You are on page 1of 10

SPE 109665

Development and Use of a Simulation Model for Mobility/Conformance Control Using a


pH-Sensitive Polymer
I. Benson, SPE, Chevron Energy Technology Co.; L.X. Nghiem, SPE, Computer Modelling Group Ltd.; S.L. Bryant, SPE,
M. M. Sharma, SPE, and C. Huh, SPE, U. of Texas at Austin

Copyright 2007, Society of Petroleum Engineers


found to be a useful dimensionless quantity for characterizing
This paper was prepared for presentation at the 2007 SPE Annual Technical Conference and acid floods with pH-sensitive polymer. Slugs of pH-sensitive
Exhibition held in Anaheim, California, U.S.A., 11-14 November 2007.
polymer improve oil recovery better than continuous polymer
This paper was selected for presentation by an SPE Program Committee following review of
information contained in a proposal submitted by the author(s). Contents of the paper, as
flooding or waterflooding.
presented, have not been reviewed by the Society of Petroleum Engineers and are subject to The simulator was successfully used to history match
correction by the author(s). The material, as presented, does not necessarily reflect any
position of the Society of Petroleum Engineers, its officers, or members. Papers presented at coreflood experiments, to model techniques to propagate low-
SPE meetings are subject to publication review by Editorial Committees of the Society of
Petroleum Engineers. Electronic reproduction, distribution, or storage of any part of this paper
pH fluids deep into reservoir, and to demonstrate the
for commercial purposes without the written consent of the Society of Petroleum Engineers is effectiveness of pH-sensitive polymer slug treatments for
prohibited. Permission to reproduce in print is restricted to a proposal of not more than 300
words; illustrations may not be copied. The proposal must contain conspicuous conformance control.
acknowledgment of where and by whom the paper was presented. Write Librarian, SPE, P.O.
Box 833836, Richardson, TX 75083-3836, U.S.A., fax 01-972-952-9435.
Introduction
Abstract Poor volumetric sweep or poor conformance, due to reservoir
Injection of a pH-sensitive polymer has been proposed heterogeneity, reduces oil recovery and increases operating
recently as a novel deep-penetrating mobility control method, costs of mature waterfloods. Mobility improvement methods
and the development of a simulation capability for its scale-up such as polymer flooding, and conformance control methods
is reported here. An aqueous dispersion of polymer microgel, such as workovers and polymer gel treatments, aim to improve
whose swelling property shows a strong dependence on pH, oil recovery by diverting injected fluid from high-permeability
preferentially flows into high-permeability zones under acidic zones to low-permeability, unswept areas of a reservoir.
conditions. Since the injected fluid viscosity is low, small Injection of a pH-sensitive polymer as a deep-penetrating
injection pressures are needed to inject the polymer. mobility control or conformance control method has been
Geochemical reactions increase the pH, causing the polymer- developed earlier in our laboratory (Huh et al. 2005; Choi et
containing fluid to experience a viscosity increase of several al. 2006). A solution of pH-sensitive polymers such as
orders of magnitude thereby altering the flow pattern of polyacrylic acid hydrogel can exhibit an orders-of-magnitude
subsequently injected fluid. Because the viscosity of microgel viscosity increase when its pH increases from acidic condition
dispersions can be controlled with adjustment of pH, this to above a threshold value. Such polymers are commercially
process can be employed both as a deep-penetrating mobility available at low cost in large quantities and are
control method (with moderate polymer viscosity) and as a environmentally benign. A major strength of the proposed
conformance control method (with immobile gel generation). method is that the pH increase needed to induce the drastic
Polymer-bank placement design and process scale-up viscosity increase in situ can occur naturally by geochemical
requires simulation of transport of microgel, acid-mineral reactions between acidic polymer solution and the reservoir
geochemical reactions, pH changes, and the coupling between rock. For the method to be effective, therefore, the ability to
aqueous phase composition and viscosity. Such a capability model the behavior of pH-sensitive polymer in situ in the
has been implemented in a commercial reservoir simulator and reservoir environment is essential. Such a capability has been
preliminary simulations verify the operation and effectiveness developed in a general-purpose reservoir simulator and is
of the complex new features, which can describe both the described in this paper.
mobility and conformance control applications. As discussed in detail earlier (Choi et al. 2006), the
Determination of reservoir mineralogy and mineral reaction controlled application of the proposed method requires a good
rates is critical to modeling in-situ pH changes accurately. understanding of its three sub-processes:
History matching of coreflood acid injection experiments was (1) dependence of polymer viscosity on ionic (pH)
used to estimate geochemical reactions and reaction rates conditions in the reservoir;
occurring in Berea cores. Linear and radial geometry floods in (2) geochemical characterization of pH change in the
2-layer reservoir models were carried out as preliminary scale- rock; and
up simulations. Acidic fluids can be propagated farther into a (3) polymer microgel transport in porous media.
reservoir in a low-pH state, using high injection velocity, an The first of the above has been addressed in detail by Huh et
acid preflush, or weak acids. The Damkohler number was al. (2005) and the second and the third by Choi et al. (2006).
2 SPE 109665

