Anda di halaman 1dari 10

Langmuir 1992,8, 838-847

Effect of Surfactant on the Solution Properties of


Hydrophobically Modified Polyacrylamide
S.Biggs, J. Selb, and F. Candau*
Institut Charles Sadron (CRM-EAHP), 6 Rue Boussingault, 67083 Strasbourg Cedex, France
Received September 26,1991. In Final Form: December 24,1991

The interactions, in aqueous solution, of sodium dodecyl sulfate (SDS)with copolymers of acrylamide
(AM) and N-(4-ethylphenyl)acrylamide (e$AM; [e$AMI/([AMl + [e$AMl) between 1and 1.3 mol 5%)
have been examined in the semidilute concentration range by rheology, fluorescence, and conductometry.
Data are discussed with regard to changes in the hydrophobe content, the molecular weight, and the
microstructure of the copolymers. The hydrophobic groups of such copolymer chains, in pure water
solution, are associated by either inter- or intrachain liaisons. The presence of interchain liaisons leads
to apparent viscositiesmuch greater than those seen for corresponding polyacrylamides. Addition of SDS
to these aqueous copolymersolutions was seen to causedramatic increasesin the viscosityat concentrations
below that of the critical micelle concentration (cmc). Examination of the viscosity both as a function
of shear rate and as a function of time (fixed shear rate) showed these solutions to have shear thickening,
rheopexic, and thixotropic behavior. This complex behavior is explained in terms of the balance between
inter- and intrachain liaisons and their effects on chain dimensions. Fluorescence and conductometric
data have shown that the SDS associates with the hydrophobicregions of the copolymers in a noncooperative
continuous binding process.

Introduction investigated:-" there are relatively few communications


on the interaction of hydrophobically associating co-
The interaction of hydrophobically associating water- polymers and surfactants. In contrast, the interaction of
soluble copolymers with small molecule surfactants, for nonionic water-aoluble homopolymers with surfactants
example sodium dodecyl sulfate (SDS), is of fundamental in solution has been extensively investigated. An excellent
importanke for many industrial applications such as comprehensive review12 of such systems has been previ-
tertiary oil recovery and latex paint technology.' This ously published and therefore only a brief review of the
interest stems from the need to accurately control the major points will be presented here.
solution viscosity under a variety of shear conditions. As The interaction of poly(ethy1eneoxide) (PEO) with SDS
a consequence,nonionic water-soluble copolymersof acryl- has been of particular interest to many research groups.
amide prepared with low amounts of a hydrophobic Studies of this interaction have been performed with a
comonomer (<1mol %) have recently received a great wide range of techniques including vis~osity,'~ surface
deal of attention in the specific Interchain tension,14dialysis equilibrium,15NMR,16 ESR,17neutron
associations of the hydrophobe units, which are dispersed scattering,'& and f l u ~ r e s ~ e n As
~ e a. ~result
~ ~ ~of~these
along the polyacrylamide backbone, lead to solutions of studies it has been concluded that PEO interacts with
high viscosity at low concentrations and low shear rates. SDS by inducing micellization of the surfactant onto the
In general, these solutions are shear thinning, as the shear polymer chains at concentrations below that of the cor-
rate is increased the interactions are disrupted and the responding critical micelle concentration (cmc) in pure
viscosity decreases. If the shear is halted, the associations water. The existence of two critical surfactant concen-
between the polymer chains re-form and the solution trations, viz. C1 and CZ,in the system SDS/PEO was first
viscosity returns to its original value. While the properties reported by Jones.14 C1 was defined as the surfactant
of these copolymers in pure solution have been thoroughly concentration at which interaction with the polymer first
occurs while C2 is that concentration at which all the
* To whom correspondence should be addressed. polymer sites available for interaction are saturated and
(1) (a) Polymers in Aqueous Media; Glass, J. E., Ed.; Advances in any further surfactant will form classic micelles free in
Chemistry Series 223;American Chemical Society: Washington, DC, 1989.
(b) Evani, S.; Rose, G. D. Polym. Mater. Sci. Eng. 1987,57,477. solution. C1 and C2 are always located below and above
(2) (a) Peer, W. J. Polym. Mater. Sci. Eng. 1987,57,492. (b) Peer, W. the corresponding cmc, respectively. The micelles which
J. in ref la, p 381. are formed at C1 have a smaller aggregation number, a
(3) (a) Valint, P. L., Jr.; Bock, J.;Schulz, D. N.Polym. Mater. Sci. Eng.
1987,57, 482. (b) Valint, P. L., Jr.; Bock, J.; Schulz, D. N. in ref la, p higher degree of charge delocalization, and a more open
399. structure19 than those of a regular SDS micelle. The
(4) (a) Bock, J.; Siano, D. B.; Valint, P. L., Jr.; Pace, S. J. Polym. polymer chain is believed to envelope the micelles with
Mater. Sci. Eng. 1987.57, 487. (b) Bock, J.; Siano, D. B.; Valint, P. L., the hydrophilic regions of the chain associated with the
Jr.; Pace, S. J.<n ref l a , p 411.
(5) (a) Siano, D. B.; Bock, J.;Myer, P.; Valint, P. L., Jr. Polym. Mater. ionic head groups and the hydrophobic regions protecting
Sci. Ene. 1987. 57. 609. (b) Siano, D. B.: Bock, J.: Mver, P.: Valint, P.
L., Jr. ref l a , p 425. (12) Goddard, E. D. Colloids Surf. 1986,19, 255.
(6) Dowling, K. C.; Thomas, J. K. Macromolecules 1990, 23, 1059. (13)Francois, J.; Dayantis, J.; Sabbadin, J. Eur. Polym. J. 1985, 21,
(7) Evani, S. U S . Patent 4 432 881, 1984. 165.
(8) Mc Cormick, C. L.; Nonaka, T.; Johnson, C. B. Polymer 1988,29, (14) Jones, M. N. J. Colloid Interface Sci. 1967, 23, 36.
731. (15) Shirahama, K. Colloid Polym. Sci. 1974, 252, 978.
(9) Schulz, D. N.; Kaladis, J. J.; Maurer, J. J.; Bock, J.; Pace, S. J.; (16) Cabane, B. J. Phys. Chem. 1977,81, 1639.
Schulz, W. W. Polymer 1987, 28, 2110. (17) Shirahama, K.; Tohdo, M.; Murahashi, M. J. Colloid Interface
(IO) Hill, A.; Candau, F.; Selb, J. Prog. Colloid Polym. Sci. 1991, 84, Sci. 1982, 86, 282.
61. (18)Cabane, B.; Duplessix, R. J.Phys. (Paris) 1987, 48, 651.
(11)Flynn, C. E.; Goodwin, J. W. In Polymers as Rheology Modifiers; (19) Zana, R.; Lianos, P.; Lang, J. J.Phys. Chem. 1985, 89, 41.
Schulz, D. N . , Glass, J. E., Eds.; ACS Symposium Series 462; American (20) Turro, N. J.; Baretz, B. H.; Kuo, P.-L. Macromolecules 1984,17,
Chemical Society: Washington, DC, 1991; Chapter 11, p 190. 1321.

