Sie sind auf Seite 1von 7

Thickening Carbon Dioxide With the

Fluoroacrylate-Styrene Copolymer
Jianhang Xu, Aaron Wlaschin, and Robert M. Enick, SPE, U. of Pittsburgh

Summary DeSimone and coworkers were the first to demonstrate that a


high molecular weight polymer, poly(1,1-dihydro-perfluorooctyl-
The fluoroacrylate-styrene copolymer is the first associative thick-
acrylate), PFOA, could exhibit remarkably high CO2 solubility in
ener that has been identified for carbon dioxide. Fluoroacrylate is
the absence of a cosolvent.1,10 Falling cylinder viscometry was
highly carbon dioxide-philic, enhancing the solubility of the co-
used to demonstrate that PFOA could also increase the viscosity of
polymer in carbon dioxide. Styrene is relatively carbon dioxide-
CO2. For example, at 323 K and 34 MPa, the addition of 6.7wt%
phobic, but promotes viscosity-enhancing, intermolecular associa-
PFOA to CO2 resulted in an increase of viscosity from 0.2 mPas
tions. The copolymer used in this study had a composition of 29
to 0.6 mPas. Although the level of viscosity enhancement was
mol% styrene—71 mol% fluoroacrylate, a number-average mo-
promising, the shear rate of the measurement was not provided,
lecular weight of 540,000, and a polydispersity index of 1.63. The
and the concentration of PFOA was high.1,10
copolymer was sufficiently soluble in CO2 at typical reservoir
Enick and coworkers designed, synthesized, and evaluated sev-
flooding conditions to induce a significant increase in viscosity.
eral CO2-soluble thickeners.11–13 Many of these thickeners were
Falling cylinder viscometry measurements at 298 K demonstrated
similar in structure to oil thickeners, but were modified to incor-
that the viscosity enhancement of CO2 increased with increasing
copolymer concentration, decreasing shear rate, and decreasing porate fluorocarbon chains (rather than hydrocarbon chains) in
temperature. Mobility measurements of the fluoroacrylate-styrene their structure. The fluoroalkyl, fluoroether, and fluoracrylate
copolymer-CO2 solutions flowing through 80 to 200 md Berea functional groups in these compounds imparted CO2-philicity to
sandstone cores at superficial velocities of 0.00035 to 0.028 cm/s the candidate thickening agent, thereby facilitating its dissolution
(1–80 ft/D) at 298 K also indicated that the copolymer was an in dense CO2. The polar, ionic, hydrogen-bonding, or carbon di-
effective thickener. At a superficial velocity of 0.00035 m/s (1ft/ oxide-insoluble groups in the thickener enhanced intermolecular
D), the addition of 1.5 wt% copolymer increased the CO2 viscosity associations with similar groups in neighboring molecules. These
by a factor of 19 relative to neat CO2. Smaller increases in vis- intermolecular associations led to the formation of macromolecu-
cosity occurred at lower copolymer concentrations and higher ve- lar structures in solution that were capable of inducing viscosity
locities. These results demonstrate that it is possible to design a increases or gelation when present in dilute concentration. Thick-
compound that can enhance the viscosity of dense carbon dioxide. ening candidates included semifluorinated trialkyltin fluorides,11
Fluoropolymers are characterized by environmental persistence, telechelic ionomers,11 small fluorinated hydrogen-bonding com-
high cost, and unavailability in bulk quantities, however. Therefore pounds, 12 fluoroacrylate homopolymers, 13 fluoroacrylate-
nonfluorous, inexpensive thickeners are currently being designed. hydrocarbon copolymers,13 and partially sulfonated fluoroacry-
late-styrene copolymers.13 Although each of these compounds ex-
hibited CO2 solubility of several weight percent, most induced
Introduction modest viscosity increases. Modest increases in carbon dioxide
