetics of Polymeric DRAs. Presented at the 8th North American Conference on
Multiphase Technology, Banff, Canada, 20–22 June. BHR-2012-A004.

February 2018 SPE Journal 27

ID: jaganm Time: 18:15 I Path: S:/J###/Vol00000/170047/Comp/APPFile/SA-J###170047


J186103 DOI: 10.2118/186103-PA Date: 12-January-18 Stage: Page: 28 Total Pages: 16

Hunkeler, D., Nguyen, T. Q., and Kausch, H. H. 1996a. Polymer Solutions Under Elongational Flow: 1. Birefringence Characterization of Transient and
Stagnation Point Elongational Flows. Polymer 37 (19): 4257–4269. https://doi.org/10.1016/0032-3861(96)00290-X.
Hunkeler, D., Nguyen, T. Q., and Kausch, H. H. 1996b. Polymer Solutions Under Elongational Flow: 2. An Evaluation of Models of Polymer Dynamics
for Transient and Stagnation Point Flows. Polymer 37 (19): 4271–4281. https://doi.org/10.1016/0032-3861(96)00187-5.
Jouenne, S., Klimenko, A., and Levitt, D. 2016. Tradeoffs Between Emulsion and Powder Polymers for EOR. Presented at the SPE Improved Oil Recov-
ery Conference, Tulsa, 11–13 April. SPE-179631-MS. https://doi.org/10.2118/179631-MS.
Keller, A. and Odell, J. A. 1985. The Extensibility of Macromolecules in Solution; a New Focus for Macromolecular Science. Colloid Polym. Sci. 263
(3): 181–201. https://doi.org/10.1007/BF01415506.
Levitt, D., Klimenko, A., Jouenne, S. et al. 2012. Design Challenges of Chemical EOR in High-Temperature High Salinity Carbonates. Presented at the
SPE Abu Dhabi International Petroleum Conference and Exhibition, Abu Dhabi, 11–14 November. SPE-161633-MS. https://doi.org/10.2118/
161633-MS.
Maerker, J. M. 1975. Shear Degradation of Partially Hydrolyzed Polyacrylamide Solutions. SPE J. 15 (4): 311–322. SPE-5101-PA. https://doi.org/
10.2118/5101-PA.
Maerker, J. M. 1976. Mechanical Degradation of Partially Hydrolyzed Polyacrylamide Solutions in Unconsolidated Porous Media. SPE J. 16 (4):
172–174. SPE-5672-PA. https://doi.org/10.2118/5672-PA.
Magueur, A., Moan, G., M., and Chauveteau, G. 1985. Effect of Successive Contractions and Expansions on the Apparent Viscosity of Dilute Polymer
Solutions. Chem. Eng. Comm. 36 (1–6): 351–366. https://doi.org/10.1080/00986448508911265.
Martin, F. D. 1986. Mechanical Degradation of Polyacrylamide Solutions in Core Plugs From Several Carbonate Reservoirs. SPE Form Eval 1 (2):
139–150. SPE-12651-PA. https://doi.org/10.2118/12651-PA.
Müller, A. J., Odell, J. A., and Carrington, S. 1992. Degradation of Semidilute Polymer Solutions in Elongational Flows. Polymer 33 (12): 2598–2604.
https://doi.org/10.1016/0032-3861(92)91143-P.
Nghe, P., Tabeling, P., and Ajdari, A. 2010. Flow-Induced Polymer Degradation Probed by a High Throughput Microfluidic Set-up. J. Non-Newton.
Fluid 165 (7–8): 313–322. https://doi.org/10.1016/j.jnnfm.2010.01.006.
Nguyen, T. Q. and Kausch, H.-H. 1986. Mechanochemical Degradation of Polymer Solution in Capillary Flow: Laminar and Turbulent Regime. Chimia
40 (4): 129–135.
Nguyen, T. Q. and Kausch, H.-H. 1988. Chain Scission in Transient Extensional Flow Kinetics and Molecular Weight Dependence. J. Non-Newton.
Fluid 30 (2–3): 125–140. https://doi.org/10.1016/0377-0257(88)85020-1.
Nguyen, T. Q. and Kausch, H.-H. 1991. Influence of Nozzle Geometry on Polystyrene Degradation in Convergent Flow. Colloid Polym. Sci. 269 (11):
1099–1110. https://doi.org/10.1007/BF00654117.
Noik, C. H., Delaplace, P. H., and Muller, G. 1995. Physico-Chemical Characteristics of Polyacrylamide Solutions after Mechanical Degradation through
a Porous Medium. Presented at the SPE International Symposium on Oilfield Chemistry, San Antonio, Texas, 14–17 February. SPE-28954-MS.
https://doi.org/10.2118/28954-MS.
Odell, J. A. and Keller, A. 1986. Flow-Induced Chain Fracture of Isolated Linear Macromolecules in Solution. J. Polym. Sci. B 24 (9): 1889–1916.
https://doi.org/10.1002/polb.1986.090240901.
Odell, J. A., Muller, A. J., Narh, K. A. et al. 1990. Degradation of Polymer Solutions in Extensional Flows. Macromolecules 23 (12): 3092–3103. https://
doi.org/10.1021/ma00214a011.
Seright, R. S. 1983. The Effects of Mechanical Degradation and Viscoelastic Behavior on Injectivity of Polyacrylamide Solutions. SPE J. 23 (3):
475–485. SPE-9297-PA. https://doi.org/10.2118/9297-PA.
Seright, R. S., Seheult, J. M., and Talashek, T. 2009. Injectivity Characteristics of EOR Polymers. SPE Res Eval & Eng 12 (5): 783–792. SPE-115142-
PA. https://doi.org/10.2118/115142-PA.
Southwick, J. G. and Manke, C. W. 1988. Molecular Degradation Injectivity and Elastic Properties of Polymer Solutions. SPE Res Eng 3 (4):
1193–1201. SPE-15652-PA. https://doi.org/10.2118/15652-PA.
Stavland, A., Asen, S. M., Mebratu, A. et al. 2016. Impact of Choke Valves on the IOR Polymer Flooding: Lessons Learned from Large Scale Tests. In
Proc., IOR Norway 2016, Stavanger, Norway, 25–26 April, 18–22. IRIS.
Vanapalli, S. A., Ceccio, S. L., and Solomon, M. J. 2006. Universal Scaling for Polymer Chain Scission in Turbulence. Proc. Natl. Acad. Sci. USA 103
(45): 16660–16665. https://doi.org/10.1073/pnas.0607933103.
Zaitoun, A., Makakou, P., Blin, B. et al. 2011. Shear Stability of EOR Polymers. Presented at the SPE International Symposium on Oilfield Chemistry,
The Woodlands, Texas, 11–13 April. SPE-141113-MS. https://doi.org/10.2118/141113-MS.