In the latter, history matching was used to estimate the mineral modeling parameters. Three corefloods performed by Choi
composition and reaction kinetics between hydrochloric acid (2005) were used for history matching the reaction rate of
and Berea sandstone. Estimation of geochemical reactions is hydrochloric acid with Berea sandstone.
critical to modeling the behavior of pH-sensitive polymer Ermel (2005) performed history matches of the same
since these reactions alter both the pH and ionic concentration corefloods using the KGEOFLOW simulator (Sevougian et al.
of the polymer solution. The geochemical characterization 1995) and concluded that a 2-mineral model could adequately
method employed in Choi et al. (2006) is significantly represent the matrix reactions. The present work focused on a
improved in the current effort. The use of the dimensionless 4-mineral Berea model composed of quartz, calcite, kaolinite,
Damköhler number (see Appendix B) was found useful for and potassium feldspar minerals. History matching
comparing pH-sensitive polymer flood simulations since it simulations were also performed with a one-mineral model,
combines the effects of geochemistry, acid concentration, which demonstrated some advantages over the four-mineral
convection rate and geometry into one value. model.
In the next Section, the new simulation capability to The corefloods were performed on 9” long, 1” diameter
model the deep-penetrating mobility control or conformance cores of Berea sandstone at 24ºC at atmospheric pressure. The
control using pH-sensitive polymer is briefly described. core porosities for Experiments 1, 2, and 3 were 0.20, 0.177,
History matching of coreflood acid injection experiments to and 0.1825, respectively. All three corefloods began with
estimate geochemical reactions and reaction rates occurring in injection of 3 wt% NaCl brine (pH = 7). Hydrochloric acid
Berea cores is then described. Preliminary scale-up (pH = 1) was then injected into the cores until the effluent pH
simulations employing linear and radial geometry floods in 2- approached 1. Fluid injection rates were varied for
layer reservoir models are given, followed by Conclusions. Experiments 1 and 3. The experiments included shut-in
periods ranging from a few minutes to several days. Figure 2,
Simulation Model Description Figure 3, and Figure 4 show the effluent pH and fluid injection
The key feature of the new simulation capability is the rates versus pore volumes of fluid injected, for Experiments 1,
implementation of the pH-sensitive polymer rheological 2 and 3, respectively. Effluent concentrations of silicon,
model, developed by Huh, Choi and Sharma (2005), into a 3- aluminum, calcium and potassium were measured for
D compositional reservoir simulator capable of modeling Experiment 2 and are shown elsewhere (Choi, 2005).
geochemical reactions (GEM-GHG). The enhanced simulator As the preliminary step for history matching, initial
allows prediction of polymer behavior in situ. The rheological simulation efforts focused on identifying the key minerals for
model predicts the apparent viscosity of polymer as a function the simulation model. Berea compositions cited in literature
of pH, shear rate, polymer concentration and ionic strength. It include a large portion of quartz, and lesser amounts of clays,
is used to calculate water phase viscosity in each grid cell carbonates and feldspars (Ermel, 2005). 1-D domains with 50
every timestep. The keywords used to enter the rheological gridblocks were employed for simulation, in which various
model constants are available elsewhere (Benson, 2007). Grid combinations of quartz, calcite, illite, kaolinite, potassium
block properties used to calculate aqueous phase viscosity feldspar, and anorthite are assumed to participate for
include pH, ion and polymer concentrations, fluid velocity, geochemical reactions. The simulations were run with pH = 1
water saturation, permeability and water phase relative hydrochloric acid injected, and the maximum pH attained
permeability. Figure 1 demonstrates the large effect of pH on when the acid was fully reacted with the rock was recorded.
polymer viscosity for a 3 wt% EZ-2 polymer solution (ionic The experiments showed effluent pH greater than 5, and this
strength = 0.5, no flow; Huh et al. 2005). preliminary step determined which minerals would allow the
The ability of the simulator to model geochemical pH to exceed 5. All geochemical models tested initially
reactions occurring between reservoir minerals and the ions in contained mineral reaction and equilibrium constants (at 25ºC)
the aqueous phase, coupled with the rheological model, allows from Stumm and Morgan (1996). High mineral reaction rates
the modeling of the desired mobility and conformance control were chosen for the simulations to ensure that the acid fully
effects. The simulator models reservoir mineral reactions as reacted with all the reservoir minerals before exiting the core.
rate-limited heterogenous reactions. Appendix A describes the Models containing only clay, potassium feldspar, or quartz did
reaction rate calculation. Mineral reactions cause minerals to not allow the pH to rise above 5. Models containing one clay
dissolve or precipitate in a reservoir and change ion mineral (illite or kaolinite), potassium feldspar and quartz
concentrations in reservoir brine. The simulator models intra- allowed the acid to form brine with a pH in excess of 5.
aqueous reactions within the water phase as instantaneous Models containing calcite allowed the formation of a brine
reactions, i.e. reactions that attain chemical equilibrium with pH~5.4, or higher if a clay, quartz and potassium feldspar
immediately. Geochemical reactions control the ionic strength were present.
and pH of the reservoir brine which are key inputs to the pH- A Berea mineralogy consisting of kaolinite, calcite, quartz
sensitive polymer rheological model. and potassium feldspar was chosen for more detailed history
matching of the effluent concentration histories of Experiment
Core Flood History Matching 2. The objective here was to determine mineral reaction rate
Proper selection of mineral components, chemical reactions constants and mineral reactive grain fractions (see Appendix
among them, mineral reaction rate models and their parameter A). Initial estimates of kaolinite, calcite, quartz and potassium
values, is necessary to model the behavior of pH-sensitive feldspar grain fractions were 6%, 3%, 88% and 3%,
polymer in situ. History matching of the effluent composition respectively. The mineral reactions, intra-aqueous reactions,
of laboratory acid corefloods was used to select geochemical and equilibrium constants shown in Table 1 and Table 2 were
SPE 109665 3

