Download as pdf or txt
Download as pdf or txt
You are on page 1of 24

ON POLYMER/SURFACTANT INTERACTION

E. D. Goddard

Specialty Chemicals Division


Union Carbide Corporation
Tarrytown, NY 10591

The field of polymer/surfactant interaction is reviewed.


A compilation of recent contributions, with special emphasis on
phase studies and hydrophobic polymers, is included. Lastly,
potential applications of polymer/surfactant interaction, with
particular reference to the cosmetics area, are outlined.

INTRODUCTION

In a variety of different fields, for example cosmetics, detergents


and foods, products are formulated to contain both surfactants and
polymers. Interaction between these two species, especially when existing
together in aqueous solution, is known to be very prevalent with attendant
implications regarding its influence on the properties of the products
themselves. In recent years a great deal of scientific study has been
carried out on the properties of polymer/surfactant mixtures in model
systems. The present article seeks to summarize the high points of these
studies, to describe some recent developments in this field, and simulta-
neously to identify areas seeming to offer opportunities for technical
exploitation of these systems with particular reference to the cosmetics
area. In the interest of limiting the length of this article biosystems,
including lipoproteins, will not be dealt with and the review will be
concerned largely with synthetic polymers and synthetic surfactants.

Before discussing mixed systems it is appropriate to briefly


describe the properties of each of these components when alone in aqueous
solution, namely the properties of surfactant solutions and of polymer
solutions. Such a discussion in a general sense will aid the understand-
ing of what interaction can be expected when the two species are present
in the same solution.

Surfactants

A surfactant is the classical ca,e of an amphipathic molecular


structure. In aqueous solution the rcaic headgroup, or the polar
headgroup (e.g. polyoxyethylene group), will be hydrated and the powerful
energetic forces of solvatior, involved are able to carry the whole

SurJactants in Solution, Volume 11, Edited by K.L. Mittal and


D.O. Shah, Plenum Press, New York, 1991 219
molecule into solution, even against the unfavorable process of transfer-
ring the alkyl group of the surfactant into an aqueous phase which
constitutes a hostile environment. But there are profound consequences of
this action. The migrations of the surfactant molecule in solution are
dominated by the desire of the headgroup to remain in solution, but of the
non-polar group to reduce contact with this hostile phase. Adsorption at
various interfaces allows the alkyl groups to self-associate and/or to
transfer to a less polar phase than water, e.g. to air, oil, or hydropho-
bic solid, as an adsorbed monolayer. Perhaps the best known manifestation
of the transfer of alkyl groups away from molecular contact with the
aqueous phase is, however, the self aggregation or micellization of
surfactants at concentrations equal to or exceeding the CMC. Micelli-
zation, in fact, represents a very delicate balance of intermolecular
forces which is strongly influenced by such factors as surfactant chain
length, temperature, added salt, and so on.

Polymers

Conventional water soluble polymers are based on hydrogen and


carbon, residing in hydrophobic groupings, together with oxygen and/or
nitrogen atoms in polar groupings to confer overall water solubility.
Thus, they too can possess an amphipathic nature. A cardinal feature of
polymers is their high molecular weight which endows the molecules with
special properties. Another feature is flexibility which allows
macromolecules to adopt various conformations in solution or at an
interface so as to achieve their most favorable energy condition. In
general terms, if the solvation forces are strong ("good solvent"
conditions) the polymer molecule will tend to expand, a readily observable
manifestation of this being maximal viscosity contribution to the
solution. On the other hand, if solvation forces barely compensate
internal cohesion forces in the macromolecule, the latter will tend to
coil and the solution viscosity will decrease. (These ideas are embraced
in Flory's concepts, where the latter state corresponds to "theta" condi-
tions and the former to "better than theta" solvent conditions. If
conditions are "much worse than theta" the polymer is insoluble l ).

WHAT HAPPENS ON MIXING A POLYMER AND A SURFACTANT?

The traditional view of polymer/surfactant interaction involved


binding of the surfactant ion (nonionic surfactants are much less prone to
interaction) on sites along the polymer backbone. This view probably
developed from the widely studied protein/surfactant, and later synthetic
polyion/surfactant, interactions where clear-cut indications of positive
ion/negative ion stoichiometry were obtained.

The sites of binding were never well specified when the polymer was
nonionic. This contrasts with the above-mentioned case of complexes
formed between a charged surfactant and an oppositely charged polyion in
which the latter's charged groups represent well-defined sites for binding
a surfactant ion by electrostatic attraction. The first step in binding
a surfactant ion to the polyion probably involves the attachment of a
single ion but it is clear that the process becomes highly cooperative
like an aggregation phenomenon. This may well be the case when the
polymer is uncharged also, but an increasingly accepted view is that
complex formation in this case is more appropriately viewed as a
perturbation of micelle formation. The exact molecular mechanism remains
to be settled and will depend on the specifics of the polymer and the
surfactant themselves: suffice it to say that the kinetics of aggregate
formation in the complex are comparable to those involved in ordinary
micellization.

220
10.0

1
~ 7.5
~
E-
O
Z
::J
0 5.0
'"
(J)
0
(J)
u..
0
I-
Z
::J 2.5
0
:;
<>:

o~ __ ~~-U~~~ __L -_ _ _ _ _ _- L_ _ _ _ _ _ ~ _ _ _ _ _ _ _ _L--"

o 0.5 1.0 1.5 2.0 2.5


CONCENTRATION OF "OUTSIDE" SOLUTION (SDS mM)

Figure 1. Binding isotherm of PEO-SDS system in O.lM NaCl. 3

Although a large number of investigative techniques have been used


to study polymer/surfactant interaction, we will restrict discussion here
to a few methods which shed light on the basic mode of interaction and
illustrate the influence of some of the main operating factors. 2

Surfactant Binding Studies

Direct measurements of surfactant binding by polymers afford a very


convincing demonstration of the formation of polymer/surfactant complexes.
Dialysis equilibrium, traditionally used for protein binding studies, has
been applied by a number of investigators to synthetic polymer surfactant
combinations. With this technique one must add indifferent electrolyte
(e.g. 0.1 M NaCl) to avoid Donnan membrane effects. As discussed later,
the presence of salt can influence the observed binding ratio. Combina-
tions of SDS and a variety of nonionic polymers (polyethyleneoxide - PEO;
polyvinyl pyrrolidone - PVP; polyvinyl acetate, alcohol - PVAc, PVOH;
methyl cellulose - MeC) have been examined by this method. Typical data
for the PEO system 3 are given in Figure 1. It is seen that below a
certain (monomer) concentration of SDS no binding occurs. The steepness
of the uptake curve suggests a highly cooperative process, and this is
followed by a plateau. Clearly, this behavior is fully compatible with
extensive results obtained by the surface tension method which formalized
the notion of surfactant concentrations, "TI" and "T2'" representing
initial binding and saturation of the polymer, respectively.4

Turning to charged polymer systems, Ohbu et al. 5 carried out


dialysis experiments to study the binding of SDS by a series of cellulosic
polymers of varying cationic substitution (CS). Binding was strong and
cooperative in all but the lowest CS case (0.05). The strength of binding
was indicated by the fact that it occurred at very low concentrations of
SDS; at 1/20th the CMC, the degree of binding, ~, had already reached the
value of 0.5 (~ =1 corresponds to 1 bound DS ion for each positive site on
the polymer). As noted above, the reason for the strong interaction in
such oppositely charge pairs is that potent electrostatic forces are
brought into play, reinforced, as will be seen, by hydrophobic interac-
tions.