Appendix A—Degradation Through a Capillary (An Abrupt Contraction)


In these series of experiments, an 800-ppm polymer solution of 3630S in a 35 g/L brine was injected through a stainless-steel capillary
of ID DC ¼ 0.18, 0.25, or 0.5 mm connected at both ends to identical PEEK tubes of IDs Dt ¼ 2, 1.6, or 1 mm. The length of the capil-
lary was constant and equal to 30 cm. The capillary and the tubes were connected with a PEEK union fitting manually drilled to prevent
any diameter variation at the interior of the fitting and to ensure that the contraction was abrupt. For PEEK tubes with IDs Dt ¼ 2 and
1.6 mm, the capillary was inserted inside the tube, as sketched in Fig. A-1. For the PEEK tube with ID Dt ¼ 1 mm, the tubes and the
capillary were put face to face inside the union fitting, as in Fig. A-1.

ΔP ΔP
1
Tube outer diameter = /16 in. Tube outer diameter = 1/16 in.
Inner diameter = 1 mm Inner diameter = 1 mm
Tube outer diameter = 1/16 in.
Tube outer diameter = 1/16 in.
Inner diameter = 0.18, 0.25 or 0.5 mm
Inner diameter = 0.18, 0.25 or 0.5 mm

Tube outer diameter = 1/8 in.