used for these history match runs. Published reaction-rate pH to decrease from ~5.4 in all the models performed in
constants (10-13, 10-8.8, 10-13.9, 10-12 mol·m-2s-1, respectively) this study.
compiled by Xu et al. (2001) were used for the initial history • Kaolinite reaction behavior is conjectured to be controlled
match runs. The mineral reaction rates were entered as by the calcite reaction through the H+ concentration.
constant values, kβ (see Appendix A). The published rate Kaolinite dissolves from the inlet until a point 3 inches
constants were too small to allow the pH = 1 acid to produce into the core. The large decrease in H+ concentration 3
pH ~5.4 effluent in Experiment 2. The acid injection rate of 2 inches from the core inlet appears to cause strong
cm3/min gave an acid residence time of about 9.7 minutes in kaolinite precipitation until a point 6.7 inches from the
the second coreflood experiment. The mineral reaction-rates acid inlet.
for the four minerals were therefore changed to 10-7.2, 10-4.07, • Quartz behavior opposes that of kaolinite. Quartz
10-6.22, and 10-5.98 mol·m-2s-1, respectively. These are referred precipitates where kaolinite is dissolving from the core
to as the “initial-match parameters” in Table 3. The mineral inlet until a point 3 inches from the inlet. Quartz is
concentrations were changed to those given in Table 4 to give strongly dissolving from 3 to 6.7 inches in the core where
a reasonable initial match for the effluent pH and ion kaolinite is precipitating. The quartz and kaolinite
concentrations for the second coreflood. reactions are strongly coupled by the silica (H4SiO4(aq))
The mineral reaction rates from the initial history match concentration in the model.
of Experiment 2 were then used in simulations of Experiments • Potassium feldspar (k-feld) dissolves continuously along
1 and 3. The effluent pH predicted from the “initial match” the core despite changes in H+ and silica concentrations.
reaction rate constants did not match observed effluent pH
data. It was necessary to increase the calcite reaction rate by a Sensitivity to One-Mineral Model with First-Order Mineral
factor of 5 to prevent the simulated effluent pH from dropping Reaction Rate
during the high acid injection rates in Experiment 3. The The 4-mineral final match coreflood simulation models were
quartz, kaolinite and potassium feldspar reaction rates were altered to create 1-mineral models with a first-order mineral
decreased by a factor of 2 to match the number of pore reaction rate. Calcite was chosen as the mineral because it
volumes of pH=1 acid that had to be injected before the demonstrated a spent-acid effluent pH close to that observed
effluent pH from the coreflood dropped from about 5 to about experimentally. The other minerals in the original simulations
1.5. The resulting “final-match” reaction parameters and core were removed. Near perfect agreement between the simulated
mineralogy are given in Table 3 and in Table 4, respectively. and experimental coreflood effluent pH resulted when a
Figures 2, 3 and 4 show, respectively, the simulation matches mineral reaction rate first-order in hydrogen concentration was
of Experiments 1, 2 and 3. The simulations agree reasonably used. The new reaction rate was modeled by setting kβ near
well with the laboratory effluent data. zero in the reaction equation (Eq. [A-1]), and setting the
exponent for the H+ activity ωβ to 1, and varying reaction
Simulated Mineral Interaction Effects constant ki,β for the hydrogen activity along with calcite
Mineral interactions in the 4-mineral simulation model created concentration until a match was achieved.
pH = 3 plateaus in the effluent in some runs. The reaction Figure 6, Figure 7, and Figure 8 show effluent pH from
affinity term for each mineral’s reaction (see Appendix A) was the simulations and the experiments. These matches had a
plotted versus pore volumes injected to understand mineral calcite volume fraction of 0.051 (compared to 0.047 in 4-
behavior in the final simulation of core flood Experiment 2. mineral model) and kH+,calcite set to 10-1.37 mol·m-2·s-1. One
⎣ ( )
The reaction affinity is calculated as ⎡1 − Qβ K eq , β ⎤ for each

drawback of this method was that the final effluent pH was 1,
not ~1.3, after all the calcite was dissolved. Another is that
mineral based on the ion activity product, Qβ, for each mineral although the pH matched well, calcite does not contain
reaction in each cell in the 1-D coreflood simulation. Positive aluminum, silicon, or potassium that were also measured in
reaction affinity values indicate mineral dissolution and the experiments. It is possible that a mineral or pseudo-
negative ones indicate precipitation. Figure 5 shows the mineral that can match those effluent readings could be found
dimensionless mineral concentration (fraction of initial or generated. However, the main objective here is to predict
concentration), reaction affinity, and selected ion the propagation of the pH front, because this controls the
concentrations for the final match simulation after 50 pore transition between small and large viscosity. From this
volumes of fluid were injected. Coreflood 2 involved injection perspective, the 1-mineral model has the advantage of being
of 11 pore volumes of pH = 7, 3 wt% brine before injection of simpler to apply.
pH = 1 hydrochloric acid began.
Figure 5 provides valuable insight on the mineral Sensitivity to Weak Acid Injection
interactions as the acid flows in the core: Simulations testing the ability of weak acids to propagate low-
• Calcite is dissolving at high rates between the core inlet pH conditions farther into a linear reservoir than a strong acid
and a point 3.5 inches from the inlet. Calcite reaction were performed. Cases were run using both the 4-mineral
affinity is +1 from the inlet to a point 3 inches in the core, matrix with fixed reaction rates and the 1-mineral matrix with
and decreases to a slightly positive value at 3.5 inches. All first-order mineral reaction rates. Both mineralogies yielded
mineral reaction rates are dependent on the mineral similar results. Simulations compared the ability of pH=2
concentration through the surface area value in Eq. [A-1]. solutions of hydrochloric acid, citric acid and acetic acid to
Arrival of the calcite dissolution front causes the effluent maintain a low-pH away from the injector. It was found that
4 SPE 109665