The mobility ratio of carbon dioxide floods is unfavorable because viscosity at 298 K were detected by falling cylinder viscometry for
the viscosity of CO2 is low relative to the oil being displaced. A most of these compounds. The poly(fluoroacrylate-styrene) co-
more favorable mobility ratio would result in an increased amount polymer (Fig. 1) and its slightly sulfonated analogs induced tre-
of oil recovery prior to CO2 breakthrough and increased rate of oil mendous increases in viscosity when present in dilute concentra-
recovery thereafter. Water-alternating-gas injection is typically tion, however.13 It was postulated that the fluoroacrylate compo-
used to improve the mobility ratio by decreasing the relative per- nent induced high CO2 solubility while the pendent phenyl groups
meability of the carbon dioxide. Alternatively, the mobility ratio associated with the styrene interacted with the phenyl group of a
could be improved if the viscosity of carbon dioxide was in- neighboring copolymer molecule via ␲-␲ stacking.14 Stacking of
creased. The use of “thickened” CO2 would also eliminate the need aromatic rings is attributed to delocalization of electrons on the
for the co-injection of water. carbon atoms of the ring and slight residual positive charge on the
The identification of a compound capable of dramatically in- hydrogen atoms. The hydrogen atoms are attracted to the more
creasing the viscosity of carbon dioxide when present in dilute electron-rich carbon atoms to give a T-shaped arrangement (the
solution has been an elusive goal. The primary obstacle that has aromatic rings stack normally to one another, forming a “T” shape
hindered the efforts of numerous investigators has been the ex- when viewed on edge).14 This hypothesis was supported by the
tremely low CO2-solubility of candidate thickeners.1 Most of the observation that many fluoroacrylate-based copolymers dissolved
pioneering efforts in this area were conducted by Heller and co- in CO2, but only the copolymer with the pendent aromatic ring
workers1–6 in their evaluation of tri-alkyl tin fluorides, polymers, thickened the CO2.13 Therefore, the thickening of the CO2 was
telechelic ionomers, and 12-hydroxystearic acid. Significant con- achieved by the intermolecular associations of the dissolved co-
tributions were also made by Irani and coworkers, who success- polymer molecules, which established an associative, viscosity-
fully thickened CO2 using silicone-based polymers that required enhancing, macromolecular network to form in dense carbon di-
the addition of an organic solvent.7 Irani and coworkers later pro- oxide. The optimal composition of the copolymer was 71 mol%
posed that polymers that were designed to exhibit low solubility fluoroacrylate and 29 mol% styrene. Compositions with greater
parameters may be effective CO2 thickeners in the presence of amounts of styrene were more difficult to dissolve and yielded
small amounts of an organic liquid cosolvent.8,9 smaller viscosity increases. This was probably caused by an in-
creased number of intramolecular interactions of the aromatic
groups that prevented the polymer coil from expanding and asso-
ciating with other copolymer molecules.13 Copolymer composi-
Copyright © 2003 Society of Petroleum Engineers
tions with less than 29% styrene are also effective because they are
This paper (SPE 84949) was revised for publication from paper SPE 71497, first presented even more soluble in CO2, but they are somewhat less effective as
at the 2001 SPE Annual Technical Conference and Exhibition, New Orleans, 30 Septem-
ber–3 October. Original manuscript received for review 25 January 2002. Revised manu-
thickeners and require higher concentrations in CO2 to attain the
script received 22 January 2003. Manuscript peer approved 28 January 2003. desired level of viscosity enhancement.13