221
1.0,------,-------------,------,

{3

0.5

-4.5 -3.5
LOG Mb
Figure 2. Binding isotherm of polyacrylate (Sxl0 4 M): comparison of
alkyl-pyridinium and -trimethylammonium ions. 9
(6)C 12PyCl; (O)C 14PyBr; ( .... )C 12TAB; (I)C 14TAB.

Suecific Ion Electrodes: The development of "membrane" electrodes,


specific for long chain surfactant ions, has provided a very useful tool
for studying the interaction of such ions with polymers. Illustrative
data are provided in the work of Kresheck and Constantinidis 6 on binding
of sodium decyl sulfate (SDecS) by PVP, presented in the form of
e. m. f. /SDecS concentration plots at various levels of added polymer.
Values of the binding obtained by these authors were 0.3 mols SDecS/base
mol PVP, in good agreement with the values obtained for SDS/PVP and
SDS/PEO in salt free solution obtained previously. Others who have used
the specific ion electrode include Birch et al. 7 and more recently
Takisawa et al. 8 who investigated the PVP/SDS and PVOH/SDS systems.

Returning to oppositely charged pairs, we refer to the extensive


studies of Kwak and co-workers who used membrane electrodes specific for
cationic surfactants in studies of their interaction with a series of
polyanions (2,9,10 and references contained therein). Typical data of
Kwak et al. are given in Figure 2. Important aspects are:

1. The binding curves are steep, indicative of cooperative binding.


2. Binding occurs at very low surfactant concentrations, « the CMC
value.
3. Considerable differences in binding characteristics in a series of
polyanions were found by Kwak. See below.

Precipitation in Oppositely Charged Pairs

The studies of Kwak and other investigators were confined to


relatively low levels of added surfactant since binding in these systems
is so energetic, i.e. occurs in most cases at concentrations a fraction of
the CMC of the surfactant. When, however, the charged surfactant and
oppositely charged polymer approach equivalence, on a charge basis,
precipitation reactions occur since the charge neutralized complex has
limited solubility. An interesting feature is that the precipitated
complex can be solubilized if excess surfactant is added to the system.
These effects are illustrated in the "solubility diagram" for the cationic
cellulosiC i Polymer JR 10/triethanolamine lauryl sulfate combination. See
Figure 3. 1 The 4So slope of the log/log plot, representing conditions
of maximum precipitation, signifies a constant composition of the insolu-
ble complex. It should be noted that resolubilization in such pairs

222
10 ~--------------------------------------------,

/
C C SP SPSP P //SP SP SP C C
1.0 •• • • • /
• • • ••
CC SP SP SP P P SP C C
•C •sp•sp• p• • T C• C• •
• • • .
C P
• • ••
CC SP SP P P T CC
0.1 • •/ • • ••
/
/
/
/ C=CLEAR
/
/ T=TURBID
P= PRECIPITATE
SP= SLIGHT PRECIPITATE

0.1 1.0 10

TEALS, WT%

Figure 3. Solubility diagram of Polymer JR 400jtriethanolamine lauryl


sulfate system. 11

depends on the specifics of both polyion and surfactant, and in some cases
the resolubilization areas are only partial or even absent. For example,
it has been found that if the charge density of the polyion is too high
resolubilization cannot be achieved. 12

Nuclear Magnetic Resonance

Early work by Muller and Johnson 13 on "F3 SDS" (SDS in which the ter-
minal CH3 group is replaced by a CF3 group) showed a characteristic NMR
shift (¢) when this surfactant underwent micellization, i.e. when the CF3
groups experienced a change from an aqueous environment to that of the
micelles. In the presence of PEO (constant level) and at low F3 SDS con-
centration the chemical shift was the same as that of submicellar con-
centrations of F 3 SDS. 14 At a certain concentration ("T 1"), lower than the
CMC of F 3 SDS, a slope change in the J vs. reciprocal concentration plot
was observed. At a second concentration ("T2 ") above the CMC the slope
changed again to a value close to that of polymer-free micellar F"3SDS
solutions. Tl was found to be independent of polymer concentration while
T2 increased with it. The similar values of J for the polymer bound
surfactant and micellized surfactant confirmed again the concept of the
surfactant molecules existing as aggregates in the former state.

In an extensive study of the PEOjSDS system, Cabane investigated the


13 C NMR shifts of the different carbons in SDS on adding increasing amounts
of PEO to micellar solutions l5 of SDS. Substantial chemical shifts were
noted for the three carbons (C l , C2 , C3 ) closest to the sulfate headgroup,
but were virtually absent for carbons C4 through C12 . Furthermore, the
shifts for C1 C2 and C3 were linear up to a concentration of added PEO
corresponding' to the T2 condition, i. e. where the polymer is just
saturated with surfactant. When the experiment was done in reverse, i.e.
SDS was added in increasing amounts to a fixed concentration solution of

223
PEO, a linear shift in the 13C line of PEO was observed up to a concentra-
tion close to the T2 value for the system.

One can draw a number of conclusions from Cabane's work. First, the
environment experienced by carbons C4 through C12 of the surfactant in the
polymer/surfactant aggregate is indistinguishable from that in regular SOS
micelles suggesting that the aggregates themselves are modified micelles.
On the other hand, the NMR data show that carbons C1 through C3 of SOS in
the polymer/surfactant aggregate are in a different environment and Cabane
suggests that EO groups of the polymer replace water molecules in the
outer region of the micelles. These ideas will be discussed later in the
article.

In a recent paper by Gao et al. 16 IH NMR studies were carried out on


mixtures of PEO and sodium W-phenyldecanoate. Clear indications of Tl and
T2 concentrations were obtained. See below. The use of FT-NMR to study
the self diffusion of dodecyltrimethylammonium (OTA+) ions in solutions
of ethylhydroxyethyl cellulose (EHEC) has recently been reported by
Carlsson et al. 17 Clear indications of binding were again obtained with
evidence of stronger binding at higher temperatures as the polymer became
more hydrophobic.