0 1992 American Chemical Society


Hydrophobically Modified Polyacrylamide Langmuir, Vol. 8, No. 3, 1992 839

the exposed hydrophobic areas of the micelle. Such a Table I. Physical Properties of the Copolymers
~

conformation which limits the contact of the hydrophobic copolymer % monomers hydrophobe) mol wt,'
regions of the micelles with water is obviouslyenergetically sample code conversiona mol % Mw (Xl0-9
favorable and therefore the micelles form at a lower sur- M 51 44.8 1.30 3.13
factant concentration, i.e. C1. These findings led Shira- M 52 52.6 1.25 3.12
hama et to propose the so-called "necklacen model of M 53 62.7 1.12 2.90
the complex in which the beads are the SDS clusters and M 54 73.5 1.09 2.67
the string is the polymer chain. In such a model the chain M 55 83.0 1.03 2.60
is elongated to reduce the electrostatic repulsion between M 56 87.8 1.03 2.57
the clusters, a proposal which is supported by viscosi- a Monomers conversion as a percentage of total initial monomers
metric data which show polyelectrolyte behavior for the *
feed in polymerization. Mol % = [hydrophobe]/([acrylamide] +
~omp1exes.l~ There is no evidence at present to suggest [hydrophobe]). Determined from light scattering measurements.
aggregates comprising more than one chain for such ho-
mopolymer/surfactant systems in the range of concen- ified hydroxyethyl cellulose copolymer in SDS solution,
trations studied. Other homopolymers investigated for both increases and decreases in solution viscosity were
their interaction behavior in aqueous surfactant solution observed as a function of increasing SDS concentration
include poly(vinylpyrro1idone) poly(viny1 acetate) ,24
,22923
below the cmc. The increase in viscosity observed at low
and p~lyacrylamide.~~ In the case of polyacrylamide it SDS concentrations was attributed to an increase in chain
has been shown that there is no interaction with SDS in cross-linking by the formation of mixed micelles. As the
aqueous solution. surfactant cmc is approached, however, there is sufficient
Studies of the solution properties of hydrophobically surfactant to solubilize each hydrophobe separately and
modified copolymers with surfactant are much less ex- so the viscosity falls. Similar behavior, in SDS solution,
tensive. For such systems it is necessary to consider the has also been reported both for hydrophobically modified
microstructure of the copolymer, the presence of charge ethoxylated urethane (HEUR) copolymers33and for a hy-
either in the chain backbone or on the pendant hydro- drophobically modified poly(sodium acrylate) in SDS
phobic groups, and the size of these pendant groups. I t is solution.34
also important to consider the chemical nature (anionic, In this paper we report the results of investigations into
cationic, or nonionic), size, and structure of the surfac- the effects of SDS on the semidilute solution properties
tant.26127The interplay of two or more of these factors can of copolymers of acrylamide and N- (4-ethylpheny1)acryl-
lead to widely differing properties. For instance, addition amide at surfactant concentrations ranging from below to
of surfactant, particularly nonionic or cationic, to an above the cmc in pure water solution. The effects of
aqueous solution of hydrophobically modified poly(s0- varying the hydrophobe content of the copolymer samples
dium acrylate) has been shown to cause a marked viscosity in this region of enhanced solution viscosity will be
enhancement2* However, results for a copolymer of acryl- discussed. Rheological behavior as a function of (i) the
amide and an N-alkylacrylamide in the presence of anionic shear rate and (ii) the shear time at constant shear rate,
surfactant suggest that the surfactant disrupts interchain in this region, is investigated. Complementary data
liaisons causing a decrease in solution viscosity.*l0 obtained by fluorescence and conductometry will also be
As outlined above, the presence of hydrophobic groups presented. In the light of these data the relative impor-
along the copolymer backbone may allow, dependent upon tance of inter- and intramolecular interactions for the
the chain flexibility,both inter- and intrachain associations copolymers as a function of shear conditions will be
which are in many ways analogous to micellar aggregates. discussed.
Small molecule surfactants, when added to such a system,
can interact with these hydrophobic groups leading to the Experimental Section
formation of mixed micelle-like These Materials. Copolymers of acrylamide andN-(4-ethylphenyl)-
mixed micelles either may promote chain cross-linking by acrylamide were prepared by free-radical copolymerization in
acting as a bridge between two polymer chains or may an aqueous micellar medium. All the copolymer samples
disrupt such linkages by solubilizing each hydrophobic examined here were prepared during a single copolymerization
group of a polymer chain into a single micelle. It has been reaction. The termination of aliquots of this reaction at different
previously reporteds32 that, for a hydrophobically mod- degrees of total monomers conversion led to copolymer samples
with differing degrees of average hydrophobe content.35 Data
(21)Shirahama, K.;Tsujii, K.; Takagi, T. J. Biochem. 1974,75, 309. for the hydrophobe content and the degree of total monomers
(22)Fishman, M. L.; Elrich, F. R. J. Phys. Chem. 1975,79,2740. conversion for these copolymers are given in Table I. The
(23)Arai, H.; Murata, M.; Shinoda, K. J. Colloid Interface Sci. 1971, copolymers chosen for study here varied in their average
37,223.
(24)Tadros, Th.F. J. Colloid Interface Sci. 1974,46,528. hydrophobic monomer content between 1 and 1.3 mol %
(25)Sabbadin,J.;Le Moigne, J.;Franpis, J. In SurfactantsinSolution; ([monomer]/ [monomers]). The distribution of the hydrophobe
Mittal, K. L., Lindman, B., Eds.; Plenum Publishing: New York, 1984; units along the acrylamide backbone has not been determined.
Vol. 2,p 1377. However, the use of a micellar reaction medium is believed to
(26)Ruckenstein, E.; Huber, G.; Hoffmann, H. Langmuir 1987,3,382. lead to copolymers with a structure in which small blocks of
(27)Nagarajan, R. J. Chem. Phys. 1989,90,1980. hydrophobe are dispersed along the polymer chain. A full
(28)Wang, K. T.; Iliopoulos, I.; Audebert, R. In Water-Soluble
Polymers. Synthesis, Solution Properties and Applications; Shalaby, discussion of the effects on the copolymer structure of using a
S.W., McCormick,C. L., Butler, G. B., Eds.; ACS Symposium Series467, micellar reaction medium is given in a previous publication from
American Chemical Society: Washington, DC, 1991;Chapter 14,p 218. this l a b ~ r a t o r y . ~The
~ molecular weights, determined from
(29)Sau, A. C.; Landoll, L. M. in ref 1, Chapter 18,p 353. classical light scattering measurements, varied in the range (2.5-
(30)Hulden, M.;Sjoblom,E. In64thACSColloidandSurfaceScience 3.2) X 108.
Symposium Abstracts; American Chemical Society: Washington, DC,
1990. Before being used in the preparation of solutions, each
(31) Dualeh, A. J.; Steiner, C. A. Macromolecules 1990,23,251. copolymer was dried for 24 h a t 50 "C to eliminate any residual
(32)Steiner, C. A. J. Appl. Polym. Sci. 1991,42, 1493. moisture. Samples were then weighed immediately and sealed
(33)(a) Lundberg, D.J.; Glass, J. E. Polym. Mater. Sci. Eng. 1989,61, prior to measurement.
533. (b) Karunasena, A.; Brown, R. G.; Glass, J. E. in ref la, p 495. (c)
Lundberg, D. J.; Ma, Z.; Alahapperuna, K.; Glass, J. E. in ref 11,Chapter
14,p 234. (35)Biggs, S.;Hill, A.; Selb, J.; Candau, F. J. Phys. Chem. 1992,96,
(34)Iliopoulos, I.; Wang, K. T.; Audebert, R. Langmuir 1991,7 , 617. 1505.
840 Langmuir, Vol. 8, No. 3, 1992 Biggs et al.
Sodiumdodecylsulfate (SDS) (TouzartandMatignon, >99.5%
pure) was used as received without further purification. The
critical micelle concentration (cmc) determined by conductivity
was 8 X 10-3 mol L-l and by fluorescence was 8.2 X mol L-l.
Pyrene was zone refined and kindly supplied by Dr. A. Mal-
liaris, NRC Democritos, Athens, Greece.
All water used here was doubly distilled from an all-glass
apparatus.
;'Ouo
W t
Methods. (1) Conductometry. Samples of the copolymer,
at appropriate concentrations, were dissolved in 20 cm3 of water
with gentle agitation over 24 h. These stock solutions were divided
into two lo-cm3 aliquots. SDS was then added to one of these
.-
VI
,x 75t
.- 50t
0