June 2003 SPE Journal 85


Fig. 1—Synthesis of the fluoroacrylate-styrene random copolymer.

Objective were employed to determine copolymer solubility.11–13 Specified


The use of an associative thickener to reduce the mobility of car- amounts of the copolymer and CO2 were introduced to the sample
bon dioxide flowing through porous media has not previously been volume. The mixture was then heated to the specified temperature
determined. Therefore, the objective of the current work was to and compressed to 45 MPa. Mixing was achieved by rocking the
determine whether the fluoroacrylate-styrene copolymer would be cell until a single, transparent phase was obtained. The sample
sufficiently soluble in CO2 in the 298 to 373 K temperature range volume was then slowly expanded by withdrawing the overburden
to decrease the mobility of carbon dioxide flowing through a sand- fluid at a constant rate until the copolymer began to precipitate.
stone core. Viscosity was initially estimated with a falling cylinder This pressure corresponded to the cloud-point pressure of the mix-
viscometer, and aluminum cylinders of varying diameter were ture of known overall composition. This procedure was repeated
used to define the non-Newtonian characteristics of the thickened for copolymer concentrations up to 5 wt%.
CO2 over a wide range of shear rates. The mobility of thickened
CO2 flowing through Berea sandstone cores at superficial veloci- Falling Cylinder Viscometry. The simplest test for estimating
ties of 0.00035 to 0.028 cm/s (1–80 ft/D) was then determined. high-pressure fluid viscosity is falling object viscometry.2,11,13,15
Single-phase solutions of the copolymer in CO2 were prepared at
Experimental 298 K and 34 MPa. An aluminum cylinder was placed within the
Copolymer Synthesis. Styrene [Aldrich, 99+%] was purified un- sample volume of hollow quartz tube (Fig. 3). Rapid inversion of
der vacuum distillation. 3,3,4,4,5,5,6,6,7,7,8,8,9,9,10,10,10- the high-pressure cell enabled the fall of the aluminum cylinder
heptadecafluorodecyl acrylate [Aldrich, 97%] was used after the through the thickened CO2 to be observed (Fig. 3). Cylinders with
inhibitor, 100 ppm monomethyl ether hydroquinone, was removed an outer diameter slightly less than the inside diameter of the
by adding the monomer dropwise to an inhibitor removal column. quartz tube were used (Table 1), thereby reducing the terminal
Bulk, free radical polymerization, with AIBN as initiator, was used velocity of the falling cylinder and minimizing the acceleration
to obtain a random copolymer product (Fig. 1). length required for the cylinder to attain its terminal velocity. The
In a typical synthesis of the 71% fluoroacrylate, 29% styrene pressure equilibration across the walls of the quartz tube elimi-
copolymer, 0.82 g (0.007873 mol) styrene, 10.00 g (0.01929 mol) nated the possibility of any change in the dimensions of the tube
fluoroacrylate, and 8.93×10−3 g (5.44×10−5 mol) of AIBN were under high-pressure conditions.
The viscosity and shear rate can be determined for Newtonian
added to a 50-ml glass ampule under an inert N2 atmosphere. The
fluids given the geometry of the system, the density of the fluid
ampule was sealed and placed in an oil bath at 338 K for 12 hours
and falling cylinder, and the terminal velocity.15 Multiple cylin-
to promote polymerization. The reaction products were then dis-
solved in 100 g 1,1,2-trichlorotrifluoroethane. The copolymer was ders must be used to characterize non-Newtonian fluids.16 The
then precipitated by washing with 300 g methanol, while the un- value of the calculated viscometer constant differed from the value
reacted monomer remained in solution. The copolymer product, a obtained by calibration with neat carbon dioxide by 0 to 50%. The
white solid, was recovered by filtration. The copolymer was magnitude of the error increased as the diameter of the cylinder de-
washed two additional times in a similar manner and dried under creased and the terminal velocity of the falling cylinder increased.
vacuum overnight. The structure of the copolymer (Fig. 1) was The aluminum cylinder was introduced to the cell prior to the
confirmed using standard IR and NMR techniques.13 The yield addition of the copolymer and the carbon dioxide. The pressure
was then increased above the cloud point, yielding a single, trans-
was 85%. This copolymer was a random copolymer of the two
parent phase within the sample volume. The entire high-pressure
monomers with an overall proportion of styrene in the copolymer
cell was then rapidly inverted (Fig. 3), causing the aluminum cyl-
of 29%. This copolymer was neither a block copolymer (a polymer
inder to fall through the neat or thickened carbon dioxide. The time
that consists of a polyfluoroacrylate section joined to a polystyrene
required for the cylinder to fall a specified distance, 2 to 10 cm,
section) nor a physical mixture of polystyrene resin and polyfluo-
was then recorded. Each measurement was repeated five times.
roacrylate resin.
These measurements indicated that the cylinder attained its termi-
nal velocity within several millimeters.
Solubility Apparatus. The phase behavior studies of CO2-
The falling cylinder results provided a convenient measure of
copolymer mixtures at 298 to 373 K were performed using the
the change in the fluid viscosity. During the initial screening of
apparatus illustrated in Fig. 2. Phase behavior measurements were
thickeners, the solution was assumed to be Newtonian.
conducted in a high-pressure, variable-volume, windowed cell
(D.B. Robinson Cell), rated to 69 MPa and 478 K. A thick-walled dVz
hollow quartz cylinder was inserted in the high-pressure windowed ␶⳱−␮ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (1)
dr
cell. A floating piston in the hollow cylinder isolated the sample
volume in the cell from the overburden fluid below the piston and Consider the aluminum cylinder falling through a dilute poly-
surrounding the tube (Fig. 3). Standard, nonsampling techniques mer solution that is considered to be a Newtonian fluid. The gov-

86 June 2003 SPE Journal


Fig. 2—Apparatus for phase behavior, falling cylinder viscometry, and mobility measurements.

erning expression for a falling cylinder viscometer under these when comparing neat CO2 to thickened CO2. Under these constraints
conditions is provided in Eq. 2. for a Newtonian fluid, the ratio of the viscosity of thickened CO2 to
the viscosity of neat CO2 was equal to the ratio of the terminal
共␳c − ␳f兲 velocity in neat CO2 to the terminal velocity in thickened CO2.
␮=C . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (2)
Vc
␮copolymer solution Vc,CO2
The viscometer constant, C, can be theoretically determined = . . . . . . . . . . . . . . . . . . . . . . . . . (3)
␮CO2 Vc,copolymer solution
from the viscometer geometry15,16 or experimentally determined
by calibrating the viscometer with neat CO2. The disparity be- The viscosity of the CO2-copolymer solutions was also mod-
tween these values increased with decreasing cylinder diameter eled using a power law expression, providing a more accurate
and increasing gap width (Table 1). The narrower cylinders were representation of the non-Newtonian nature of the solution. (Note
also prone to wobbling and rotation during their fall. Therefore it that if non-Newtonian fluid approaches the Newtonian behavior, n
should be emphasized that the primary utility of falling cylinder approaches unity as m approaches ␮.)