Gel Filtration

A direct and convenient way to obtain information on the formation


and properties of polymer/surfactant complexes involves the use of gel
permeation. Sasaki et al. 18 investigated the PEO 6000/S0S system using a
Sephadex GIOO column and an electrical conductivity detector and showed
that elution volumes, corresponding to decreasing molecular volumes,
increased in the order, complex < micelles < single surfactant ions.
Complex formation was found to occur above a concentration (T 1 ) of SOS of
4xlO-3 M. Illustrative elution plots have been given for SOS concentra-
tions (1) below T 1 , (2) above T 1 , but below T2 , (3) above T2 . For
condition (2) only one elution peak was observed for the polymer
suggesting that the complexed SOS is shared equally among the PEO
molecules.

Other Techniques

Several methods, in addition to the above, have been used to study


polymer/surfactant interactions. They include surface tension,
electrical conductivity, electrophoresis, ultracentrifugation, electro-
optics, optical rotatory dispersion, photochemistry, fast kinetics
techniques, fluorescence, calorimetry, electron spin resonance, X-ray
diffraction, electron microscopy, and small angle neutron scattering.
Typical results are summarized and reviewed in Reference 2.

FACTORS INFLUENCING COMPLEX FORMATION

Surfactant Structure

It has frequently been reported that interaction between uncharged water


soluble polymers and surfactants is much more facile with anionic surfac-
tants than with cationic surfactants and, in turn, is very much stronger
than with nonionic surfactants. If the polymer is sufficiently hydroph-
obic, clear-cut evidence of interaction with cationic surfactants and even
certain nonionic surfactants has now been obtained. This will be referred
to again under "Hydrophobic Polymers". A pronounced influence of the
particular counterion present in the cationic surfactant on its reactivity

224
with a polymer has been found by Saito. 19 This author20 also reported
that modification of the structure of an alkylsulfate by insertion of
ethyleneoxide groups considerably weakened its interaction with a polymer
(PVP). Branching of a surfactant molecule is also expected to weaken the
interaction. We refer again to the binding of tu-phenyldecanoate by PEO
studied by Kwak and co-workers using IH NMR. 16 Their data lead them to
the conclusion that in the complex both the phenyl group of the surfactant
and most of the PEO groups are deep in the interior of the mixed
aggregates. This result would not have been predictable, a priori.
Examples of the interaction of certain nonionic surfactants with
"hydrophobic" water soluble polymers are given later in article. Brackman
and Engberts 21 report that alkyl phosphates apparently react more feebly
with nonionic polymers than do alkyl sulfates, and that the interaction
decreases as the phosphate charge, Zo, increases from 1 to 2.

Chain Length

In a homologous series of alkylsulfates, the initial binding


concentration, T 1 , decreases with increasing chain length. 22-23 A linear
relation between log TI and n, the number of carbons in the alkyl chain,
viz. ,

InTI = n LV/kT + constant

exists as has been found between log CMC and n. For PVP/SDS mixtures Arai
et al. 23 found a value for uJ of -1.1 kT. This corresponds to the free
energy change per CH2 group on transferring the surfactant from the
unassociated state in water to the complex, and is comparable to the value
for the analogous transfer of the surfactant molecule from solution to an
ordinary micelle.

Goddard and Hannan l2 studied the solubility diagrams of a homologous


series of alkyl sodium sulfates in mixture with the cationic cellulosic
polymer, Polymer JR. In each case it was found that, in the limit, the
slope of the points of maximum insolubility in the plot of log polymer
versus log surfactant concentration changed from 45 0 to 90 0 (i.e. became
independent of polymer concentration, if the latter were reduced below a
certain value). Mathematically, the result could be expressed as

Ceexp (n W/kT) = constant

where Ce is the polymer concentration- independent surfactant concentration


corresponding to maximum precipitation. A value of W of -l.l kT was
derived, suggesting that the environment of surfactant molecules in the
complex resembles that of micelles in this case also.

Similar trends for the binding of cationic surfactants to polyanions


(e.g. polyacrylate) have been reported ~ Kwak and co-workers, previously
referenced, by Binana-Limbele and Zana2 and by Chu and Thomas. 25

Effect of Salt

In the case of unionized polymer/charged surfactant systems, addition


of salt depresses the TI values, i. e. promotes the formation of complexes.
Murata and Arai, for example, found the log-log plot of TI against sodium
ion concentration for the PVP/SDS association to be linear with a slope
exactly the same as that of the CMC/sodium ion concentration plot for this
surfactant. 26 Addition of salt also increased the binding ratio of
surfactant to polymer, i.e. extended the T 1 ,T 2 range. For example, in the
case of PVP/SDS, addition of 0.1 M NaCl increased the ratio to 0.9 mol SDS

225
per base mol PVP from the 0.3/1 ratio observed in water. A similar effect
occurs with PEO/SDS. 27

A very different effect of salt occurs if the polymer is a polyion


and the surfactant bears an opposite charge to it. In this case, while
the addition of salt increases the steepness, i. e. the "cooperativity",
observed in the binding isotherm, it substantially reduces the affinity
of binding as evidenced by a steady increase in the concentration at which
binding commences. 28 This result clearly points to the importance of
electrostatic attraction between the polyion and the oppositely charged
surfactant as being a primary driving force for the association. That
electrical screening by the salt is the operating mechanism in weakening
attraction is verified inasmuch as the effect is magnified if the salt
contains divalent ions, as opposed to monovalent ions, bearing a charge
opposite in sign to that of the polyion. 29

The Polymer

A minimum molecular weight of polymer is apparently required to


ensure "complete" interaction with the surfactant,30 and this value is
4000 for PEO and PVP. Below values of ca. 1500 the interaction tendencies
with these polymers are restricted.

Until quite recently, the list of unionized polymers showing the


ability to form complexes with ionic surfactant was quite small. The
"traditional" list included PEO, PVP and PVOH. The lack of reactivity in
this respect of other polymers, such as hydroxyethylcellulose, HEG, was
thought to be due to a lack of macromolecular flexibility, but even the
more flexible polysaccharide dextran shows little tendency to interact
with SDS or dodecylbenzenesulfonate, DDBS. Likewise, the relative
inactivity of polyacrylamide, PAAm, remained a puzzle. It was recognized
that reactivity can be induced in a polymeric structure by introducing
hydrophobic sites in the macromolecule. Examples are methylcellulose,
MeG, versus HEG; low molecular weight polyalkyleneoxides (PPO) in which PO
replaces EO; and PVOH specimens prepared from, but still containing,
residual amounts of PVAc. 31 In fact, the reactivity of polymers seems to
correlate with a kind of "hydrophobicity index". For anionic surfactants,
Breuer and Robb 32 listed polymers in the following order of increasing
reactivity:

PVOH < PEO < MeG < PVAc < PPO ro./ PVP

and for cationic surfactants:

PVP < PEO < PVOH < MeG < PVAc < PPO

Recently, the notion of the importance of hydrophobicity in the


polymer in promoting reactivity with surfactants has had strong reinforce-
ment. It has been shown by Winnik et al. 33 and by Lindman and co-workers
that HPG, hydroxypropyl cellulose 34 , and EHEG, ethylhydroxyethyl cellu-
lose,34 and of course MeG, 34,35 all display pronounced association
tendencies towards SDS and cationic surfactants, and it can be expected
that "reactivity" can be induced in PAAm by methyl (or other alkyl
substitution) of the nitrogen in the amide group of this polymer. Indeed,
Schild and Tirrel1 36 recently demonstrated reactivity of N-isopropyl
substituted PAAm with a range of alkyl sulfates. These results support
the earlier data of Murai et al. on the reaction of SDS with a series of
model polypeptides. 37 It should be mentioned that very recent work on so-
called "associative polymers" (alkyl substituted water soluble polymers)
is clearly indicating that these structures have pronounced interaction

226
tendencies with added surfactants. This important subj ect is treated
later.