>
fractions to give a solution of polymer a t 1% (w/w) surfactant
concentration in water.
The conductivity measurements at a fixed concentration of 25 t i
copolymer, as a function of surfactant concentration, were then
performed by titrating the portion a t 1% SDS into the nonsur-
0 1
factant containing portion in small aliquots. Conductivities were 1.0 1.1 1.2 1.3 1.4
recorded using a Universal Bridge B905 (Wayne-Kerr) and a percentage hydrophobe
conductivity cell (Tacussel CM02/550). Figure 1. Variation of the apparent viscosity as a function of
(2) Fluorescence. From a filtered stock solution of water hydrophobe content (mol % ) for copolymers of acrylamide and
saturated with pyrene, copolymer/SDS solutions were prepared N-(4-ethylpheny1)acrylamide in aqueous solution at a concen-
at appropriate copolymer and surfactant concentrations. Com- tration of 1%(w/w) (shear rate, y = 1 s-l).
plete dissolution of all the components was performed with gentle
agitation for 24 h. Before measurement, the solutions were left
without agitation a further 24 h.
Steady-state fluorescence measurements of the emission
spectra for all samples were performed using a Hitachi F-4010
spectrophotometer. The excitation wavelength used was 335 nm
with a bandwidth of 1.5nm. The ratio 13/11of the intensities of
the third and the first vibronic peaks for the emission spectrum
for pyrene provides an estimate of the local environment, i.e.
hydrophobic or hydrophilic, sensed by the fluorescent probe.36
In general, for the spectra, the first and third vibronic peaks
were situated a t 374 and 385 nm, respectively.
(3) Rheological Measurements: Viscosity-time measure-
menta were performed using a Contraves Low Shear 30 instrument
interfaced to a personal computer and driven by a software
package supplied by the manufacturers.
Viscosity-shear data were obtained using a Carri-Med con-
trolled stress rheometer with a cone-plate geometry: (acrylic cone, 1
diameter 6 cm, angle lo). 1 10 100 1000
Copolymer/surfactant solutions were prepared by dissolution shear rate s-'
of the copolymers, at appropriate concentrations, in stock aqueous Figure 2. Effect of shear rate on the apparent viscosity for an
solutions of SDS. These stock solutions were pre-prepared at a aqueous solution of copolymer sample M 51 at a concentration
range of concentrations between 0.01% and 0.4% (w/w). No of 0.5% (w/w).
agitation was applied during dissolution of the copolymers which
was usually complete a t between one and two days. For certain The removal of any hydrophobic chain from aqueous
samples of high viscosity a slow tumbling procedure was applied solution, into a micellar aggregate, is always associated
to ensure complete homogeneity of the final solution.
with both a large increase in entropy and a small enthal-
Results and Discussion pic decrease.37 The entropy gain is the driving force,
therefore, for the aggregation of hydrophobic groups in
(1) Solution Properties of the Copolymers in the aqueous solution. This entropy gain originates from the
Absence of Surfactant, Apparent viscosities, measured breakdown of a highly ordered structure of the water
at a single shear rate (y = 1 s-l), as a function of average around the hydrophobe as it is removed from solution.
hydrophobe content, in the absence of SDS, are shown in Any aggregation of hydrophobic groups which leads to a
Figure 1. Details of the hydrophobe contents and mo- net decrease in the number of hydrophobe/water contacts
lecular weights of all the copolymers investigated here are will therefore be favored due to a large decrease in the free
given in Table I. Clearly, the viscosity at a constant energy.
copolymer concentration is directly related t o the average Data from fluorescence measurements of 13/11 as a
hydrophobe content of the copolymer samples. A typical function of copolymer concentration (for sample M55) in
example of the viscosity variation as a function of shear the absence of surfactant are given in Figure 3. The values
rate, for copolymer sample M51, is presented in Figure 2. of Z3/11 are seen to rise steadily, as a function of copolymer
It should be noted that all the copolymer samples concentration, indicating that at all concentrations, or at
investigated here exhibited shear thinning behavior in least those as low as 0.001 96 (w/w), hydrophobic regions
solutions free of surfactants. are found in the solutions. Asimilar increase of the solution
Before considering the results of this study in detail, it hydrophobicity, with copolymer concentration, has been
is necessary to define why hydrophobic groups interact in reported previously, for copolymers of acrylamide-alkyl-
aqueous solution and at what concentration of copolymer acrylamide, when studied by fluorescence using either
such interchain associations may occur'. pyrene" or 8-anilino-1-naphthalene~ulfonate~ as a probe.
(36) Kalyanasundaram, K.; Thomas, J. K. J.Am. Chem. SOC.1977,99, (37) Tanford, C. The Hydrophobic Effect. Formation of micelles and
2039. biological membranes, 2nd ed.; Wiley-Interscience: New York, 1980.
Hydrophobically Modified Polyacrylamide Langmuir, Vol. 8, No. 3, 1992 841