冏冏
viscometry is the rapid detection of significant increases in vis- n−1
cosity. The density of CO2 was estimated with a precise equation dVz dVz dVz
of state derived solely for carbon dioxide.17 The density of the ␶ = −m = −␩ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (4)
dr dr dr
thickened carbon dioxide was assumed to be comparable to the
density of neat carbon dioxide because the concentration of the The relative increase in the viscosity of carbon dioxide exhib-
copolymer was relatively low, and the same cylinder was used ited by the non-Newtonian copolymer solution was defined as the
following ratio18:
␩copolymer solution
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (5)
␮CO2

Flow Through Porous Media. The most accurate and relevant


viscosity measurements were obtained by displacing the high-
pressure fluid through sandstone. A core holder rated to 100 MPa

Fig. 3—Falling cylinder viscometry.

June 2003 SPE Journal 87


(Temco, Tulsa) was used to retain the 15.24-cm long, 2.54-cm Solubility Results. The copolymer is soluble in dense carbon di-
diameter, clean, homogeneous Berea sandstone core [80 to 200 oxide at concentrations up to 5 wt% at 298, 323, 348, and 373 K
md, American Stone, ∼13% porosity]. The core sample was held (Fig. 4). The cloud-point curve shifted toward higher pressure with
within a rubber sleeve, which had two pressure taps that permitted increasing concentration and temperature. The cloud-point pres-
pressure measurements to be made 5.08 and 10.16 cm from the sure at any given temperature must be contrasted with the mini-
inlet face of the core. Neat CO2 and thickened CO2 were prepared mum miscibility pressure (MMP) at the same temperature to es-
in the Robinson cell at 298 K and 20 MPa to ensure that a single tablish the solubility of the copolymer at reservoir conditions. For
phase of copolymer-CO2 solution at 298 K was attained. The radial example, consider a carbon dioxide-miscible project with a mini-
overburden pressure of silicone oil was maintained at 23.5 MPa mum miscibility pressure of 18 MPa (2616 psia). Fig. 4 indicates
with a positive displacement pump (Ruska, Houston). A dual posi- that at a temperature of 298 K and a pressure of 18 MPa, roughly
tive-displacement proportioning pump (Ruska) was used to ensure 3 wt% of the copolymer could be dissolved in CO2, more than
that the high-pressure fluid (CO2 or thickened CO2) was displaced enough to significantly thicken the CO2. At a reservoir temperature
into the core at the same rate that the volume for the effluent fluid of 323 K and a pressure of 18 MPa, about 0.5 wt% of the copoly-
from the core was expanded. The use of a dual-proportioning mer could be dissolved in CO2—roughly the lowest amount of
pump eliminated the need for a backpressure regulator. The entire copolymer needed to make a significant (e.g., doubling) increase in
system was charged to the desired pressure, and the high-pressure CO2 viscosity. At a pressure of 18 MPa and 348 or 373 K, the
fluid was then displaced from the Robinson cell into the sandstone solubility of the copolymer in CO2 would be much less than 0.25
face at the same volumetric rate that the effluent solution was wt%, insufficient to induce an appreciable viscosity change. If
being withdrawn from the porous media. Therefore the experi- greater copolymer solubility is required at the reservoir conditions
ments were easy to control, monitor, and reproduce, despite the than is possible with the 79%fluoroacrylate-21%styrene copoly-
very small pressure gradients across the high-pressure core. High- mer with a number average molecular weight of 540,000, then a
total-pressure, low-pressure-drop differential pressure transducers copolymer of comparable molecular weight with a slightly greater
(Validyne Engineering Co., DP303 transduce, Northridge, Califor- fluoroacrylate composition and/or a lower molecular weight would
nia) were used to monitor the pressure drop along each third of the be required,13 thereby making the copolymer more CO2-soluble.
core. Equivalent pressure drops along each section indicated that Yet this modification would reduce the thickening capability of the
the copolymer was flowing uniformly through the core rather than copolymer, and a slightly greater concentration of the copolymer
being retained at the inlet face. would be required to attain a desired level of viscosity. Future
Each core was initially saturated with neat carbon dioxide. studies are directed at the design of inexpensive, environmentally
Carbon dioxide was then displaced through the core, followed by benign, nonfluorous thickeners, however, rather than further opti-
80 to 90 cm3 of thickened carbon dioxide. Finally, neat CO2 was mization of the fluorinated copolymer composition and molecu-
again introduced to the core to determine if any permeability re- lar weight.
duction had occurred. Fluid viscosity was determined by rearrang-
ing Darcy’s Law to solve for viscosity. Falling Cylinder Viscometer Results. The falling cylinder results
for CO2-rich solutions at 298 K and 34 MPa containing 0.2 to 5.0
⌬P K wt% copolymer are presented in Fig. 5. Experimental data are
␮= . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (6) provided along with curves that correspond to the power law cor-
L v
relation obtained using the parameters listed in Table 3. The plots
The pressure drop measurement associated with the flow of of ␩solution/␮CO2 as a function of shear rate illustrate that the poly-
neat carbon dioxide was used to determine the core permeability. mer solutions are shear-thinning in nature and approach Newto-
Using this permeability value, the pressure drop associated with nian behavior with diminishing copolymer concentration. No dis-
the flow of the thickened carbon dioxide was used to determine its cernible increase in viscosity was detected at copolymer concen-
viscosity. Because the same core was used for the experiments trations less than 0.2 wt%. This falling cylinder viscometer was not
with neat CO2 and thickened CO2, the ratio of the viscosity of suited for the study of dilute copolymer solutions at the very low
thickened CO2 to neat CO2 was estimated as the ratio of the shear rates associated with creeping flow through porous media
respective pressure drops along the length of the core. (1-10s−1). Still, the results clearly indicate that substantial in-
creases in viscosity can be realized at extremely low shear rates
␮copolymer solution 共⌬P Ⲑ L兲copolymer solution with the fluoroacrylate-styrene copolymer present at concentra-
= . . . . . . . . . . . . . . . . . . . . . (7) tions of at least 0.2 wt%.
␮CO2 共⌬P Ⲑ L兲CO2