Inner diameter = 2 or 1.6 mm Union fitting –1/16 in.–1/8 in. Tube outer diameter = 1/8 in. Union fitting –1/16 in.–1/8 in.
Te fitting 1/8 in.–1/8 in. Inner diameter =1 mm Te fitting 1/8 in.–1/8 in.

Fig. A-1—Description of the experimental setup. The small-diameter capillary is either inserted into the higher-diameter inlet and
outlet tubes (left) or put face to face (right).

28 February 2018 SPE Journal

ID: jaganm Time: 18:15 I Path: S:/J###/Vol00000/170047/Comp/APPFile/SA-J###170047


J186103 DOI: 10.2118/186103-PA Date: 12-January-18 Stage: Page: 29 Total Pages: 16

Polymer solution was injected at controlled flow rate with an Isco 1000D pump thermostated at 30 C. The tubes and the capillary
were also submerged in a thermostated water bath at 30 C. Flow rate was increased step by step until a pressure drop of 20 bar was
reached. At each step, the pressure drop through the capillary/inlet contraction/outlet expansion was measured. Effluent fluid was col-
lected and degradation was evaluated by the viscosity loss at 73 s1 with Eq. 1.

Influence of the Inlet-Tube Diameter at Constant Capillary Diameter. The pressure drop is the sum of the entrance-pressure drop,
the exit-pressure drop, and the frictional-pressure drop with the capillary wall. As seen in Fig. A-2, the diameter of the inlet tube has no
effect on the pressure drop over the whole length of the capillary, whatever the flow rate. This suggests that the entrance-pressure drop
is negligible compared with the shear-pressure drop all along the capillary, or that the diameter of the inlet tube has no influence on the
entrance- and exit-pressure drops.

250 60 60

Degradation at 73 s–1 (%)

Degradation at 73 s–1 (%)


50 50
Flow Rate (cm3/min)

200

40 40
150
30 30
100
20 20
Dt = 1 mm; Dc = 0.5 mm; L = 30 cm
Dt = 1 mm; Dc = 0.5 mm; L = 30 cm Dt = 1 mm; Dc = 0.5 mm; L = 30 cm
50 Dt = 1.6 mm; Dc = 0.5 mm; L = 30 cm
Dt = 1.6 mm; Dc = 0.5 mm; L = 30 cm 10 Dt = 1.6 mm; Dc = 0.5 mm; L = 30 cm 10
Dt = 2 mm; Dc = 0.5 mm; L = 30 cm
Dt = 2 mm; Dc = 0.5 mm; L = 30 cm Dt = 2 mm; Dc = 0.5 mm; L = 30 cm
0 0 0
0 5 10 15 20 25 0 5 10 15 20 25 0 50 100 150 200 250
ΔP (bar) ΔP (bar) Flow Rate (cm3/min)

Fig. A-2—Degradation of an 800-ppm 3630S polymer solution in a 35 g/L brine through a capillary (DC 5 0.5 mm) connected to inlet
and outlet tubes (Dt 5 2, 1.6, or 1 mm). (Left) Flow rate vs. DP through the capillary. (Center) Degradation vs. pressure drop. (Right)
Flow rate through the capillary. Degradation is calculated from viscosity measurements at 73 s21.

The dependence of degradation on flow rate and DP is plotted for all the tests in Fig. A-2. All the curves are superposed (the higher
degradation measured with Dt ¼ 2 mm is within the uncertainty of the degradation measurement). These results are similar to those
obtained with a capillary DC ¼ 0.25 mm (Fig. A-3) or DC ¼ 0.18 mm (Fig. A-4).