0.7 c I I for those associating copolymer systems which have been


investigated at 1 % concentration,4l the molecular weights
were of the order of lo5 and not lo6 as here.
Flynn and Goodwinll have reported the presence of a
critical hydrophobe concentration cc, for similar copoly-
mers in solution, above which rapid increases in the number
of interchain associations are observed. Increases in the
hydrophobe content of their copolymers led to a decrease
of the critical concentration and a more rapid increase in
the viscosity with concentration after cc. At any constant
IP concentration above cc, therefore, an increase in hydro-
phobe content was seen to lead to an increase in solution
viscosity as observed here.
(b) Copolymer Molecular Weight. As shownin Table
I , I, the copolymer weight average molecular weights (M,)
0.0 0.5 1.0 1.5 2.0 decrease with increasing monomers conversion. A full
copolymer concentration (wt.%) discussion of this point is given elsewhere35and will not,
therefore, be considered here. Instead, we will concentrate
Figure 3. Variation of Z3/Z1 for a pyrene fluorescent probe in on the differences in copolymer solution viscosities (at
aqueous solutions of copolymer sample M 55, as a function of constant concentration) when employing copolymers of
copolymer concentration (wt 5% 1. different molecular weights. It is known for polyacryla-
mide40that, at a solution concentration of 1 % (w/w), the
Thus, in hydrophobically associating copolymer solutions, viscosity, TI,is approximately proportional to M,.2.5 Thus,
intrachain associations may be considered to occur at any a change of 20% in the M , can lead to changes of up to
concentration. However, definition of the presence or 1.5 times in the apparent viscosity. While in the case of
absence of a minimum concentration for interchain liaisons hydrophobically modified copolymers of acrylamide the
is not at all clear. Winnik38has reported the presence of inter- and intrachain liaisons will complicate such a
interchain interactions at copolymer concentrations con- calculation, it is clear that the variation in M , of the
siderably lower than the critical overlap concentration (e*), samples (Table I) will also play an important role in the
a result supported by extensional flow data for copolymers viscosity variations observed here.
in solution.39 It is certain, therefore, that for copolymer ( c ) Hydrophobe D i s t r i b u t i o n w i t h i n t h e
concentrations equal to or greater than e*, substantial Copolymers. The distribution of hydrophobe units within
interchain liaisons may occur. In our case, the copolymer the copolymers used here has not, as yet, been absolutely
concentration was always equal to or greater than 0.5% determined. The following points are, however, known to
(w/w). If we assume that c * [ d is of the order of 1 and be true.35 Micellar copolymerization has been shown to
using a known value of the intrinsic viscosity’0 [TI = 700 give copolymers which, at low levels of monomer conver-
for a polyacrylamide of M , = 2.5 X lo6, we calculate. a sion, have a composition very rich in hydrophobe, with
value of c* = 0.14% (w/w). This is 4 times smaller than the hydrophobe being distributed along the acrylamide
our working concentration and so, even allowing for the backbone in small blocks. As the copolymerization
diminished values of the intrinsic viscosity4Jl (-2 times proceeds, the amount of hydrophobe incorporated into
smaller) seen for associating copolymers, we may expect each chain falls, with both the size of the blocks and their
interchain liaisons to occur if they are energetically number decreasing. It is important to recall that all the
favorable. copolymers used here were prepared during one copo-
The copolymer samples investigated here exhibit a lymerization reaction, therefore those copolymers recov-
relatively large viscosity change on passing from a sample ered at high levels of conversion contain a mixture of all
with an average hydrophobe content of 1.3 mol % to one the samples at lower conversion levels plus further
of 1.03 mol % (Figure 1). This change from 122 to 38 cP, copolymers at lower contents of hydrophobe. The average
a t a copolymer concentration of 1 % (w/w) in water, is due hydrophobe content of the samples falls therefore with
to a combination of factors: the hydrophobe content, the increasing conversion (Table I). This polydisperse “blocky”
copolymer molecular weight, and the distribution of this nature of the copolymers, characteristic of micellar co-
hydrophobe within the copolymer samples. The effects polymerization,contrasts with other associatingcopolymer
on the solution viscosity, in the semidilute concentration systems in which the hydrophobic moieties of the copol-
range, of altering either the hydrophobe content or the ymers are introduced by chemical modification of pre-
molecular weight have been previously i n v e ~ t i g a t e d . ~ ~ ~formed homopolymers28~29and will have therefore a
However, systematic studies of the effects of changes in relatively regular distribution of single hydrophobe units
the hydrophobe distribution have not been reported. The throughout the copolymer samples. These differences are
role of each of these three factors, in this system, will represented schematically in Figure 4.
therefore be considered separately. It is well-known that the strength, or more precisely the
(a) Hydrophobe Content. Previous reports have free energy change, of any hydrophobic interaction de-
shown that, at a fixed concentration, an increase in the pends on both the number of hydrophobic units and their
hydrophobe content causes an increase in the solution relative hydrophobicity. For example, changing the size
v i s c o ~ i t y . ~ Direct
~ ~ J ~ comparisons with this work are of the hydrophobic monomers from an aliphatic CSto an
difficult, however, since for comparable acrylamideIN- aliphatic C12 has been shown to substantially increase the
alkylacrylamidecopolymer systems no published data exist solution viscosity for otherwise identical copolymers.8
at concentrations greater than 0.4% (w/w).~ Furthermore, While the aromatic CShydrophobic monomers used here
(38) Winnik, F.M. Macromolecules 1989,22, 734.
may not have a high hydrophobicity when employed
(39) Odell, J. A.; Keller, A.; Muller, A. J. in ref la,Chapter 11, p 193.
(40) Kulicke, W.M.; Kniewske, R.; Klein, J. Prog. Polym. Sci. 1982, (41) Jenkins, R.D.Ph.D. Thesis, Lehigh University, Bethlehem PA,
8,373. 1990.
842 Langmuir, Vol. 8,No. 3, 1992 Biggs et al.