In the case of ionized polymers, if the charges on the polyion and


the surfactant have the same sign then strong forces of electrostatic
repulsion are involved and little to no association can be expected.
Examples are NaCMC or ionized polyacrylic acid in the presence of anionic
surfactants. The position is reversed in the opposite case, and in fact
the polyion now presents well-defined electrostatic binding sites for the
oppositely charged surfactant ions. The process is strongly reinforced by
alkyl chain association of the adsorbing surfactant molecules. Examples
are the interaction of alkyl sulfate surfactants with cationic-cellulosic
or -vinyl polymers.

Several properties of the polyion influence the reaction with


oppositely charged surfactants. This is apparent in the binding sequence
observed by Kwak et al. 38 ,39 for DTAB, viz., polystyrene sulfonate >
dextransulfate > polyacrylate > DNA > alginate ~ pectate > NaCMC.
Involved evidently are structure, charge density, hydrophobicity, macromo-
lecule flexibility and other factors. The reader is referred to Kwak's
papers for more detailed information. It should again be mentioned that
the ability of excess surfactant to resolubilize the insoluble poly-
mer/surfactant complex formed under "stoichiometric conditions" also
depends on the charge density and structure of the polyion.

INTERACTION MODELS

For both types of polymer/surfactant systems, viz., polymers charged


or uncharged, theories developed to account for the formation of poly-
mer/surfactant complexes have postulated structures with the surfactant
molecules in the form of aggregates or clusters. Much information,
including NMR data (referred to above) and solubilization data (referred
to below), points to individual surfactant molecules experiencing an
environment in the complexes similar to that which they encounter in
regular surfactant micelles.

Most models for interaction of surfactant with uncharged polymer can


be regarded as elaborations of the simple model of Smith and Muller. 14 In
explaining their data for the SDS/PEO system these authors postulated that
each macromolecule consists of a number of effective segments of mass Ms
and total concentration [P] which act independently of each other and are
able to bind n surfactant anions, D-, in a single step according to

with the equilibrium constant being given by

K is obtained from the half saturation concentration, viz.,

The data indicated the cluster size, n, to be about 15 (which is much


smaller than in regular micelles) and Ms to be 1830 which explains the
experimental finding that PEO of MW 1500 is relatively ineffectiv~ for
surfactant binding while higher molecular weight PEO's are effective. The
free energy of binding is given by

227
and the value obtained, viz., -5.07 kcal mole-I, is close to that of
micelle formation.

Shirahama3 introduced a similar model to account for his binding data


of SDS on PEO in the sense that cluster formation of the bound surfactant
molecules (with n - 20 molecules) was invoked and the degree of binding,
e, was expressed by a modified Langmuir isotherm of the form

More comflete models have been proposed by Nagaraj an40 and by Gilanyi
and Wolfram. 4 Nagaraj an, for example, sets up a mass balance equation for
the total surfactant in the system which includes surfactant in single ion
form, in clusters of size gb in the polymer/surfactant complexes, and in
micelles of size gf. Binding (Kb) and micellization (Kf) constants are
also assigned. Depending on the relative values of Kb and Kf' one can then
ei ther encounter complex formation followed by micellization or only
micellization, and the steepness of the binding curve will be determined
by the degree of cooperativity or gb. With this model Nagarajan was able
to account quantitatively for observed binding data, while confirming the
essential similarity of the processes of complex formation and of micelle
formation. The picture which emerges is that these two processes are
essentially competitive. If Kb >Kf complex formation will occur until the
binding process is complete after which, on further increase of surfac-
tant, regular micelles will form. (This would lead to phase diagrams of
the type developed by Sasaki et al. IS and other workers 27 ,40). Another
factor supporting the similarity of the two processes comes from the fast
kinetics measurements of Wyn-Jones.42~ A surprising result is that the
kinetics of aggregate formation are at least as fast as those of
micellization, itself a very fast process with relaxation times in the
microsecond to millisecond range. A further point of interest is that
the activation energy for surfactant clustering in complex formation is as
little as one third that involved in micelle formation of the same sur-
factant. 39 More recently, models have been developed by Ruckenstein et
al. 45, Nagaraj an46 and Hall. 47

A simple explanation of all these phenomena is seen in the picture of


the surfactant/(unionized) polymer complex put forward by Nagarajan,40
Shirahama,4S Landoll 49 and Cabane. I5 See Figure 4. It is noted first that
in the formation of regular ionic surfactant (spherical) micelles, the
classical case of a cooperative aggregation phenomenon, a major resisting
force is the crowding together of ionized headgroups at the periphery of
the micelle and the development of a high electrostatic potential which
can be offset only partially by counterion binding. Furthermore, in the
well-accepted spherical, or Hartley, micelle there is a considerable
distance, on a molecular scale, between the headgroups at the periphery,
if only for geometrical reasons of packing. Some of this space will
accommodate conterions, but most will comprise areas of the hydrocarbon
chains exposed to water - an obviously unfavorable situation. Early NMR
data indicated that the first few carbons (measured from the headgroup) of
micellized molecules of SDS remain in contact with water. One can easily
imagine a "loopy" configuration of water soluble polymer, associating with
a micellar array of surfactants, which allows ion-dipole association of
the hydrophilic groups of the polymer and the ionic headgroups of the
surfactant and, in addition, contact between the hydrophobic segments of
the polymer and the "exposed" hydrocarbon areas of the micelle--in effect
resulting in screening of the electrical charges and diminution of the

228
a

b
Figure 4. Schematic diagram of a) anionic surfactant micelle and
b) polymer/surfactant complex. 40

extent of these exposed areas. Consequences of the above would include


several features already observed, such as

1. A more favorable free energy of association, as manifested in a


lowered "CMC" (i.e. Tl < CMC).
2. Increased ionic dissociation of the aggregates.
3. An altered environment in the CH2 groups of the surfactant near the
head group, as seen in 13 C_NMR results.
4. Increased associating tendency as the polymer becomes more hydropho
bic.