60 60
25
Dt = 1 mm; Dc = 0.25 mm; L = 30 cm
Degradation at 73 s–1 (%)

Degradation at 73 s–1 (%) Dt = 1 mm; Dc = 0.25 mm; L = 30 cm


50 50
Flow Rate (cm3/min)

Dt = 1.6 mm; Dc = 0.25 mm; L = 30 cm


20 Dt = 1.6 mm; Dc = 0.25 mm; L = 30 cm
Dt = 2 mm; Dc = 0.25 mm; L = 30 cm
40 40 Dt = 2 mm; Dc = 0.25 mm; L = 30 cm

15
30 30

10
20 20

5 Dt = 1 mm; Dc = 0.25 mm; L = 30 cm 10 10


Dt = 1.6 mm; Dc = 0.25 mm; L = 30 cm
Dt = 2 mm; Dc = 0.25 mm; L = 30 cm
0 0 0
0 5 10 15 20 25 0 10 20 30 40 0 10 20 30 40 50

ΔP (bar) ΔP (bar) Flow Rate (cm3/min)

Fig. A-3—Degradation of a 3630S polymer solution at 800 ppm in 35 g/L brine through a capillary (DC 5 0.25 mm) connected to inlet
and outlet tubes (Dt 5 2, 1.6, or 1 mm).

6 60
60
Degradation at 73 s–1 (%)

Degradation at 73 s–1 (%)

Dt = 1 mm; Dc = 0.18 mm; L = 30 cm


5 50 Dt = 1 mm; Dc = 0.18 mm; L = 30 cm
50
Flow Rate (cm3/min)

Dt = 1.6 mm; Dc = 0.18 mm; L = 30 cm


Dt = 1.6 mm; Dc = 0.18 mm; L = 30 cm
Dt = 2 mm; Dc = 0.18 mm; L = 30 cm
4 40 40 Dt = 2 mm; Dc = 0.18 mm; L = 30 cm

3 30 30

2 20 20
Dt = 1 mm; Dc = 0.18 mm; L = 30 cm
1 Dt = 1.6 mm; Dc = 0.18 mm; L = 30 cm 10 10
Dt = 2 mm; Dc = 0.18 mm; L = 30 cm

0 0 0
0 5 10 15 20 25 0 20 40 60 80 0 4 8 12 16 20 24

ΔP (bar) ΔP (bar) Flow Rate (cm3/min)

Fig. A-4—Degradation of a 3630S polymer solution at 800 ppm in a 35 g/L brine through a capillary (DC 5 0.18 mm) connected to
inlet and outlet tubes (Dt 5 2, 1.6, or 1 mm).

February 2018 SPE Journal 29

ID: jaganm Time: 18:15 I Path: S:/J###/Vol00000/170047/Comp/APPFile/SA-J###170047


J186103 DOI: 10.2118/186103-PA Date: 12-January-18 Stage: Page: 30 Total Pages: 16

Influence of Capillary Diameter at Constant Inlet- and Outlet-Tube Diameter. As seen in Fig. A-5, when the capillary diameter is
varied (for inlet and outlet tubes Dt ¼ 2 mm), degradation is no longer a function of the flow rate or pressure drop. A decrease of the
capillary diameter dramatically increases pressure drop (which varies as 1=D4 in laminar flow) at fixed flow rate. The counterintuitive
result is that degradation is minimal with the smallest-diameter capillary. Reducing the capillary diameter is thus a good way to create
frictional pressure drop without degradation. The dependence of degradation on shear rate at the capillary wall (_c c ) and velocity at the
capillary entrance (VC ) is plotted for all the tests in Fig. A-6 (three capillaries with diameters DC ¼ 0.5, 0.25, and 0.18 mm connected to
tubes of diameters Dt ¼ 2, 1.6, and 1 mm). Across the different combinations tested (Dt =DC ), degradation is comparable at similar ve-
locity VC , rather than similar shear rate c_ C , in accordance with the observations of Nguyen and Kausch (1991).
Velocity at the capillary entrance (VC ) is calculated by use of

VC ¼ Q=SC ¼ Q=ðpD2C Þ; . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ðA-1Þ

where Q is the volumetric-flow rate, SC is the cross-sectional area of the capillary, and DC is the internal-capillary diameter.