Micellar Chemical ? I 1
copolymerisation modification

I increasing
conversion
4
C
0 5
L
(d
a
chemical
reaction 2 0
0.0 0.1 0.2

.; .;
percentage SDS (w/w)
Figure 5. Apparent viscosities as a function of the SDS
concentration (wt 5% ) for copolymers of differing molar hydro-
phobe contents: M 51,O;M 52,O; M 53, 0 ;M 54, M 55,
M 56, A. (Concentration of copolymers = 0.5% (w/w);shear
rate, y = 0.5 s-l.)
measured at a single shear rate ($ = 0.5 s-l) are given. The
viscosity values shown are those observed when at equi-
librium with respect to shear time. For many of the
copolymer/surfactant systems, time-dependent shear be-
havior was seen initially (see later), although after a certain
Figure 4. Schematicrepresentation of (i) the changing average period (usually 3-5 min) a constant value was always
distribution of hydrophobic monomer within any copolymer
sample as a function of micellar copolymerization time and (ii) attained. Since we discuss these constant equilibrium
the distributionof hydrophobe units in copolymers prepared by values in this section, we use the term steady state to
chemical modification: monomer units, 0 . describe the viscosity values given.
Three regions of behavior may be identified from these
separately, if they are found as small blocks within the curves: region I occurs at very low SDS concentrations
polymer backbone their hydrophobicity can be considered (<0.01% (w/w)) where the viscosity is not seen to change
to be much greater and they will be more likely to form greatly from that of the pure copolymer solution; region
either inter- or intrachain aggregates. It was previously I1is characterized by large viscosity enhancements; region
reported that copolymers with the same molar quantity I11 is located at SDS concentrations equal to or greater
of hydrophobe but with one having a blocky structure and than the cmc where the viscosity is observed to be
the other a structure in which the monomer is dispersed approximately the same as the pure copolymer solution
as discrete units will have very different viscosities in and to show very little variation. From the measurements
solution.10 The blocky copolymer was shown to have much performed here, differences in the relative size of each
higher viscosities in aqueous solution at all the concen- region as a function of the average hydrophobe content
trations examined. are not evident. It is clear, however, that the magnitude
At low levels of conversion virtually all of the hydro- of the peak viscosity is very sensitive to the amount of
phobe is found in small blocks and will therefore prefer hydrophobe in each sample. At the highest hydrophobe
to form either inter- or intrachain liaisons protected from contents measured, the absolute maximum in the apparent
solution by the hydrophilic backbone. At higher levels of viscosity was too large to be measured with the LS 30
conversion, those copolymer chains with the large blocks rheometer. However, the points on either side were
are still present, although their concentration will be recorded giving an indication of the peakmagnitudes. Data
decreased by dilution with chains containing fewer more on the exact position of the maxima were also not
singly distributed hydrophobes. These single units are determined due to the preclusive number of solutions
more easily solubilized as discrete units, and therefore the which would be required for each copolymer sample.
proportion of interchain liaisons will fall. Thus, the The results of Figure 5 appear to be in agreement with
decrease of viscosity with increasing monomers conversion previously reported data.24-34Once again, however, direct
(Table I) or, in other words, decreasing hydrophobe content comparison is difficult because of the very different nature
(Figure 1) must also be due, in part, to the increasing of the copolymer samples. Despite this, it should be noted
proportion of copolymers with both smaller and fewer that the region of increased viscosity seen here was at
blocks at higher levels of monomers conversion. approximately the same concentrations of surfactant given
It is clear, therefore, that each of these parameters plays in these earlier reports.
an important role in the viscosity of these hydrophobi- Before discussion of the more complex rheological
cally modified copolymers in solution. The relative behavior in section 3, it is necessary to consider why the
importance of each, however, has yet to be determined. presence of surfactant causes an increase in these “steady-
(2) “Steady State” Solution Properties of the Co- state”solution viscosities. For a copolymer solution above
polymers in the Presence of Surfactant. Addition of c*, in any hydrophobically modified copolymer chain there
SDS to the aqueous copolymer solutions was seen to will be both inter- and intrachain liaisons of the hydro-
dramatically alter their rheological behavior. In Figure 5 phobe. The hydrophilic polymer chains of the copolymers
the apparent viscosities as a function of SDSconcentration, will then surround these hydrophobe regions in a highly
for copolymers of differing average hydrophobe content, coiled conformation to protect them from the solvent.
Hydrophobically Modified Polyacrylamide Langmuir, Vol. 8, No. 3, 1992 843
Obviously such a highly coiled and restricted conformation
does not represent an energetically favorable state for the
hydrophilic chains. However, the gain in entropy of
removing the hydrophobe from solution is much greater
and therefore overrides this loss of chain entropy. The
presence of surfactant micelles allows the solubilization
of the hydrophobe and its protection from the water by
the surfactant head groups. This in turn liberates the
hydrophilic chains, enabling them to rearrange into a more
energetically favorable state. As the surfactant is initially
added, at a concentration of 0.5% (w/w) copolymer, there
is already a substantial quantity of hydrophobe in the
solution. It would seem feasible therefore that the sur-
factant associates directly with those hydrophobic regions
of the copolymer solution in a noncooperative process, a
hypothesis which is supported by the conductance and
fluorescence data (see later). Similar behavior has been
proposed for a variety of systems in which hydrophobic
domains exist in solution before the addition of any
~ u r f a c t a n t . ~As
~ - the
~ ~ concentration of surfactant in-
creases, a point is reached at which there is sufficient to
effectively solubilize the hydrophobes, forming mixed mi-
celles, without the assistance of the hydrophilic chains.
This corresponds to the onset of the dramatic increases
in viscosity, region I1 in Figure 5. There is not, as yet,
sufficient surfactant to solubilize each region of hydro-
phobe separately, and so two or more regions, either of the
same chain or of others, are incorporated into the same
micelle. Obviously, as further surfactant is added and we
approach the cmc, each hydrophobe region may become
solubilized by a single micelle leading to a decrease in the
viscosity. These processes are represented schematically
in Figure 6.
The large increases in the observed viscosity for all the
samples in the presence of SDS suggest that in pure water Figure 6. Schematic representation of the interaction of the
copolymerswith SDS at three differentsurfactantconcentrations
solution the number of interchain liaisons is considerably correspondingto the three regions displayed in Figure 5. SDS,
smaller than those of an intrachain nature, a result --; hydrophobe blocks, 0.
predicted theoretically in a recent publication by Balazs
and H u . ~ ~ 1.0 I""''''\
The magnitude of the maxima (Figure 5 ) were, as in the
absence of surfactant, disproportionally larger with higher
levels of hydrophobe (cf. Figure 1). Addition of the sur-
factant facilitates interchain linking, as described above.
As discussed in section IC, the chains at low levels of
monomer conversion have a very blocky composition and
therefore an increased hydrophobicity (cf. Figure 4). Thus,
addition of surfactant will result in a large lowering of the
free energy both from the formation of mixed micelles
and also from the release of the restriction of the hydro-
philic chain motion. At higher levels of monomer con-
version there is an increased proportion of hydrophobe
units which are distributed as smaller blocks. These blocks
will not have the same tendency to form mixed micelles
due to the much smaller energy gain of forming a mixed
0.5 ' 0.00 1
I ,