A major point of difference from the above systems exists when the
polymer and surfactant are oppositely charged; as we have pointed out, in
this case there are discrete binding sites for the surfactant ions and
binding is reinforced by alkyl chain association, and this can also be
considered a special case of surfactant aggregation. There are strong
analogies between the process of surfactant adsorption, leading to complex
formation of the surfactant with the polyelectrolyte, and adsorption,
leading to hemi-micelle formation, of ionic surfactants on the surface of
oppositely charged solids, such as minerals. In both cases an ion-
exchange process is involved in which the counter ion of the polyelectro-
lyte (or charged surface) is replaced by the surfactant ion and binding
commences at a concentration orders of magnitude below the CMC of the
surfactant. A differentiating factor in the case of the polyelectrolyte
is the molecular flexibility of the charge-bearing substrate meaning that
its properties, such as conformation, can be substantially altered by the
adsorption process and actually reinforce it. One of the models, developed
by Satake and Yang,50 for binding is, in fact, based on the Zimm-Bragg
theory for coil-helix transitions of polymers as adapted to the coopera-
tive binding process. Precisely the same relationships as those obtained

229
&----POLYMER
/
POLYMER
PRECIPITATION
ZONE

CONCENTRATION SURFACTANT-

Figure j. Conditions in the bulk and surface of solutions containing a


polycation and increasing concentration of an anionic
surfactant. Simple gegenions are depicted only in the surface
zone. 2

by Satake were derived by Shirahama51 -52 who employed a statistical


mechanical treatment of the binding process. Lastly, Delville53 has
presented a model based on two additive effects, one due to Poisson-
Boltzmann distribution of the surfactant ion and the other a contribution
due to cooperative binding, to describe the process. At high surfactant
concentration, in the post-precipitation or re-solubilization zone,
"string of beads" structures have been invoked, in which the beads are
surfactant clusters and the polyion is the string. Possible structures
are depicted in Figure 5 which links the surface and bulk behavior with
compositional changes in the system. 2

SOME PROPERTIES AND OPPORTUNITIES RELEVANT TO COSMETIC SCIENCE

There are many changes in solution and surface properties which occur
as a result of polymer/surfactant association. Some which would appear
relevant to cosmetic science follow.

Viscosity Enhancement

It is known that polyelectrolytes tend to adopt a linear configura-


tion in aqueous solution, unless the ionic strength of the solution is
high. This configuration favors an increase in viscosity of the solution.

When an unionized polymer binds an ionic surfactant, the polymer will


acquire a charge and hence, on the basis of the "polyelectrolyte effect",
an increase in viscosity would be anticipated. This area has. been
investigated by a number of authors, e.g. Jones 4 and Francois et al. 54 who
examined the effect of adding SDS to a series of PEO polymers. A sudden
increase in viscosity occurs at a certain concentration (T 1 ) of surfac-
tant, independent of polymer molecular weight, and this increase can be as
high as five-fold, which is consistent with a polymer charging effect.

230
5000 sos+
/ 1 % POLYMER JR-400
4000 .,#-

>-
f- 3000
Ui I
0 2000 I
U
r/) I
:; 1000
~
w c ~)
~ PRECIPITATION:
3 125 ZONE I
0 I
t:. 100
I
w I
> I
~
..J
75
W
a: 50

25

0.0001 0.001 0.01 0.1 1_0 10.0


SOS CONCENTRATION %

Figure 6. Relative viscosity of 1% Polymer JR400 as a function of added


SDS concentration.

Even larger increases can be seen with certain oppositely charged


polyelectrolyte/ionic surfactant pairs as illustrated in Figure 6. In
this case involving the cationic cellulosic, Polymer JR400, relatively low
levels of added SDS lead to substantial viscosity increases (as large as
200 fold), or even weak gel formation, in the immediate pre-precipitation
zone. 55 The most likely explanation is that "super macromolecules" are
formed through association of the alkyl groups of surfactant ions bound to
different polymer molecules. Such a structure would have pronounced shear
thinning characteristics, thus providing opportunities for rheology
control. It should be mentioned that thickening effects on addition of
surfactant are dependent on the structure of the polyelectrolyte. Thus
the much more flexible polycation, Reten (Hercules), based on vinyl
chemistry, did not show the viscosity increase displayed by the "stiff
backboned" cationic cellulosic/SDS combination. Strong gel formation by
the highest MW grade of Polymer JR(30M) with added SDS will be described
later.

Solubilization

Micellar surfactant solutions are well known for their ability to


dissolve oil soluble materials, e.g. dyes, hydrocarbons, esters, perfumes,
and so on. To the extent that complex formation with a nonionized polymer
can be regarded as a depression of the aggregation concentration of the
surfactant (i.e. Tl < CMC) enhanced solubilization by the complex can be
anticiEated. This effect has been confirmed using water insoluble
dyes 22 ,56,57 and sparingly soluble fluorescers. 58 Much more pronounced
effects have been found for polyelectrolyte/ionic surfactant pairs. 55 For
the cationic cellulosic/SDS pair a solubilization region for the dye,
Orange OT, occurs at very low concentration and the "main" solubilization
zone is also widened (shifted to lower concentration) as compared to
simple SDS solutions. 55 Solubilization at the very low concentration of
SDS signifies clustering around the polymer's positive charges in the
initial binding process.

Since polymers and surfactants can associate in solution it would not


be surprising if they could influence each other's solubility as well as

231
that of a third component. Perhaps the best known case of this effect was
reported by Isemura and Imanishi 59 who showed that a PVAc polymer of very
low solubility could be solubilized in solutions of SDS. Another specific
case of solubilization of polymers by surfactants is treated in the next
section.

In an interacting nonionized polymer/ionic surfactant pair it is


logical to expect that increased solubility could be manifested in the
opposite sense, i. e. the polymer could increase the solubility of the
surfactant since the monomer concentration required for aggregation of the
surfactant is lowered in the presence of the polymer. Such an effect has,
in fact, been reported by Schwuger and Lange who showed that PVP can
reduce the Krafft point of sodium hexadecyl sulfate by close to 10°C. 60

Note: It is well known that many conditioning polymers are


polycationic, and we have pointed out that precipitation zones exist at
certain ratios in combinations of such polyelectrolytes with anionic
surfactants. In most cases, however, such precipitates can be solubilized
in the presence of excess surfactant or prevented by the co-presence of a
nonionic surfactant. 12,61,62

Cloud Point Elevation of Polymers

Most uncharged polymers owe their solubility to the presence of polsr


groups, such as ether, hydroxyl, amide, carboxyl, etc., which will hydrate
in the presence of water. This hydration, especially of an ether group,
can diminish progressively with temperature and the critical balance
governing solubility can be upset at a specific temperature ("cloud
point") at which the polymer comes out of solution. If the polymer can
acquire charges, e.g. by ionization of acidic or basic groups or by the
adsorption of a charged species, such as a surfactant, enhanced solubility
or elevation of the cloud point can be expected.

There is much evidence in the literature to illustrate these effects.