Results and Discussion


Polymer Characterization. The number-, weight-, and Z-average
molecular weights, Mn, Mw, and Mz, respectively, were determined
using gel permeation chromatography (American Polymer Stan-
dards Corporation, Mentor, Ohio). These results are presented in
Table 2, along with the polydispersity index (PDI), composition,
and melting point of copolymer. The PDI is the ratio of Mw to Mn.

Fig. 4—Effect of temperature on the solubility of the copolymer


in carbon dioxide.

88 June 2003 SPE Journal


centration for a constant superficial velocity (Fig. 7). Copolymer
concentrations as low as 0.25 wt% were capable of thickening
CO2. Higher copolymer concentrations at a constant superficial
velocity led to remarkable increases in viscosity at low superficial
velocities. The 1.0 and 1.5 wt% copolymer solutions were 8 and 19
times more viscous than neat CO2, respectively, at a velocity of
0.00035 cm/s (1 ft/D). Because of the shear-thinning nature of the
Fig. 5—Relative viscosity increase vs. shear rate at 298 K, 34 copolymer, increases in superficial velocity led to diminished
MPa, power law model, falling cylinder viscometry. thickening. For example, the addition of 1.5 wt% copolymer in-
creased the CO2 viscosity by a factor of only 2 at a superficial
velocity of 0.028 cm/s (80 ft/D). The permeability of the core to
Effect of Temperature on Copolymer Thickening Capability. neat CO2 that was injected after the copolymer solution agreed
The copolymer is capable of increasing the viscosity because of its with the original measurement of permeability to within 1%.
high molecular weight and the intermolecular associations of the Therefore, the flow of the copolymer solution through the sand-
aromatic groups.13,14 Although Fig. 5 demonstrated effective stone core did not result in any detectable permeability reduction.
thickening at 298 K, reservoir temperatures can be much higher. Copolymer adsorption and changes in the wettability of the porous
An increase in temperature may influence the strength and geom- media were not determined.
etry of the stacking of the phenyl groups, which may result in the
change of local molecular environment, thereby effecting the ␲-␲ Next-Generation Thickeners. Silicone polymers have been pre-
stacking of the aromatic groups and the thickening capability of viously used to thicken carbon dioxide flowing through porous
the copolymer. Therefore the aluminum cylinder with a diameter media, as evidenced by decreased mobility and improved recov-
of 3.157 cm was used to estimate viscosity through falling cylinder ery.7 Substantial amounts of an organic cosolvent were required,
viscometry. Fig. 6 demonstrates that the effectiveness of the thick- however, to attain a single-phase solution.7 The fluoroacrylate-
ener at concentrations of 1.0 and 0.5 wt% decreased slightly with styrene copolymer is a direct thickener in that it does not require
increasing temperature in the 298 to 373 K range at 34 MPa. a cosolvent. Nonetheless, this copolymer is a highly fluorinated
copolymer that is expensive, unavailable in large quantities, and
Measurement of Thickened CO2 Mobility in Sandstone. The biologically and environmentally persistent. Rather than continu-
pressure drops along the core were on the order of 100 to 100,000 ing to modify the structure, composition, or molecular weight of
Pa, depending on the superficial velocity and the copolymer con- this fluoroacrylate-styrene copolymer, a new generation of CO2
centration. This pressure drop was small relative to the pressure at thickeners is being designed. Future studies will focus on inex-
the inlet face of the sandstone, which was maintained at 20 MPa. pensive, biodegradable, nonfluorous direct thickeners that can be
The pressure drops along each third of the core varied by no more produced in large amounts and do not require an organic cosolvent
than 5% for every mobility measurement. These variations were to achieve dissolution. The same strategy used in the development
random, with no section of the core (particularly the inlet section) of the fluorinated thickener will be employed in the development
exhibiting a consistently greater pressure drop than the other sec- of the nonfluorous thickener.
tions. Therefore, no evidence of copolymer retention at the inlet First, novel copolymers composed of carbon, hydrogen, oxy-
face of the core was detected. Pressure drops of the neat and gen, and nitrogen will be designed or selected that exhibit remark-
thickened CO2 flowing at the same superficial velocity were used
to estimate the increase in CO2 viscosity. This viscosity increase at
298 K and 20 MPa is presented as a function of copolymer con-

Fig. 6—Temperature effect on copolymer-CO2 solution relative Fig. 7—Relative viscosity increase vs. concentration at 298 K,
viscosity increase, 34 MPa, falling cylinder viscometry. 20 MPa, flow through ∼100 md Berea sandstone.