0.01 0.1
I

micelle with less hydrophobe and the very small release Concentration SDS (mol/l)
of chain restriction. The dominance of larger hydrophobe Figure 7. Variation of 13/11 for a pyrene fluorescent probe in
blocks for the formation of mixed micelle interchain aqueous copolymer solutions as a function of the SDS concen-
liaisons therefore means that those samples with the hy- tration (log scale): (a) SDS, A; (b) M 54, 0;(c) M 52, 0.
drophobe distributed as blocks will have a much greater
viscosity increase due to the formation of disproportion- Data of 13/11versus SDS concentration for two copolymer
ately more linkages. samples are given in Figure 7. The curve for SDS in the
absence of copolymer is also given for reference. The
(42) Winnik, F. M.; Ringsdorf, H.; Venzmer, J. Langmuir 1991,7,905. curves of SDS/copolymer solutions were seen to rise
(43) Winnik, F. M.; Ringsdorf, H.; Venzmer, J. Langmuir 1991,7,912. gradually toward a plateau region which corresponded to
(44) Benrraou, M.; Zana, R.; Varoqui,R.;Pefferkorn, E. Submitted for those SDS concentrations of maximum viscosity. After
publication in J. Phys. Chem. this region the curves were seen to follow closely the
(45)Hu, Y.-Z.;Zhao, C.-L.; Winnik, M. A.; Sundararajan,P. R. Lang-
muir 1990, 6, 880. behavior for that of the pure SDS solution after the cmc.
(46) Balazs, A. C.; Hu, J . Y. Langmuir 1989,5, 1230 and 1253. Up to the cmc the difference in 13/11values in the presence
Biggs et al.

600
.
500 - 45.1
oomo-0%
.-x
c
UI
9f
h .d' 0

.-0>
v)

3400 - %
; I
I 4-
al C
:
2 300 -
omC
%
"
om! I
I
; E0
0

as 3 ? I a
n t
1
3
C 0% I
8 200 - 4 I
I
I 1
1 10 100 1000
100-
Pg
,f
I
I
I shear rate s-'
I
L? I Figure 9. Effect of the shear rate on the apparent viscosity for
I
8'
I CYC
an aqueous solution of copolymer sample M 52 (concentration
I , = 0.6% (w/w)) at three concentrations of SDS (w/w):(a) 0.01%;
01 '
(b) 0.02%; ( c ) 0.25%.
This continuous aggregation process contrasts with the
earlier studies on water-soluble homopolymers and sur-
factants in which a cmc was observed at some concentration
lower than the cmc of the surfactant in pure water solution.
While there is no lower critical surfactant concentration,
and absence of copolymer can be interpreted as the CI, as defined by Jones,14we might expect to see differences
increased hydrophobic character of the copolymer solu- in the upper critical concentration, CZ. However, a t the
tions. different concentrations of copolymer and hydrophobe
The absence of a sharp cmc for the copolymer/surfac- contents measured here, no measurable differences in the
tant solutions (Figure 7) supports the previous hypothesis Cp values were recorded. This is probably due, however,
of a continuous association process of the SDS with the to the relatively low overall amounts of hydrophobe in the
hydrophobicregions of the copolymers. These data agree system. In the case of homopolymers which interact with
directly with a recent report by Winnik et al.42for hy- surfactants, the number of interaction sites is very large.
drophobically modified copolymers of poly(N-isopropyl- Therefore a change in concentration of the polymer causes
acrylamide) in the presence of SDS. Their fluorescence a large and easily measurable change in the number of
results show a gradually increasing value of 1 3 / 1 1 with interactions. It was already noted that polyacrylamide
increasing surfactant concentration which was attributed does not interact with SDS,25and therefore in our case the
to a noncooperativebinding mechanism of the surfactant only possible sites of cooperationare with the hydrophobe
to the hydrophobic regions of the copolymers. groups. Since these groups are considered to be grouped
The conductance data as a function of SDS concen- along the polyacrylamide backbone in small blocks, at a
tration for two copolymer samples at a concentration of concentration of between 1 and 1.3 mol %, they will be
0.5% (w/w) are presented in Figure 8. Data for one of separated by relatively large blocks of acrylamide. Thus
these copolymer samples at a lower concentration (0.1% any electrostatic interactions of associated micelles will
(w/w))are also given as is the curve for a pure SDS solution. be very slight, contrary to the homopolymerlSDS systems
The presence of the copolymer, at a concentration of 0.5 5% where the adsorption of micelles confers a polyelectrolyte
(wlw), is seen to cause a decrease in the conductance at behavior on the polymer. So, the increased viscosities
all concentrations of SDS due to association of the sur- seen here in the presence of SDS are caused solely by
factant with the macromolecules. Below the surfactant changes in the numbers of chain-chain attractions and
cmc the difference between the conductance values with not to increased coil dimensions as a result of polyelec-
and without copolymer was seen to increase with increasing trolyte charge repulsions.
surfactant concentrations. Increasing the hydrophobe (3) Viscosity of Copolymer/Surfactant Solutions
content of the copolymer was seen to cause a slight decrease as a Function of Shear Rata and Shear Time. Viscosity
in the conductivity. At the two copolymer concentrations as a function of shear rate for a single copolymer sample,
for the same sample, the conductivity was lower at the but within the three regions of existence described above,
higher concentration. Above the cmc of the SDS each are shown in Figure 9. It can be seen that for regions I
curve was seen to be parallel with equivalent gradients, and I11 (curves a and c) the copolymer/SDS mixtures
the difference in conductance increasing from the values exhibit shear thinning with increasing shear rate, a
of SDS with increasing contents of hydrophobe. comportment seen also for the pure copolymer solutions
Further evidence for noncooperative binding is found (cf. Figure 2). For the sample located in region I1 (curve
in these results. If the surfactant is initially free in solution, 9b) the viscosity-shear rate behavior is very different, the
as discrete molecules, it may be expected that the curves sample initially exhibiting shear thickening before be-
in the presence of copolymer would correspond exactly coming shear thinning at higher shear rates. The shear
with that of the pure SDS solution at low surfactant rate range of this shear thickening region was seen to have
concentrations. A deviation of the curves would only be only a small dependence on the hydrophobe content of
seen at some concentration which correspondsto a mixed the copolymer, while the height of the maximum was again
micelle cmc. No such curves were obtained at all con- sensitive to the amount of hydrophobe in the copolymer.
centrations and hydrophobe contents examined here. A closer examination of those samples located in region
Hydrophobically Modified Polyacrylamide Langmuir, Vol. 8, No. 3, 1992 846