Although PED itself is not amenable to such studies (its cloud point
exceeds 100°C) related polymers provide such information: PPD, by dint of
possessing hydrophobic methyl groups, has much lower water solubility than
PED: PPD's of MW 1025 and 2000 have ctfud points of about 40°C and 20°C,
respectively. Pletnev and Trapeznikov showed that SDS and sodium dodecyl
benzene sulfonate can raise these values to above 90°C. (In one sense,
standard nonionic surfactants can be regarded as PEO polymers: that one
can raise their cloud point by addition of ionic surfactants is a fact
very well known to formulators). There are several references to
increases of the cloud point of PVOH polymers by addition of anionic
surfactants and similar elevation is well known for MeC. This subject is
discussed further below.

Reduction of Monomer Concentration

The fact that, in the presence of polymer, aggregates of surfactant


can form at concentrations lower than the CMC means that the maximum
monomer concentration of surfactant is reduced. For nonionized polymer/
ionic surfactant combinations the region involved would be in the Tl T2
concentration range. For polyelectrolyte/ionized surfactant pairs this
effect would be maximal in the pre-precipitation binding zone. Because of
the strong bonding forces involved in the latter case, the reduction in
monomer concentration would tend to be much higher.

Although the picture concerning the irritation to skin caused by


exposure to surfactants, in particular anionic surfactants, is not
completely clear much evidence exists which suggests that lowered monomer

232
concentration of the surfactant can correspond to lowered irritation. 64
Coupled with this, there is evidence 65 ,66 that addition of selected
polymers to solutions of anionic surfactants can reduce the irritation
caused by the latter. The implication is that formulations of lowered
irritation potential could be more reliably created on this basis if a
knowJedge of the binding characteristics of the particular polymer/surfac-
tant combination chosen were established. On the other hand, a definite
possibility always exists that any observed reduction of irritation
occasioned by the presence of a polymer may involve more than one
mechanism. 67

Surface Activity and Adsorption Characteristics

It has been pointed out that nonionized polymers may reduce the
surface activity (air/water interface) of ionic surfactants by binding the
latter in the form of weakly surface active complexes. On the other hand,
the formation of highly surface active complexes between polyelectrolytes
and oppositely charged surfactants provides a synergistic enhancement of
the surface activity of the two components ~ Consequences of this
enhancement could include improved emulsifying abilit~ and, especially,
foaming: instances of the latter have been reported. 68 ,69 At the same
time, several references in the literature can be found on improvement of
the foamability and foam quality of ionic surfactants by the addition of
nonionized polymers, such as PVOH, PEO, modified starches, cellulosics and
so on. Such implied alteration of surface properties can be understood
if one considers the adsorbed layer of surfactant to resemble a surfactant
micelle (in this case a "semi-infinite hemi-micelle") and be able to
interact with the polymer in a way similar to that depicted in Figure 5,
top center. In other words, in this case also, surfactant and polymer
could influence each other's adsorption characteristics, and hence surface
properties, including foaming, would be affected, even though the energy
of association of the surfactant and the base monomer may be weak.

As regards solid surfaces, since both polymer and surfactant can


adsorb on such surfaces there has been much interest and considerable work
done to determine what effect each has on the extent of adsorption of the
other. Most of the work done has involved mineral (or latex) solids 2 and
will not be detailed here. Suffice it to say that positive and negative
effects have been found for both adsorbing components depending on
conditions, i.e. the actual components themselves, addition sequence, the
solid surface, the pH, and so on. As the adsorption energy per monomer
unit of (especially) a nonionic polymer can be quite weak it is not
surprising that its adsorption overall can be affected by an added
surfactant. The clear implication is that opportunities exist to modify
the surface characteristics of chosen solids by appropriate choice of
surfactants and polymers. Since the solid introduces a new phase (as in
conditioning) and since other ingredients may also be present, determina-
tion of improved adsorption characteristics has in most cases to be done
empirically.

Polymeric Surfactants

There is at present a great renewal of interest in "polymeric surfac-


tants" and growing commercial recognition of their importance as so-
called "associative thickeners" for use in latex and other formulated
product systems. Th~se materials are in effect conventional water soluble
polymers which have been modified by inclusion of hydrophobic moieties, in
particular alkyl groups. 70 They combine the properties of a surfac-
tant/polymer mixture in one molecule and therefore display some of the
properties of such mixtures. Thus, when dissolved in water they tend to

233
self-associate, generating association structures of high molecular weight
and hence substantial increases in viscosity of their solutions result.
Several recent papers on the subject of associative thickening polymers
may be found in Reference 71, as pointed out below. Other consequences of
the associative tendency of molecules with this type of structure are the
development of solubilizing properties for water insoluble materials,
including dyes, and also a strong tendency to produce foams, especially,
foams of unusual stability: Goddard and Braun in fact showed that a
hydrophobica11y modified cationic cellulosic polymer, Quatrisoft LM (Union
Carbide Corp.) itself could form the basis of an aerosol mousse. 72

SOME RECENT DEVELOPMENTS

The field of polymer/surfactant interaction is currently experiencing


considerable activity. It is convenient to group (much of) these
activities into three, somewhat overlapping, areas.

Phase Behavior

Elevation of the cloud point of various unionized polymers (generally


containing ether oxygens) by added surfactants has been referred to.
Cloud point alteration has usually been considered to reflect a monotonic
increase with increase in ionic surfactant concentration but recent work
has shown the first additions of ionic surfactant can sometimes lower the
cloud point, especially when salt is present. 34,73-75 In fact, the
phenomenon seems to be unusually sensitive to the presence of salt.
Kar1strom et al. 34 have studied the phenomenon in detail and have presented
phase relationships for a number of cellulosic polymer (especially
EHEC) /ionic surfactant pairs. The actual cloud point vs. surfactant
concentration plot is affected by a number of factors, viz., (1) the
polymer itself; (2) the surfactant (structure and chain length); (3) the
presence of salt and its concentration; (4) the particular salt chosen.
Specific ion effects, well known in salting out/in of polyether nonionic
surfactants and polymers, are pronounced. In a qualitative way it is
proposed that the cloud point of the polymer is raised by association with
an ionic surfactant because of electrical repulsion between the (now
charged) polymer molecules. Evidently, this effect is very sensitive to
and can be offset by electrical screening on adding salt.

In a very interesting development this group34 has also attempted to


model the phase behavior of mixed systems of polymer and surfactant using
a Flory-Huggins approach. The components are (1) water, (2) polymer, and
(3) surfactant micelle. By making various assumptions, and having the
"degree of polymerization" of the micellar species and the various
interaction parameters as adjustable, rather good agreement with observed
behavior could be obtained.