Dt = 2 mm; Dc = 0.18 mm; L = 30 cm


60 60
Degradation at 73 s–1 (%)

Degradation at 73 s–1 (%)


Dt = 2 mm; Dc = 0.25 mm; L = 30 cm

50 Dt = 2 mm; Dc = 0.5 mm; L = 30 cm


50

40 40

30 30

20 20
Dt = 2 mm; Dc = 0.18 mm; L = 30 cm
10 Dt = 2 mm; Dc = 0.25 mm; L = 30 cm 10
Dt = 2 mm; Dc = 0.5 mm; L = 30 cm
0 0
0 100 200 300 0 5 10 15 20
Flow Rate (cm3/min) ΔP (bar)

Dt = 2 mm; Dc = 0.18 mm; L = 30 cm 60


60
Dt = 2 mm; Dc = 0.25 mm; L = 30 cm
Degradation at 73 s–1 (%)

Degradation at 73 s–1 (%)

50 Dt = 2 mm; Dc = 0.5 mm; L = 30 cm 50

40 40

30 30

20 20
Dt = 2 mm; Dc = 0.18 mm; L = 30 cm

10 Dt = 2 mm; Dc = 0.25 mm; L = 30 cm


10
Dt = 2 mm; Dc = 0.5 mm; L = 30 cm
0 0
0 100,000 200,000 300,000 0 5 10 15 20

Shear Rate at the Capillary Wall (s–1) Velocity at the Capillary Entrance (m/s)

Fig. A-5—Degradation of a 3630S polymer solution at 800 ppm in a 35 g/L brine through capillaries of different IDs (DC 5 0.5, 0.25,
or 0.18 mm) connected to inlet and outlet tubes (Dt 5 2 mm).

60 60
Degradation at 73 s–1 (%)

Degradation at 73 s–1 (%)

50 50

40 40
Dtube = 2 mm; Dcap = 0.5 mm; L = 30 cm
30 30 Dtube = 1.6 mm; Dcap = 0.5 mm; L = 30 cm
Dtube = 1 mm; Dcap = 0.5 mm; L = 30 cm
Dtube = 2 mm; Dcap = 0.25 mm; L = 30 cm
20 20 Dtube = 1.6 mm; Dcap = 0.25 mm; L = 30 cm
Dtube = 1 mm; Dcap = 0.25 mm; L = 30 cm
Dtube = 2 mm; Dcap = 0.18 mm; L = 30 cm
10 10 Dtube = 1.6 mm; Dcap = 0.18 mm; L = 30 cm
Dtube = 1 mm; Dcap = 0.18 mm; L = 30 cm
0 0
0 100,000 200,000 300,000 0 5 10 15 20
Shear Rate at the Capillary Wall (s–1)
Velocity at the Capillary Entrance (m/s)

Fig. A-6—Degradation of a 3630S polymer solution at 800 ppm in a 35 g/L brine through capillaries of different IDs (DC 5 0.5, 0.25,
or 0.18 mm) connected to inlet and outlet tubes (Dt 5 2, 1.6, and 1 mm).

30 February 2018 SPE Journal

ID: jaganm Time: 18:15 I Path: S:/J###/Vol00000/170047/Comp/APPFile/SA-J###170047


J186103 DOI: 10.2118/186103-PA Date: 12-January-18 Stage: Page: 31 Total Pages: 16

Shear rate at the capillary wall (_c c ) is calculated by use of

c_ C ¼ 8V=DC : . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ðA-2Þ

Degradation Through a Capillary. In the range of contraction ratio investigated (inlet-tube diameter Dt ¼ 2, 1.6, and 1 mm; capillary
diameter DC ¼ 0.5, 0.25, and 0.18 mm), the following conclusions can be drawn:
• At constant capillary diameter, the inlet-tube diameter has negligible effect on the polymer degradation.
• When comparing capillaries of different diameters, the degradation depends on velocity at the capillary entrance rather than shear
rate at the capillary wall.

Appendix B—Parameters Affecting the Transient-Degradation Master Curve


Influence of the Shear Rate at Which Viscosity Is Measured. We may consider an experiment in which the solution is sheared in a
blender for a time interval t1 at a rotational speed V1 and then sheared again for a time interval t2 at a rotational speed V2 (Fig. 7). By
normalization of the shearing time, we obtained a transient-degradation master curve at 73 s1 (Fig. B-1). As demonstrated in Fig. B-2,
the viscosity at low shear rate is a monotonic function of viscosity at high shear rate. Thus, although the exact shape of the master curve
may change, conclusions drawn at 73 s1 will apply at other shear rates.