June 2003 SPE Journal 89


ably high CO2 solubility. Polymers containing carbonate and ether ␳f ⳱ fluid density, m/L3, g/cm3
groups, 19 sugar acetates, 20,21 acetate, 22 carbonyls, 19,22 and ␶ ⳱ shear stress, m/Lt2, N/m2
ethers23 are particularly attractive candidates. For example, we
have recently demonstrated that poly(vinyl acetate) is the most Acknowledgments
CO2 soluble, nonfluorous, high-molecular-weight, commodity The authors would like to express their appreciation to the U.S.
polymer that can readily dissolve in CO2 over a wide temperature D.O.E. Natl. Energy Technology Center for their support of this
range.22 The pressure required for dissolution of the polymer is program through Natl. Petroleum Technology Office contracts
much greater than even the highest MMP values at reservoir tem- DOE-RA26-98BC15108 and DE-FC26-01BC15315. Aaron
perature, however. Therefore nonfluorous polymers with an even Wlaschin was supported by an NSF grant for Research Experience
greater affinity for CO2 than poly(vinyl acetate) must be designed for Undergraduates, EEC 0097664. We would also like to thank
and synthesized. Bayer, Cabot Oil, and Gas and Air Products for their financial
Second, CO2-phobic, associating groups will be incorporated support. Dr. Mark Thies of Clemson U. assisted us in the charac-
into the highly CO2-philic polymer to impart thickening capabili- terization of the copolymer.
ties. The associating group content of the copolymer thickener
composition will be adjusted to attain a reasonable balance of CO2 References
solubility and thickening capability. 1. Enick, R.M.: “A Literature Review of Attempts to Increase the Vis-
cosity of Dense Carbon Dioxide,” NETL report DE-AP26-97FT25356,
Conclusions U.S. D.O.E., Washington, DC (October 1998) www.netl.doe.gov.
This study has established that it is possible to design a polymer 2. Heller, J. et al.: “Direct Thickeners for Mobility Control of CO2
that exhibits both high carbon dioxide solubility and the ability to Floods,” SPEJ (October 1985) 679.
enhance carbon dioxide through intermolecular associations using 3. Dandge, D.K., Taylor, C., and Heller, J.P.: “Associative Organotin
a fluorinated copolymer. A 71 mol% fluoroacrylate-29 mol% sty- Polymers: 1. Symmetric Tributyltin Fluorides, Synthesis and Proper-
rene random copolymer was synthesized through bulk, free-radical ties,” J. of Polymer Science—Part A Poly. Chem. (February 1989)
polymerization. This copolymer, which had a number-average mo- 1053.
lecular weight of 540,000 and a polydispersity index of 1.63, ex- 4. Dandge, D.K. et al.: “Associative Organotin Polymers,” J. of Macro-
hibited remarkably high solubility in dense carbon dioxide because molecular Science, Chem A 26 (1989) 18, 1451.
of the high fluoroacrylate content of the copolymer. The thicken- 5. Dandge, D.K., Singh, P.K., and Heller, J.P.: “Associative Organotin
ing capability of the copolymer over the 298 to 373 K temperature Polymers in Miscible Gas Enhanced Oil Recovery,” in Oilfield Chem-
range was attributed to intermolecular stacking of the aromatic istry—Enhanced Recovery and Production Stimulation, J.K. Borchardt
rings associated with the styrene monomer. Falling cylinder vis- and T.F. Yen (eds.), ACS Symposium Series No. 396, American
cometry results were used to develop a power law model of the Chemical Society, Washington, DC (1989) 529–546.
pseudoplastic solutions of the copolymer in CO2. Pressure-drop 6. Gullapalli, P., Tsau, J.-S., and Heller, J.P.: “Gelling Behavior of 12-
measurements of high-pressure solutions flowing through Berea Hydroxystearic Acid in Organic Fluids and Dense CO2,” paper SPE