For systems consisting of po1yion/opposite1y charged surfactant


pairs, very little phase work has been done until fairly recently.
Goddard and Hannan 12 in the mid 1970' s determined crude "solubility
diagrams" for pairs of po1ycations and anionic surfactants. A much more
complete investigation of the phase behavior in ionic binary systems has
recently been presented by Tha1berg et a1. 76 They studied sodium hyaluro-
nate (a high molecular weight copolymer of (3 - acetyl glucosamine and
glucuronic acid)/a1ky1trimethylanunonium bromide combinations in detail and
have mapped the phase behavior in ternary diagrams. See, as an example.
Figure 7. For such systems also, the behavior was successfully modelled
using F1ory- Huggins concepts.

234
High Viscosity Systems and Gel Formation

A special (extreme) case of viscosity development on adding a


surfactant to a polymer solution is the formation of gels. The demonstra-
tion by Leung et al. 55 of pronounced viscosity enhancement in the "pre-
precipitation" range of cationic cellulosic Polymer JR 400/SDS systems
prompted this group77,78 to examine SDS combinations with a higher MW
analog (JR 30M) where chain entanglements required for formation of the
gel structuring network would be enhanced. Strong gels were indeed formed
at concentrations of polymer of -1% and about one tenth this amount of
SDS. Rheological characterization showed that the complex modulus G*, was
dominated by the elastic component, G'; also the "zero frequency" dynamic
viscosity, was very high (-10 kPas). G' was found to increase with SDS
concentration, up to 0.15%, and then to decrease. Similar trends were
found with several other types of anionic surfactant (alkyl-ethoxysulfa-
tes, -benzenesulfonate and -sulfosuccinate). It seems likely that gel
formation in these sydtems results both from an increase in effective MW
of the polymer through cross-linking of bound clusters as well as from
increased chain entanglement.

10

H20----~~------~------r-----_7---NaHY
10 20 30 40
%NaHy - -
Figure 7. Phase diagrams for sodium hyaluronatej(MW 2.5 X 105 ), water and
surfactants C10 -' C12 - and C14 - TAB. 76

Recently Carlsson et al. 79, in an extension of their work on the phase


diagrams of nonionic cellulosic (usually EHEC) /charged surfactant systems,
demonstrated that the formation of high viscosity systems occurred under
particular conditions ("windows") of temperature and added concentration
of surfactants (usually CTAB) in the single phase zone. See Figure 8. A
rationalization of the behavior was as follows: As the temperature is

235
80 Two-phase One-phase
70 region region

60

--
U
0
50
0-

~ 40 Enhanced viscosity J
30
20
~
10
0 5 10 15 20 25
CTAB cone / mmolal

Figure 8. Part of the phase diagram for 1% EHEG and GTAB in water. 79

increased the polymer becomes less soluble, i.e. more hydrophobic, and
more prone to interact with the surfactant, then the tendency to crosslink
via adsorbed alkyl chain increases and the viscosity rises progressively
until gels are formed, i.e. in those systems evincing gel formation. A
slow decrease of viscosity/gelling tendency on further increasing the
temperature may reflect the tendency of the polymer itself to undergo
thinning at higher temperature. An optimum range of surfactant concentra-
tion for gelling is easy to explain: a minimum amount of surfactant is
required to initiate binding to the polymer and generate a sufficient
number of mutually interacting adsorbed alkyl chains. When the polymer is
saturated with surfactant, internal interactions (i.e. within one polymer
molecule) are more likely and the clusters are probably more micelle-like
and hydrophilic so reducing the tendency for the chains to form cross-
links. Addition of excess surfactant, in fact, constitutes a viable way
to "de-gel" these systems_

A very interesting case of "de-gelling" caused by polymer-surfactant


interaction was reported by Brackman and Engberts. 80 Polypropylene oxide
was added to a gel formed by combination of the long chain cetyltrime-
thylammonium cation and the salicylate anion, a system well known to form
viscoelastic solutions owing to the presence of long rod-like micelles.
Evidently the competing interaction between the GTA+ ions and polypropyl-
ene oxide is strong enough to disrupt the gel structure.

"Hydrophobic" Water Soluble Polymers

The favorable influence of some hydrophobic character in the water


soluble polymer, especially if uncharged, on its interaction with
surfactants has already been mentioned. A simple indicator of reactive-
ness is the surface tension of aqueous solutions of the polymer by itself.

Thus Goddard at al. 81 have shown that the relatively inert polymer,
HEG, is surface inactive while the much more reactive methylcellulose is
known to have appreciable surface activity. By the same token, inactive
polyacrylamide has very limited surface activity and is unreactive towards
SDS and sodium dodecanesulfonate 82 By contrast, it is to be expected that
N-alkyl substituted PAAm would be both surface active and reactive, as has

236
indeed been found for poly(N-isopropylacrylamide) and a series of sodium
alkyl sulfates,36 as was mentioned above. We would term such polymers as
having "accessible" hydrophobic groups providing potential sites for
surfactant adsorption.

In the same way, highly surface active polypropylene oxide is much


more interactive than less surface active polyethylene oxide (see Engberts
and co-workers 83 ,84). One penalty of higher surface activity of the
polymer, however, is that the very useful surface tension method of
exam~n~ng the interaction pattern with a surfactant becomes less
informative and more difficult to interpret. Fortunately, as pointed out
above, there are a large number of alternative methods to study the
interactions. As recent examples, the use of cloudpoint determinations,
phase mapping, rheology and NMR (for self diffusion) in studies of a
series of surface active cellulose ethers (EHEC, MeC, HPC) and ionic
surfactants by Lindman's group has been referred to. Clear indication of
pronounced interactions was obtained in all cases. An interesting
variation on the theme of cellulose ether interaction with a surfactant
was presented by Winnik85 who studied the interaction of pyrene labelled
HPC with the nonionic surfactants, n-octyl ~-D-(thio)glucopyranosides (OG
and OTG). Evidence of clustering of the surfactant molecules around the
pyrene chromophores was obtained. These results are reminiscent of the
earlier work of Brackman et al. 83 who showed that, if the water soluble
polymer is sufficiently hydrophobic (PPO, MW 1000), association with a
nonionic surfactant (the same OTG) does indeed occur. Other studies on
the HPC/SDS system have been reported by Winnik and Winnik86 who have
deduced that the surfactant cluster size increases with SDS concentration
above T 1 , unlike the case of SDS clusters on PEO. Pyrene labelling groups
on a fairly hydrophilic polymer, e.g. PEO, provide hydrophobic centers for
interaction. In mixtures with SDS at lower concentration (-8XI0 4 M) the
polymer chain cyclized, with the pyrene groups being in a simple micellz
as shown by the fluorescence data of Hu et al. 87 ; at higher SDS concentra-
tion the pyrene groups of the polymer were found in separate micelles. 87
In a general sense, the subject of polymer/surfactant aggregation has been
modelled by computer simulation by Balazs. 88 Predictions of a critical
concentration for interaction and of an optimal chain length of the
surfactant for promoting associations have been confirmed.