T
0
0.6
4

x 0.5
n
a
50.4
h
20.3
0

’5 {;I I .%0.2
3
C
co.1
2a t
20.0
- 0 2 4 6 8 10 0 2 4 6 8 1012
time (mins) time (mins)
Figure 10. Apparent viscosity as a function of shear time, at a Figure 12. Apparent viscosity as a function of shear time at a
constant shear rate (y = 0.5 s-l) for copolymer sample M51 at fixed shear rate of y = 1 s-1. Before each measure the sample
three different concentrations of SDS: (a) 0.05%; (b) 0.1%; (c) was subject to a shear rate of y = 20 s-l for 5 min duration,
0.15%. followed by a relaxation time, t (at zero shear): (a) t = 30 a; (b)
t = 9 O s ; (c) t = 150s. For reference, curve d, is the initial measure
at y = 1 s-l before any shear at y = 20 s-l. (Copolymer sample,
M 52;concentration = 0.5% (w/w); SDS concentration = 0.05%
(w/w).)

initially recorded peak viscosity and that of the plateau


value decreases. At the shortest relaxation time, 30 s, the
initial peak viscosity value is approximately that recorded
for the higher shear rate. The viscosity-time profile for
i the same sample at a shear rate of y = 1 s-l, but before
it was subjected to a higher shear rate, is also given in

4 t
i1 Figure 12 (curve d). The curves obtained both before and
after the higher shear rate appear to approach the same
equilibrium viscosity as a function of time.
The shear thinning behavior observed both for copol-
$2
$ 0
4 o
2 3 4 5 1
ymer solutions in the absence of surfactant (Figure 2) and
for those copolymer/surfactant solutions which do not show
any enhanced viscosification (Figure 9, curves a and c) is
time (mins) comparable to the classic response of a polymer solution
Figure 11. Apparent viscosity as a function of shear time a t a to increased shear stress. In this case, such behavior can
fixed shear rate of y = 5 s-l: copolymer sample, M 52; be attributed both to the disentanglement of the isolated
SDS concentration = 0.02% (w/w).
concentration = 0.5% (w/w); coils and to the breakdown of hydrophobic interactions.
A previous study of copolymersof acrylamide and N-alkyl-
I1 reveals a number of interesting rheological properties. acrylamides? which had been partially hydrolyzed (18mol
The viscosity of the samples was measured as a function %), suggested the presence of a critical copolymer con-
of time at a variety of different shear rates between y = centration (0.2% (w/w)) and hydrophobe content (1.25
0.5 s-l and y = 50 s-l. Rheopectic behavior (i.e. increasing mol % 1, after which shear thickening is observed, even in
viscosity with time a t a fixed shear rate) was observed for the absence of surfactants. Critical values of these
all samples, a t all shear rates examined, the viscosity parameters leading to shear thickening were also observed
increasing with time to a plateau value (Figure 10). The by Jenkins41for hydrophobically end capped poly(ethy1ene
magnitude of the viscosity increases and the time required oxide) (M,,= lo5). This behavior was seen for shear rates
to obtain the plateau value were dependent on the hy- of 1 s-l < y < 10 s-l in the absence of surfactants. No
drophobe content of the copolymers, the concentration of critical hydrophobe content was observed here between
the SDS, and the shear rate. For certain samples, the 1.03 and 1.3 mol % ,at a copolymer concentration of 0.5%
rheopectic effect was followed by a smaller thixotropic (w/w). Also, a previous report on similar copolymers
effect (Le. decreasing viscosity with time at a constant showed no critical concentration at up to 3% (w/w) for a
shear rate) before attainment of the plateau value of the hydrophobe content of 1 mol % .lo
viscosity. An example of such a curve is given in Figure Shear thickening was observed,however,within a certain
11. range of surfactant concentrations, for all the different
Rheopectic behavior was further investigated by sub; hydrophobe contents measured (Figure 9, curve b). The
mitting a sample to a series of alternate shear rates of y shear.thickening occurred at low to moderate shear rates
= 1 s-l and y = 20 s-l and recording the viscosity as a (1 < y < 20 s-l) and its magnitude was dependent on the
function of time. After each measure at y = 20 s-l (of amount of hydrophobe in the sample. Similar shear
fixed 5 min duration) increasing periods of relaxation in thickening has also been reported for the HEUR copol-
the absence of shear and before the measure at y = 1 s-l y m e r with
~ ~ ~anionic surfactants although in this case at
were employed. The results obtained at the shear rate of shear rates of between 100 and 1000 s-l. This dilatant
y = 1 5-l are presented in Figure 12 (curves a-c). As the behavior may be explained if we consider the relative
relaxation time is increased, the difference between the amounts of inter- and intrachain associations as a function
846 Langmuir, Vol. 8, No. 3, 1992 Biggs et al.