80 100
V2 after V1

Degradation at 0.1 s–1 (%)


Degradation at 73 s–1 (%)

V2 80
60
V1
60
40
V2 after V1
40

V2
20
20
V1

0 0
0.01 1 100 0.01 1 100

Normalized Time (minutes) Normalized Time (minutes)

Fig. B-1—3630S polymer solutions at 1,100 ppm in a 6 g/L brine. Degradation in a laboratory blender at a rotational speed V1 or V2
or a combination of V1 and V2 . (Left) Degradation at 73 s21 vs. normalized time. (Right) The same normalized time, but degradation
is calculated at 0.1 s21.

40 100
Degradation at 0.1 s–1 (%)

35
Viscosity at 0.1 s–1 (%)

V2 after V1
80
30
V2
25 60
V1
20 V2 after V1
15 40
V2
10
20 V1
5
0 0
0 2 4 6 8 10 0 20 40 60 80
–1
Viscosity at 73 s (%) Degradation at 73 s–1 (%)

Fig. B-2—(Left) Relationship between viscosity measured at 0.1 s21 and that one measured at 73 s21 for samples with different
degradation histories. (Right) Relationship between degradation calculated from viscosity measurements at 0.1 and 73 s21.

Fig. B-1 presents transient-degradation master curves for 73 and 0.1 s1. The shapes differ because of the nonlinear correlation
between degradations at 0.1 and 73 s1.

Influence of Polymer Concentration. 3630S polymer solutions at different concentrations (100, 300, 800, and 1,400 ppm) in a brine
at 6 g/L were sheared in a laboratory blender at constant speed. The evolution of the viscosity at 0.1 s1 as a function of the shearing
time is plotted in Fig. B-3. The transient-degradation curve at 0.1 s1 of each solution is normalized by multiplying the shearing time
by a shift factor (graph at the center). If now, the degradation is calculated from viscosity measurements at 73 s1, the curves are no lon-
ger superposed. To keep the superposition, it is necessary to adjust the shift factors to new values. In fact, degradation measured at dif-
ferent shear rates is no more correlated for the different polymer concentrations, as seen in Fig. B-4. As previously explained, the
degradation measurement is not an intrinsic property of the polymer solution. Instead of using the degradation value, it would be neces-
sary to define an “absolute” parameter representative of the degradation, such as the average molecular weight, which would be inde-
pendent from physical-chemistry parameters. In place of the degradation defined as the viscosity loss, it would be interesting to follow
the loss of elastic properties, such as through screen-factor measurement.

February 2018 SPE Journal 31

ID: jaganm Time: 18:15 I Path: S:/J###/Vol00000/170047/Comp/APPFile/SA-J###170047


J186103 DOI: 10.2118/186103-PA Date: 12-January-18 Stage: Page: 32 Total Pages: 16

100 ppm in 6 g/L


100 100 100
300 ppm in 6 g/L

Degradation at 73 73 s–1 (%)


800 ppm in 6 g/L

Degradation at 0.1 s–1 (%)


Viscosity at 0.1 s–1 (%)
80 80
1,400 ppm in 6 g/L
10
60 60

40 100 ppm in 6 g/L 40


1 100 ppm in 6 g/L
300 ppm in 6 g/L
300 ppm in 6 g/L
800 ppm in 6 g/L
20 20 800 ppm in 6 g /L
1,400 ppm in 6 g/L
1,400 ppm in 6 g/L
0.1 0 0
0.01 1 100 0.01 1 100 0.01 1 100
Shearing Time (minutes) Normalized Time (minutes) Normalized Time (minutes)

Fig. B-3—(Left) Viscosity at 0.1 s21 of 3630S solutions at varying concentrations (100, 300, 800, and 1,400 ppm) vs. shearing time in
a laboratory blender at constant rotational speed. (Center) Viscosity measurements are converted in degradation and shearing
time is normalized. (Right) The same normalized time, but degradation is calculated from viscosity measurements at 73 s21 instead
of 0.1 s21.