the weak acids (citric acid and acetic acid) propagated a low- the high-permeability zone were modeled by changing the
pH fluid front farther into the reservoir than the strong acid fixed reaction constant of calcite to 10-10.447, 10-9.447, 10-8.447,
case (hydrochloric acid). This is true only when the injected 10-7.447 and 10-6.447 mol·m-2s-1, respectively. The flow rate into
acid pHs are similar, not when the injected acid concentrations the high-permeability zone, estimated by prorating q by the
are similar. The simulations showed drawbacks of using weak layer kh products, was used in these Damkohler numbers. The
acids are that large acid concentrations are required to achieve actual fluid velocity changes in both zones as the pH-sensitive
low-pH, and they achieve a lower equilibrium pH when they polymer alters flow distribution between the two zones. The
are spent. reaction rate constants used here are orders of magnitude
smaller than the history matched reaction constants found
Conformance Control Using pH-Sensitive Polymer: from the Berea coreflood.
Linear Geometry Model Figure 10 shows the volume of oil produced for 1000-cp
Conformance control is a method that improves oil recovery pH-sensitive polymer floods simulated at various Damkohler
and production rates by diverting injected fluid in a flooding numbers (NDa). Predictably, the NDa = 0.01 case behaved more
operation into lower permeability unswept regions. 2-layer like 1 cp water and the NDa = 10 and NDa = 100 cases behaved
simulations with linear and radial grid geometries were as if they were 1000-cp fluids. The NDa = 1 case is remarkable
created to test the ability of pH-sensitive polymer to improve since it greatly improved recovery at early times (before 0.2
conformance. Comparisons were made between pore volumes injected) but ultimately caused recovery to
waterflooding, conventional polymerflooding, pH-sensitive suffer. This important result occurred because at early times
polymer flooding and polymer slug treatment cases. All the pH-sensitive polymer penetrated the high-permeability
polymer and pH-sensitive polymer cases neglected shear- zone, increased its flow resistance and automatically diverted
thinning rheology though this feature exists in the rheological polymer into the low-permeability zone which improved early
model. Effectively the fluids had Newtonian behavior. recovery. Unfortunately, the polymer injected into the low-
The linear and radial models documented here share many permeability zone experienced a viscosity increase too and
similarities. The injection rate was set at 25 m3/d, though a 15 then plugged off the low-permeability zone almost
MPa pressure limit was applied. The reservoir contained a completely. This result occurs when the ratio of the pH-
0.047 volume fraction of calcite, and the reaction rate constant sensitive polymer viscosity at final and initial conditions is
was varied to provide a range of Damkohler numbers (see greater (say 10 times or more) than the permeability contrast
Appendix B) in the pH-sensitive polymer flood cases. Initial between the two zones.
oil saturation was 0.8, residual oil saturation was 0.3. The Polymer slug injection simulations were performed to
hydrocarbon pore volume (HCPV) was 16,000 m3, pore build on the partial success of the NDa=1 case. The concept is
volume was 20,000 m3, and moveable oil was 62.5% HCPV. that by injecting a slug of polymer into both zones at low
Oil viscosity was 1.09 cp and brine viscosity was 1 cp. All viscosity, the polymer will preferentially penetrate the high-
cases had a producer BHP and initial reservoir pressure of 10 permeability zone. Polymer injection can then be terminated.
MPa. The polymer will reach maximum viscosity in situ, and water
Figure 9 is a schematic of the linear 2-layer model created injected afterwards will have an increased preference to
for sensitivity runs to investigate the effects of various invade the low-permeability zone. This is conceptually very
polymer treatments on oil recovery. While the linear geometry similar to conventional gel treatments, except that the
is highly simplistic, it may approximately represent injection viscosity increase is not triggered by environmentally harmful
from a horizontal well or from a vertical fracture. and possibly carcinogenic crosslinkers (Kabir, 2000). In
Table 5 gives the reservoir properties. The linear model addition other problems such as poor in-situ mixing,
has 100 cells in horizontal direction with 10:1 grid refinements chromatographic separation of the cross-linker and possible
near the injection well. Linear cases presented here do not interference with other metal ions are avoided.
allow crossflow between layers but such cases are documented Figure 11 shows oil recovery versus volume of fluid
by Benson (2007). injected for some polymer slug treatments. The pH-sensitive
Water and 10-cp, 100-cp, and 1000-cp viscosity polymers polymer treatment was injected at a Damkohler Number of
(approximately representing the low-shear limit viscosities in 0.1. Results from slugs of regular polymer are shown although
the rheological model; Huh et al. 2005) were injected into the it may prove difficult to inject fluid with viscosities in excess
linear model. Figure 10 shows the volume of oil produced of 10 cp in a real reservoir. All slug volumes were 0.05 HCPV
versus volume of fluid injected. Increasing the viscosity of 3
(800m ) and injected beginning at time zero. Shut-in time was
injected fluid improves recovery (and conformance) until allowed for the pH-sensitive polymer to react, then
viscosity reaches ~ 100 cp. Injection of high viscosity fluids waterflooding commenced. The improvement in oil recovery
reduces injectivity greatly. dwarfs that observed for continous injection of high viscosity
A pH=1 mixture of hydrochloric acid and pH-sensitive polymer. Figure 12 shows the water phase viscosity
polymer (EZ-2, 3 wt% NaCl; Huh et al. 2005) was injected immediately after the pH-sensitive polymer slug fully reacted,
into the model at various Damkohler numbers. A polymer and the position and reduced slug viscosity after a total of 0.5
concentration of 1.93×10-5 mol/kg gave a polymer viscosity of pore volumes of fluid were injected. This result does not
2.2 cp at pH = 1 and a final viscosity of 1000 cp in situ when include the negative effects caused by crossflow and viscous
the acid reached equilibrium with the calcite reservoir at fingering that would occur in a reservoir but qualitatively
pH=5.39. Damkohler numbers of 0.01, 0.1, 1, 10 and 100 in
SPE 109665 5