- Chain
re-coiling

Figure 13. Schematic representation of the effectsof shear time, at a constant shear rate, on the interactions in an aqueous solution
of copolymer/SDS: SDS, --; hydrophobe blocks, 0.

of shear. As the copolymer chains are extended under chains to a suddenly imposed shear strain, is practically
increased shear, the number of hydrophobic groups instantaneous. The second will be relatively fast also, since
available for interchain associations is increased due to it concerns only the motion of the small molecule sur-
the breakdown of intramolecular associations which re- factants and their attainment of the minimum free energy
quire a coiled polymer conformation. In the presence of state in the mixed micelles with respect to aggregation
micelles, these newly liberated hydrophobic groups can numbers and SDS/hydrophobe ratios. However, the third
be solubilized, leading to an increase in the number of stage which involvesthe relaxation of the copolymer chains,
interchain liaisons and hence the viscosity. As the shear in opposition to the imposed shear stress, in order to
rate is increased further, however, this structure is broken remove excess hydrophobe from solvent contacts may be
down in the shear flow and the solution becomes shear considered to be a slow process. Therefore, for a copolymer
thinning. This is further evidence therefore, that the sample at a high level of SDS there is sufficient surfactant
interactions between copolymer hydrophobic groups via to solubilize more of the initially liberated hydrophobe
the mixed micelles are considerably stronger than those than is the case at lower levels. Hence, the amount of
seen in the absence of surfactant, since the same elongation hydrophobe which must be protected from the solvent by
under shear of the copolymer coil is occurring but in the a rearrangement of the chain is lower and so the third
second case no dilatant behavior is observed. stage has a decreased importance. Thus the time required
The time required to reach an equilibrium value of the to reach the equilibrium plateau viscosity decreases with
apparent viscosity under a constant shear rate when not increasing amounts of surfactant.
in this shear thickening region is very short, usually of the In some cases, a small thixotropic effect was observed
order of seconds. However, when within the region of after the rheopexy and before attainment of the equilib-
shear thickening, the copolymer relaxation times increase rium value of the viscosity. This may be explained using
substantially, observed values often being greater than 5 similar arguments to those outlined above. The initial
min. These samples exhibit rheopectic behavior generally extension of the chain is followed by the formation of mixed
(Figure lo), with certain showing both rheopexy followed micelles with the SDS which facilitates interchain liaisons
by a smaller thixotropic effects (Figure 11). The rheopexy and hence causes a rise in viscosity. Those micelles formed
may be explained with similar arguments to those em- initially may be at a ratio of surfactant/hydrophobe which
ployed for the shear thickening above, i.e. under a shear does not give the lowest free energy for the system but
the copolymer chain has an extended conformation and does give a free energy level which is very favorable
more hydrophobe units are liberated for interchain as- compared to the free hydrophobe in solution. These mi-
sociations. The time required to arrive at an equilibrium celles will be kinetically, but not thermodynamically,
value and the magnitude of this viscosity are very sensitive stable, and so with time the surfactant will rearrange to
to the amount of SDS for any one copolymer sample at give surfactant/ hydrophobe ratios which offer a lower free
a fixed concentration. This is shown in Figure 10, where energy level at the same time as the copolymer chain relaxes
increasing amounts of SDS are seen to lead to decreasing removing the excess hydrophobe from solution. Hence,
equilibration times. The variation in the plateau values the viscosity undergoes a small decrease before attaining
of the viscosity with SDS concentration have already been equilibrium.
discussed in section 2 for the “steady state” values of the
viscosity (Figure 5). Therefore, no further discussion will Conclusions
be given here.
The time required to obtain the plateau value can be Rheological data on the interaction of hydrophobically
more easily understood if we consider the rheopectic modified polyacrylamide copolymers in aqueous SDS
process to occur in three stages. In the first, the copolymer solutions have shown that the relative numbers of inter-
chains undergo chain extension liberating a certain number and intrachain hydrophobic liaisons are very sensitive to
of hydrophobe units, from intrachain association, into the amount of surfactant and/or shear stress imposed on
solution. This is followed by a rearrangement of the SDS the system.
to solubilize the liberated hydrophobe forming mixed mi- SDS when added to an aqueous solution of copolymer
celles which may form a linkage between two or more binds to the copolymeric regions in a noncooperative
copolymer chains. In the final stage, the copolymer process causing a shift from intra- to interchain liaisons.
undergoes a conformational rearrangement to remove from This is evidenced by a peak in the apparent viscosity at
solution any of those hydrophobes liberated in the first surfactant concentrations below that of the cmc. Above
stage which have not been solubilized into the mixed mi- the surfactant cmc intra- and interchain interactions are
celles. These stages are shown schematically in Figure destroyed, each hydrophobe region of each copolymer
13. The first stage, involving the elastic response of the chain being solubilized in a single micelle.
Hydrophobically Modified Polyacrylamide Langmuir, Vol. 8, No. 3, 1992 847
In this region of enhanced solution viscosity, shear ment sites are very specific and are considered to be
thickening was observed at shear rates below 20 s-l. This separated from each other by long chains of noninter-
too is attributed to an increase in the number of inter- acting polyacrylamide.
chain liaisons as the copolymer chain is elongated under
the influence of the shear stress liberating further hy-
drophobe from intrachain liaisons. Solutions in this shear Acknowledgment. We thank Alain Hill for many
thickening domain were seen to exhibit rheopectic and helpful discussions. S.B. expresses his gratitude to the
thixotropic behavior. Such observations are attributed SERUNATO for the provision of a European post-
to the delicate free energy balance in these systems which doctoral fellowship. The financial assistance of ARTEP
is controlled by the entropy changes of removing or (Association de Recherches sur les Techniques
replacing hydrophobic groups in aqueous solution and of d'Exploitation du PBtrole) and PIRSEM (Programme In-
restrictions of the copolymer chain motion. terdisciplinaire de Recherches sur les Sciences pour
Contrary to homopolymer/surfactant systems none of l'Energie et les MatiBres PremiBres) is also gratefully
the rheological effects are attributed to a polyelectrolytic acknowledged.
effect as micelles become attached to the polymer. In the Registry No. SDS, 151-21-3; (H&=CHCONH*)-
case of hydrophobically modified copolymers, the attach- (H&=CHCONHCeHI-p-Et (copolymer), 112218-43-6.

Anda mungkin juga menyukai