sandstone demonstrated that the copolymer could effectively re- 28979 presented at the 1995 SPE International Symposium on Oilfield
duce the mobility of carbon dioxide. For example, the apparent Chemistry, San Antonio, Texas, 14–17 February.
viscosity of CO2 increased by a factor of 8-6 at superficial veloc- 7. Bae, J.H. and Irani, C.A.: “A Laboratory Investigation of Viscosified
ities of 0.00035 to 0.0035 cm/s (1 to 10 ft/D) for a 1 wt% solution CO2 Process,” SPE Advanced Technology Series (April 1993) 166.
of the copolymer in CO2. Higher velocities and lower copolymer 8. Irani, C.A., Harris, T.V., and Pretzer, W.R.: “Polymer Containing Pen-
concentrations resulted in smaller mobility reductions. Although dant Vinyl Ether Groups Useful in EOR Using Carbon Dioxide Flood-
this copolymer successfully thickened CO2, fluorinated polymers ing,” U.S. Patent No. 4,945,990 (1990).
are expensive, unavailable in bulk quantities, and environmentally 9. Irani, C.A., Harris, T.V., and Pretzer, W.R.: “Polymer Containing Pen-
persistent. Therefore, we are currently investigating nonfluorous dant Tertiary Alkyl Amine Groups Useful in EOR Using Carbon Di-
thickeners, which are composed solely of carbon, hydrogen, oxy- oxide Flooding,” U.S. Patent No. 4,945,989 (1990).
gen, and nitrogen. Acetate-rich polymers appear to be particularly 10. McClain, J.B. et al.: “Characterization of Polymers and Amphiphiles in
promising candidates. Supercritical CO2 using Small Angle Neutron Scattering and Viscom-
etry,” 1996 Spring Meeting of the ACS, Division of Polymeric Mate-
Nomenclature rials Science and Engineering, New Orleans (1996).
C ⳱ falling cylinder viscometer constant 11. Shi, C. et al.: “Semi-Fluorinated Trialkyltin Fluorides and Fluorinated
K ⳱ permeability, L2, md Telechelic Ionomers as Viscosity-Enhancing Agents for Carbon Diox-
L ⳱ length, L, m ide,” Industrial & Engineering Chemistry Research (2001) 40, 908.
12. Shi, C. et al.: “The Gelation of CO2: A Sustainable Route to the
M ⳱ power law model constant
Formation of Microcellular Materials,” Science (19 November 1999)
MMP ⳱ minimum miscibility pressure, m/Lt2, Pa 1540.
Mn ⳱ number-average molecular weight, m/mole, g/mol 13. Huang, Z. et al.: “Enhancement of the Viscosity of Carbon Dioxide
Mw ⳱ weight-average molecular weight, m/mole, g/mol Using Styrene/Fluoroacrylate Copolymers,” Macromolecules (2000)
Mz ⳱ Z-average molecular weight, m/mole, g/mol 33, 5437.
n ⳱ power law model constant 14. Weber, E.: Topics in Current Chemistry: Design of Organic Solids,
P ⳱ pressure, m/Lt2, Pa Springer-Verlag, Hamburg, Germany (1998).
PDI ⳱ polydispersity index 15. Iezzi, A.: MS thesis, U. of Pittsburgh (1989).
r ⳱ r coordinate, L, m 16. Ashare, E. and Bird, R.B.: “Falling Cylinder Viscometer for non-
Tm ⳱ melting point, T, K Newtonian Fluids,” AIChEJ (1965) 11, 910.
v ⳱ superficial velocity, L/t, m/s 17. Span, R. and Wagner, W.: “A New Equation of State for Carbon
Vc ⳱ falling cylinder terminal velocity, L/t, m/s Dioxide Covering the Fluid Region from the Triple-Point Temperature
to 1100 K at Pressures up to 800 MPa,” J. Phys. Chem. Ref. Data
Vz ⳱ velocity fraction in z direction, L/t, m/s
(1996) 25, 1509.
x ⳱ mole fraction styrene in copolymer
18. Enick, R., Beckman, E., and Hamilton, A.: “Novel CO2-Thickeners for
y ⳱ mole fraction fluoracrylate in copolymer Improved Mobility Control,” Quarterly Report, April 1–June 30, 2000,
z ⳱ z coordinate, L, m DOE-RA26-98BC15018, Washington, DC (July 2000).
␩ ⳱ non-Newtonian fluid viscosity, m/Lt, Pa-sec 19. Sarbu, T., Styranec, T., and Beckman E.J.: “Non-fluorous Polymers
␮ ⳱ Newtonian fluid viscosity, m/Lt, Pa-sec with Very High Solubility in Supercritical CO2 Down to Low Pres-
␳c ⳱ cylinder density, m/L3, g/cm3 sures,” Nature (May 2000) 165.