suggests that slug treatments may improve conformance the reactive reservoir mineral is consumed near the wellbore.
significantly. This allows the polymer to penetrate farther away from the
well as a low viscosity fluid before its pH reaches it maximum
Conformance Control Using pH-Sensitive Polymer: value in situ. The wavy flow rate profile for the pH-sensitive
Radial Geometry Model polymer case in Figure 17 was observed for many radial
To test conformance control in a radial geometry (Figure 13), simulations but not in any linear simulations.
the linear model described above was converted to a radial
model with the injector in the center. The reservoir and fluid Conclusions
volumes were kept the same. The well radius was 0.1 m, and ƒ A numerical simulator was successfully used to model the
the drainage radius was 42.4 m. The grid had 61 cells with behavior of pH-sensitive polymer in reactive porous
radial dimensions increasing away from the wellbore. The media. Preliminary simulations show that pH sensitive
radial reservoir model has high pressure drops near the polymer slug treatments can improve conformance in 2-
injector and is much less sensitive to viscosity changes away layer radial and linear geometry floods.
from the wellbore. Sorbie and Mackay (2005) discuss the ƒ History matching of coreflood experiments allows
large differences between polymer flood behavior in radial suitable mineralogy and kinetics to be found for acid-
and linear geometry floods. mineral reactions. A 4-mineral model of Berea sandstone
Water and 10-cp, 100-cp and 1000-cp fluid were injected matched effluent data well. An approximate 1-mineral
into the central injector to examine the effect of viscosity on model with a first-order in hydrogen reaction rate also
oil recovery. Figure 14 shows oil recovery versus volume of matched effluent data reasonably well.
fluid injected. Increasing the viscosity of the injected fluid has ƒ pH-sensitive polymer should be positioned in reservoir as
a much smaller impact on oil recovery in the radial a low viscosity fluid before fully reacting with reservoir
simulations compared to linear simulations. minerals to form a high viscosity plug. Simulations
Figure 15 shows oil recovery versus the volume of fluid showed high injection rates and weak acids to be effective
injected when a 0.05 HCPV polymer slug was injected to both in forming such a plug.
layers before waterflooding. The 1000-cp pH-sensitive case ƒ The use of conventional high-viscosity polymers and pH-
recovered oil most efficiently in terms of volume of fluid sensitive polymer improves vertical conformance in linear
among the slug treatments, and the slug treatments did not reservoirs with and without vertical crossflow.
improve recovery much compared to the waterflood (1-cp Continuous injection of pH-sensitive polymer in these
fluid) case. The 100-cp fluid case recovers oil faster than any reservoirs is not recommended since this can plug the
of the slug cases for the radial simulations without crossflow. low-permeability zones in a reservoir. Simulations
The aqueous-phase viscosity distribution in the radial reservoir suggest that the use of pH-sensitive polymer slug
during the 1000-cp pH-sensitive polymer slug case (not shown treatments is the preferred way to improve vertical
here) reveals that the slug reaches high viscosity near the conformance in a linear geometry during waterflooding.
injection well in both layers due to geochemical reactions. ƒ Improved injectivity can be achieved using pH-sensitive
However the slug does not remain stationary in the high- polymer compared to a conventional polymer. Shear-
permeability layer but is forced out by waterflooding. thinning affects both types of polymer and was not
According to the viscosity profile after 0.1 PV injected, high- considered here.
viscosity fluid remains in the low-permeability zone, which
explains why slug treatments injected into both layers tend to Acknowledgements
lower oil recovery. We appreciate the efforts of Suk Kyoon Choi and Farshad
Figure 16 shows oil recovery versus the volume of fluid Lalehrokh to convert the pH-sensitive polymer rheological
injected for radial cases with and without vertical crossflow model into a computer routine. Thanks to Joanna Castillo and
when a 0.05 HCPV polymer slug is injected only into the Roger Terzian for supporting the simulator software and
high-permeability layer before waterflooding. These pH- hardware. The first author was supported by DOE grant DE-
sensitive polymer treatments significantly improved oil FC26-04NT15520.
recovery over the cases where polymer was injected into both
layers. This suggests the slug treatments can be effective for Nomenclature
radial reservoirs when they are applied to the high- Âβ = reactive surface area of reactive mineral β per unit
permeability thief zones only. The volume of the slug and the bulk volume of porous medium (m2/m3)
final viscosity of the slug will greatly affect the performance Sw = water saturation
of the treatment. Crossflow appears to lower the effectiveness kβ = rate constant of mineral reaction β (mol/m2·s).
of the high-permeability zone treatments considerably. ki,β = rate constants of mineral reaction β related to
activity of each species (default ki,β = 0 in GEM).
Improved Injectivity Using pH-Sensitive Polymer Index “i” refers to each species (both hydrocarbon
A radial simulation was performed to demonstrate improved and aqueous) in the simulation.
injectivity of 100-cp pH-sensitive polymer compared to 100- ω
ai ,βi = activity of ions for mineral reaction β
cp fluid. Both cases have a constant injection pressure of 15
MPa, and shear thinning was not incorporated into either Keq,β = chemical equilibrium constant of mineral reaction
model. Figure 17 shows the injected fluid rates versus time. Qβ = activity product of mineral β dissolution reaction
The pH-sensitive polymer injectivity increases with time as rβ = dissolution/precipitation rate per unit bulk volume
6 SPE 109665