100
Degradation at 0.1 s–1 (%)

80

60

100 ppm in 6 g/L


40
300 ppm in 6 g/L

20 800 ppm in 6 g/L

1,400 ppm in 6 g/L


0
0 10 20 30 40 50 60 70 80 90 100

Degradation at 73 s–1 (%)

Fig. B-4—Relationship between degradation calculated at 0.1 and 73 s21 for 3630S solutions at different concentrations (100, 300,
800, and 1,400 ppm) and different degradation levels.

Influence of Brine Salinity. 800-ppm polymer solutions prepared in two brines at 6 and 35 g/L were sheared in a laboratory blender at
constant speed. As described previously, shift factors (a) can be found that result in superposition of transient-degradation curves, but
shift factors calculated by use of degradation at 0.1 s1 do not result in superposition at 73 s1, as shown in Fig. B-5.

12 100 100 800 ppm in 35 g/L


800 ppm in 6 g/L
Degradation at 0.1 s–1 (%)

Degradation at 73 s–1 (%)

800 ppm in 35 g/L


Viscosity at 0.1 s–1 (%)

10
800 ppm in 6 g/L 80 80

8
60 60
6
40 40
4
800 ppm in 35 g/L
2 20 20
800 ppm in 6 g/L

0 0 0
0.01 1 100 0.01 1 100 0.01 1 100
Shearing Time (minutes) Normalized Time (minutes) Normalized Time (minutes)

Fig. B-5—(Left) Evolution of the viscosity at 0.1 s1 as a function of the shearing time for 3630S solutions at 800 ppm in brines at 6
and 35 g/L. (Center) Viscosity measurements are converted in degradation and shearing time is normalized. (Right) The same nor-
malized time, but degradation is calculated from viscosity measurements at 73 s1 instead of 0.1 s1.

Parameters Affecting the Transient-Degradation Master Curve. An empirical approach is presented that allows normalization of
the evolution of the viscosity measured at a given shear rate as a function of the number of degrading events.
In the absence of absolute measurement of degradation, such as an average Mw , the viscosity loss is normalized through the calculation
of the degradation with Eq. 1. The number of degrading events is normalized by multiplying the time in a mixer, the number of passes in a

32 February 2018 SPE Journal

ID: jaganm Time: 18:15 I Path: S:/J###/Vol00000/170047/Comp/APPFile/SA-J###170047


J186103 DOI: 10.2118/186103-PA Date: 12-January-18 Stage: Page: 33 Total Pages: 16

capillary, or the length in a porous medium by a shift factor that results in a superposition of the curves. The shape of this master curve
depends on the shear rate at which degradation is calculated, but not the salinity or polymer concentration. The shift factors necessary to
achieve superposition depend on all three factors.

Stephane Jouenne is a research engineer at Total, where he has led a team dedicated to surface and subsurface topics for
polymer flooding since 2005. His research interests include oil-recovery mechanisms with polymers, physical chemistry, rheology,
and process engineering. Jouenne has authored or coauthored 25 technical papers and holds eight patents. He holds a mas-
ter’s degree in chemical and process engineering from Nancy Process Engineering School, France, and a PhD degree in physi-
cal chemistry of materials from Paris 6 University, France.
Hafssa Chakibi is a PhD degree student at IFP Energies Nouvelles. Her research interests include comprehension of physical/
chemical interactions involved in produced-water treatment. Chakibi is an SPE member. She holds a master’s degree in chemi-
cal science and engineering from Chimie ParisTech.
David B. Levitt is an enhanced-oil-recovery subject-matter expert at Total, currently assigned to the Abu Dhabi National Oil Com-
pany. Previously, he worked at the Total Platform for Experimental RESEARCH center, as well as at the Abu Dhabi Marine Operat-
ing Company. Levitt has authored or coauthored more than 30 papers and patents. He has been a member of SPE since 2005
and has served on the technical committee for the SPE Improved Oil Recovery Conference in Tulsa since 2012. Levitt holds mas-
ter’s and PhD degrees, both in petroleum engineering, from the University of Texas at Austin.

February 2018 SPE Journal 33

ID: jaganm Time: 18:15 I Path: S:/J###/Vol00000/170047/Comp/APPFile/SA-J###170047

S-ar putea să vă placă și