90 June 2003 SPE Journal


20. Raveendran, P. and Wallen, S.L.: “Sugar Acetates as Novel, Renew- direct carbon dioxide thickeners for enhanced oil recovery,
able CO2-Philes,” J. Am. Chem. Soc. (2002) 124, 7274. funded by the U.S. D.O.E. Xu holds BS and MS degrees in
21. Potluri, V. et al.: “Per-acetylated Sugar Derivatives Show High Solu- chemical engineering from the East China U. of Science and
Technology. Aaron Wlaschin is a PhD candidate at the U. of
bility in Liquid and Supercritical Carbon Dioxide,” Organic Letters
Minnesota in the Chemical Engineering and Materials Science
(2002) 4, 2333. Dept., where he is advised by Dr. Friedrich Srienc. As an under-
22. Shen, Z. et al.: “CO2-Solubility of Oligomers and Polymers that Con- graduate he participated in an NSF-sponsored Research Ex-
tain the Carbonyl Group,” Polymer (2003) 44, 1491. periences for Undergraduates program at the U. of Pittsburgh
23. O’Neill, M.L. et al.: “Solubility of Homopolymers and Copolymers in in Bob Enick’s lab. In 2002 he received an NIH Biotechnology
Carbon Dioxide,” Ind. Eng. Chem. Res. (1998) 37, 3067. training fellowship. His thesis research focuses on using anaero-
bic fermentation and elementary mode modeling to optimize
the synthesis of biodegradable polymers. Wlaschin holds a BS
degree in chemical engineering from the U. of Nebraska. Bob
Enick is a professor in the Dept. of Chemical and Petroleum
SI Metric Conversion Factors
Engineering at the U. of Pittsburgh. e-mail: enick@engrng.
cp × 1.0* E–03 ⳱ Pa⭈s pitt.edu. He joined the university immediately after his gradu-
ft × 3.048* E–01 ⳱ m ation in 1985, and he currently holds the James T. MacLeod
Professorship in the School of Engineering. He is a long-standing
mL × 1.0* E+00 ⳱ cm3
member of SPE, and his current research focuses on the de-
psi × 6.894 757 E+00 ⳱ kPa sign, synthesis, and evaluation of hydrocarbon-based, carbon
*Conversion factor is exact. dioxide-soluble thickeners and surfactants for improved mobil-
ity control. Enick is also an Oak Ridge Fellow at the U.S. D.O.E.
Natl. Energy Technology Center, where he works on the de-
velopment of hydrogen and carbon dioxide membranes. En-
Jianhang Xu is a PhD student in the Dept. of Chemical and ick holds BS, MS, and PhD degrees in chemical engineering, as
Petroleum Engineering at the U. of Pittsburgh. e-mail: well as an MS degree in petroleum engineering, from the U.
jianhang+@pitt.edu. He is currently working on the study of of Pittsburgh.

June 2003 SPE Journal 91

Das könnte Ihnen auch gefallen