of porous medium [mol/(m3·s)] Appendix A: Mineral Reaction Rate Equation Used in


Rmn = number of mineral reactions GEM-GHG Simulator
ζβ = parameter in general reaction rate expression GEM-GHG uses the following formula to calculate mineral
(default = 1) reaction rates for each grid cell:
NDa = Damkohler number ξβ ς β
⎡ ⎛ Qβ ⎞⎤ ⎛ naq
⎞ ⎛ Qβ ⎞
L = length of medium (m) ⎟⎟ ⎥ Â β S w ⎜⎜ k β + ∑ ki ,β ai , β i
ω
rβ = sgn ⎢1 − ⎜ ⎟⎟ 1 − ⎜⎜ ⎟⎟
A = area perpendicular to flow (m2) ⎢⎣ ⎜⎝ K eq , β ⎠ ⎥⎦ ⎝ i =1 ⎠ ⎝ K eq , β ⎠
q = acid flow rate (m3/s)
a = H+ molality (mol/kg) [A-1]
rH + = reaction-rate of H+ (mol/m3·s). In the above equation:
νH+ , νβ = stoichiometric coefficient of H+ and mineral β in
reaction equation
1) The sign of the reaction affinity term, ⎡1 − Qβ K eq , β ⎤ ,
⎣ ⎦ ( )
determines whether a mineral is dissolving or precipitating.
References 2) Fixed reaction rate constants (kβ), reaction rate dependent
1. Benson, I., “Numerical Simulation of pH-Sensitive on individual ion activity (ki,β), or combinations thereof can be
Polymer Injection as a Conformance Control Method”, used. The best history matches were obtained for the
M.S. Thesis, University of Texas at Austin, May, 2007. hydrochloric acid flood experiments when reaction rates were
2. Ermel, Y. M., “Application of Geochemical Modeling to first-order in hydrogen concentration.
Optimize a pH-Sensitive Gel System”, M.S. Thesis, 3) The mineral surface area in the reaction equation makes
University of Texas at Austin, May, 2005. reactions first-order in mineral concentrations in the simulator.
3. Huh, C., Choi, S. K., and Sharma, M. M., “A Rheological 4) Setting the ζβ exponents to zero removes the ability of the
Model for pH Sensitive Ionic Polymer Solutions for reaction affinity value to control the reaction rate.
Optimal Mobility Control Applications”, SPE 96914
presented at SPE Annual Technical Conf., Dallas, Texas, Appendix B: Calculation of Damköhler Number
October 9-12, 2005. The Damköhler number, NDa, is a dimensionless number used
4. Kabir, A. H., “Chemical Water & Gas Shutoff to compare the acid reaction rate to acid convection rate. This
Technology”, SPE 72119 presented at SPE Asia Pacific value includes the effects of mineral-acid stoichiometery,
Improved Oil Recovery Conf., Kuala Lumpur, October 8- kinetics, acid concentration and Darcy velocity.
9, 2001. The Damköhler number is based on hydrogen
5. Choi, S. K., “A Study of a pH-Sensitive Polymer for Novel consumption rate, which is closely related to the mineral
Conformance Control Applications”, M.S. Thesis, reaction rate (rβ) given in Equation A-1. The total hydrogen
University of Texas at Austin, December, 2005. consumption rate must be used when more than one reactive
6. Choi, S. K., Ermel, Y. M., Bryant, S. L., Huh, C., and mineral is present:
Sharma, M. M., “Transport of a pH-Sensitive Polymer in νH +

Porous Media for Novel Mobility Control Applications”, rH + = r [B-1]


υ min eral β
SPE 99656 presented at SPE/DOE Symp. Improved Oil
Recovery, Tulsa, OK, Apr. 22-26, 2006.
7. Nghiem, L., “Component and Reaction Data Input for The value of NDa depends on the flow layer geometry. For
GEM-GHG (Version 2004.19A)”, Computer Modelling linear or 5-spot geometry the equation is:
Group, December 16, 2005.
8. Palandri, J. L., Kharaka, Y. K. 2004. “A Compilation of LArH +
N Da = [B-2]
Rate Parameters for Water-Mineral Interaction Kinetics for ρ acid qa
Application to Geochemical Modeling”, U. S. Geological
Survey, Open File Report 2004-1068, Menlo Park,
where L = length of medium (m), A = area perpendicular to
California, March 2004.
flow (m2), q = acid flow rate (m3/s), a = H+ molality (mol/kg),
9. Sorbie, K. S., and Mackay, E. J., “Scale Inhibitor
Placement: Back to Basics – Theory and Examples”, SPE rH + = reaction-rate of H+ (mol/m3·s) and ρacid = density of the
95090 presented at SPE International Symp. Oilfield Scale, aqueous phase (the injected acid), taken to be 1000 kg/m3 for
Aberdeen, United Kingdom, May 11-12, 2005. the purpose of evaluating Eq. [B-2].
10. Stumm, W., and Morgan, J. J., Aquatic Chemistry rd For radial flow layer geometry the equation is:
Chemical Equilibria and Rates in Natural Waters, 3
Ed., John Wiley and Sons, New York, 1996. π ⋅ Re2 ⋅ h ⋅ rH +
11. Xu, T., Apps, J. A., and Pruess, K., “Analysis of Mineral N Da = [B-3]
Trapping for CO2 Disposal in Deep Aquifers”, Lawrence ρ acid qa
Berkeley National Laboratory, Paper LBNL-46992, 2001. where Re is the outer radius of the reservoir layer (assuming
Re >>Rwellbore)
SPE 109665 7

Table 1: Mineral Reactions and Equilibrium Constants (at Table 4: Mineral Concentrations, Reactive Surface Areas
25°C) and Activation Energies

Mineral Reaction log Keq Specific


Reactive
Activation
Grain Volume Area
Quartz SiO2(s) + 2 H2O = H4SiO4(aq) -3.98 Mineral Energy
Fraction Fraction (m2/[m3 of
(J/mol)
Bulk
Al2Si2O5(OH)4(s) + 6 H+(aq) = Volume])
Kaolinite 7.435
2 Al3+(aq) + 2 H4SiO4(aq) + H2O
Quartz 0.846 0.696 6960 87500
Calcite CaCO3(s) = Ca2+(aq) + CO3--(aq) -8.42
Kaolinite 0.061 0.05 50000 62760
Potassium KAlSi3O8(s) + 8 H2O =
-20.573 Calcite 0.057 0.047 470 41870
feldspar K+(aq) + Al(OH)4-(aq) + 3 H4SiO4(aq)
Potassium
0.036 0.03 300 67830
Table 2: Intra-Aqueous Reactions Equilibrium Constants feldspar
for Four-Mineral Simulations
Table 5: Properties of Two-Layer Linear Model
Equilibrium Reaction Log Keq @ 25ºC Property Layer 1 (top) Layer 2 (bottom)

HCO3- = H+ + CO3-- -10.329 Porosity 0.177 0.177

Al+++ + 4 H2O = Al(OH)4- +4 H+ -22.7 Height 18 m 2m

H2CO3 = 2 H+ + CO3-- -16.6 Length 100 m 100 m

H2O = H+ + OH- -14 Width 56.497 m 56.497 m


X & Y-direction k 100 md 10000 md
Table 3: Mineral Reaction Constants (at 25°C) from Z-direction k 0 md or 10 md 0 md or 1000 md
Literature and History Matching
krw 0.3[(Sw-0.2)/(1-0.2-0.3)]2
Rate constant kβ, moles m-2 s-1 kro 0.7[(1-Sw-0.3)/(1-0.2-0.3)]2
Mineral
Reaction U.S.G.S.
Initial-match
(Palandri and Final-match
Kharaka, 2004)
-13.4 -6.22 -6.52 1000000
Quartz 10 10 10
-13.2 -7.2 -7.5 100000
Kaolinite 10 10 10
Polymer Viscosity, cp

-0.3 10000
10 [H+] -4.07 -3.37
Calcite – 5.81 10 10
+10 1000

Potassium -12.4 -5.98 -6.28


feldspar 10 10 10 100

10

1
0 2 4 6 8 10
pH

Figure 1: Calculated Polymer Viscosity for 3wt% EZ-2


Polymer Solution (Ionic Strength = 0.5, No Flow)
8 SPE 109665

12 12 1.5
Core Flood Effluent pH

Reaction Affinity
1
Initial Effluent pH Match
10 10 0.5

Fluid Injection Rate, cc/min


Final Effluent pH Match
0
Brine Injection Rate
8 8 -0.5
pH 1 HCl Acid Injection Rate
-1
pH

6 6 -1.5 Calcite Kaolinite


K-feldspar Quartz
-2
4 4
1.25

Mineral Conc.
2 2 1

0.75
0 0
0 50 100 150 200 250 0.5
Pore Volumes Injected
0.25 Calcite Kaolinite
Figure 2: pH versus Volume of Acid Injected for Coreflood K-feldspar Quartz
Experiment 1 0
12 12 0.1

Ion Concentration
Core Flood Effluent pH
Initial Effluent pH Match 0.001
10 Final Effluent pH Match 10
1E-05
Fluid Injection Rate, cc/min

Brine Injection Rate H+


pH 1 HCl Acid Injection Rate H4SiO4
8 8 1E-07
K+
1E-09 Ca++
Al+++
pH

6 6
1E-11
0 1 2 3 4 5 6 7 8 9
4 4
Distance From Inlet (inches)
2 2
Figure 5: Reaction Affinity (for Mineral Reactions),
0 0 Mineral Concentrations and Select Ion Concentration
0 50 100 150 200 250 Profiles for Experiment 2 After 50 Pore Volumes Injected
Pore Volumes Injected
(Using Final-Match Reaction Parameters)
Figure 3: pH versus Volume of Acid Injected for Coreflood
Experiment 2 12
12 12 Core Flood Effluent pH
Core Flood Effluent pH
Initial Effluent pH Match 10 Final Effluent pH Match
10 Final Effluent pH Match 10
Fluid Injection Rate, cc/min

Brine Injection Rate Calcite Model with First Order


8
pH 1 HCl Acid Injection Rate Mineral Reaction
8 8
pH

6
pH

6 6

4
4 4

2
2 2

0
0 0 0 50 100 150 200 250
0 50 100 150 200 Pore Volumes Injected
Pore Volumes Injected
Figure 6: Effluent pH for Calcite Model with First-order
Figure 4: pH versus Volume of Acid Injected for Coreflood Calcite Reaction (Experiment 1)
Experiment 3
SPE 109665 9

12 0.5
Core Flood Effluent pH Target Qinj = 25 m3/d for all cases
NDa = 100
-Only 1 cp, 10 cp and 1000 cp
Final Effluent pH Match fluid cases subject to 15000 kPa
10

Produced Oil Volume (HCPV)


0.4 injection well pressure limit
Calcite Model with First Order Mineral Reaction

8 1000 cp fluid (gray)


NDa = 1 NDa = 10
0.3
pH

0.2
4 NDa = 0.01
1 cp fluid (gray)
0.1
2 NDa = 0.1

0 0
0 50 100 150 200 250 0 0.2 0.4 0.6 0.8 1
Pore Volumes Injected Pore Volumes Injected
Figure 7: Effluent pH for Calcite Model with First-order Figure 10: Produced Oil Volume versus Volume of Fluid
Calcite Reaction (Experiment 2) Injected (Linear Case-No Crossflow)
12
Core Flood Effluent pH
0.7
10 Final Effluent pH Match
0.6
Calcite Model with First Order Mineral Reaction 1000 cp pH sens. polymer slug

Produced Oil Volume (HCPV)


1000 cp fluid slug
8
0.5
1000 cp fluid flood (gray)
pH

6 0.4

4 0.3
100 cp fluid slug
0.2
2
1 cp fluid flood (gray)
0.1
0 10 cp fluid slug
0 50 100 150 200 0
Pore Volumes Injected
0 0.2 0.4 0.6 0.8 1
Figure 8: Effluent pH for Calcite Model with First-order Pore Volumes Injected
Calcite Reaction (Experiment 3) Figure 11: Oil Recovery versus Volume of Fluid Injected
Using 0.05 HCPV Slug Treatment Followed by
Waterflooding (Linear Case, No Crossflow)
Injection Production
Well Well

Layer 1 (h1, Φ1, k1)

Layer 2 (h2, Φ2, k2)


X

Figure 9: Schematic of Two-Layer Linear Oil Reservoir


Simulation Grid

Figure 12: Aqueous Phase Viscosity Distribution in Linear


Reservoir During 1000-cp pH-Sensitive Polymer Slug
Treatment Simulation (Linear Case, No Crossflow)
10 SPE 109665

Injection Well Production Well * 0.7

1000 cp pH-sensitive
0.6
polymer slug treatment

Produced Oil Volume (HCPV)


(reservoir without crossflow)
Zone 1 0.5 1000 cp pH-sensitive
polymer slug treatment
(reservoir with crossflow)
0.4

0.3
Possible Flow Barrier

Zone 2 0.2

0.1
* Fluid production is from entire perimeter of reservoir
1 cp waterflood (with crossflow, gray)
Figure 13: Schematic of Two-Layer Radial Oil Reservoir 0
Simulation Grid 0 0.2 0.4 0.6 0.8 1
Pore Volumes Injected

0.25 Figure 16: Oil Recovery versus Volume of Fluid Injected


Using 0.1 HCPV Slug Treatment Followed by
NDa = 0.1
Waterflooding (Radial Case)
Produced Oil Volume (HCPV)

0.2
1000 cp fluid flood (Gray) NDa = 0.01 400
pH Sensitive Polymer (100 cp max)
1 cp fluid flood (Gray) 350
0.15 100 cp fluid

Fluid Injection Rate (m3/d)


300

0.1 250
NDa = 10 NDa = 100
200
NDa = 1
0.05
150

100
0
0 0.2 0.4 0.6 0.8 1
50
Pore Volumes Injected

Figure 14: Oil Recovery versus Volume of Fluid Injected 0


Using pH-Sensitive Polymer (Radial Case, No Crossflow) 0 100 200 300 400 500 600 700 800
Time (Days)

0.2 Figure 17: Injection Rate versus Time (Constant Injection


1000 cp fluid flood (gray) Pressure, Radial Case, No Crossflow)
10 cp fluid slug
Produced Oil Volume (HCPV)

0.15
100 cp fluid slug

0.1
1000 cp fluid slug

0.05
1 cp fluid flood (gray)
1000 cp pH sens. polymer slug

0
0 0.2 0.4 0.6 0.8 1
Pore Volumes Injected

Figure 15: Oil Recovery versus Volume of Fluid Injected


Using 0.05 HCPV Slug Treatment Followed by
Waterflooding (Radial Case, No Crossflow)

You might also like