The subject of water soluble polymers with "well-defined" hydrophobic


groups is enjoying a considerable revival. This is a natural outgrowth of
the work in the 1950' sand 1960' s on "po1ysoaps" or "polymeric surfac-
tants", notably vinyla1ky1ether/maleic anhydride copolymers, alkyl
derivatives of polyvinyl pyridine, and so on. Analogies between polymeric
surfactants and conventional surfactants have led several authors to
obtain evidence, chiefly by fluorescence methods, of self-association of
the hydrophobic groups of such polymers in solution. See papers by Binana-
Limbele and Zana24 and Hsu and Strauss, 89 and references contained therein,
for studies of vinylalkylether/maleic anhydride copolymers, and by Chu and
Thomas 90 and Shih et al. 91 ,92 for studies of eX -olefin/maleic anhydride
copolymers--in the latter case examined by SANS. The self association of
pyrene -labelled - PE093 ,94 and _HPC 95 provides further illustrations of this
effect as does the self association of a PEO polymer terminally tagged
with alkyl groups.96 Viscosity studies on alkyl substituted HEC ("HM-HEC")
solutions have been carried out by Landoll,97 Gelman and Barth,98 and by
Goodwin et al. 99, all of whom obtain evidence of polymer aggregation
beyond a certain concentration. For a recent compilation of contributions
in this area, we recall Reference 71. Lastly, a model for the formation
of "micelles" by these types of amphiphilic polymers has very recently
been developed by Hamad and Qutubuddin. 1OO Since the present article is

237
Figure 9. Depiction of association structures of hydrophobically
modified polymers, end-substituted and "comb" (Courtesy
D. R. Bassett).

concerned with polymers with added conventional surfactants this important


area is not pursued further here.

D. R. Bassett lOI has prepared a qualitative depiction of possible


structures involving association between hydrophobic polymeric species,
including one in which micelles of added simple surfactant participate.
See Figures 9 and 10. On defining "polymeric surfactants" as water
soluble polymers with alkyl group sub~tituents of chain length> C10 ' one
finds relatively few published papers on the subject of mixtures of these
polymers with conventional surfactants. We will refer briefly to a few

Figure 10. Depiction of associating polymer structure with surfactant


micellar bridge (Courtesy D. R. Bassett).

238
such studies, all based on uncharged hydrophobically modified cellulose
ethers. Sau and Landoll 102 have reported substantial viscosity boosting
effects on adding a nonionic surfactant to a dilute solution of HM-HEC.
Viscosity peaks with added anionic surfactants were also found by these
authors and by Dualeh and Steiner 103 and Lindman and co-workers 79 , with
added SDS yielding gels under certain conditions. l03 Variable effects for
such systems are evident in the work of Sivadasan and Somasundaran104 who
also report pronounced interaction between HM-HEC and the nonionic
surfactant C12EO g as seen by fluorescence techniques. This method and also
electrical conductivity were used by Dualeh and Steiner l03 to verify the
pronounced interaction of HM-HEC and SDS.

t
Lastly, with a hydrophobicall (C 12 ) substituted cationic cellulosic
polymer, Ananthapadmanabhan et al. l 5 showed by pyrene fluorescence methods
that association of SDS and polymer occurs at a much lower concentration
than it does with a conventional cationic cellulosic, Polymer JR, that has
a much higher degree of cationic substitution. This shows again the
important effect on the association of the presence of hydrophobic groups
in the polymer.

REFERENCES

1. P. J. Flory, "Principles of Polymer Chemistry," Cornell University


Press, Ithaca, New York, 1953.
2. E. D. Goddard, Colloids and Surfaces, 19, 255, 301 (1986)
3. K. Shirahama, Colloid Polym. Sci., 252, 978 (1974).
4. M. N. Jones, J. Colloid Interface Sci., 21, 36 (1967).
5. K. Ohbu, o. Hiraishi and I. Kashiwa, J. Am. Oil Chern. Soc., 59, 108
(1982).
6. G.C. Kresheck and 1. Constantinidis, Anal. Chern., 56, 152 (1984)
7. B. J. Birch, D. E. Clarke, R. S. Lee and J. Oakes, Anal. Chim.
Acta, 70 417 (1974).
8. N. Takisawa, P.Brown, D. Bloor, D. G. Hall and E. Wyn-Jones, J. Chern.
Soc., Faraday Trans, I, 85, 2099 (1989).
9. A. Malovikova, K. Hayakawa and J. C. T. Kwak, in "Structure
Performance Relationships in Surfactants", M. J. Rosen, Editor, ACS
Symposium Series No. 253, p. 225, American Chemical Society,
Washington, D.C., 1984.
10. K. Hayakawa, J. P. Santerre and J. C. T. Kwak, Macromolecules, 16,
1642 (1983).
11. E. D. Goddard, T. S. Phillips and R. B. Hannan, J. Soc. Cosmet.
Chern., 26, 461 (1975).
12. E. D. Goddard and R. B. Hannan, J. Am. Oil Chern. Soc., 54, 561
(1977) .
13. N. Muller and T. W. Johnson, J. Phys. Chern., 11, 2042 (1969).
14. M. L. Smith and N. Muller, J. Colloid Interface Sci., 52, 507 (1975).
15. B. Cabane, J. Phys. Chern., 81, 1639 (1977).
16. Z. Gao, R. Wasylishen and J. C. T. Kwak, J. Colloid Interface Sci.,
137, 137 (1990).
17. A. Carlsson, G. Karlstrom and B. Lindman, J. Phys. Chern., 93, 3673
(1989).
18. T. Sasaki, K. Kushima, K. Matsuda and H. Suzuki, Bull. Chem. Soc.
Jpn., 53, 1864 (1980).
19. S. Saito and M. Yukawa, J. Colloid Interface Sci., 30, 211 (1969).
20. S. Saito, J. Colloid Interface Sci., 15, 283 (1960).
21. J. C. Brackman and J.B.F.N. Engberts, J. Colloid Interface Sci.,
132, 250 (1989).
22. H. Lange, Kolloid Z. Z. Polym., 243, 101 (1971)

239
23. H. Arai, M. Murata and K. Shinoda, J. Colloid Interface Sci., 37, 223
(1971).
24. W. Binana-Limbele and R. Zana, Macromolecules, 20, 1331 (1987).
25. D. Y. Chu and J. K. Thomas, J. Am. Chem. Soc., 108, 6270 (1986).
26. M. Murata and H. Arai, J. Colloid Interface Sci., 44, 475 (1973).
27. B. Cabane and R. Duplessix, J. Physique, 43, 1529 (1982).
28. A. Malovikova, K. Hayakawa and J. C. T. Kwak, J. Phys. Chem., 88, 1930
(1984).
29. K. Hayakawa and J. C. T. Kwak, J. Phys. Chem., 87, 506 (1983).
30. M. J. Schwuger, J. Colloid Interface Sci., 43, 491 (1973)
31. H. Arai and S. Horia, J. Colloid Interface Sci., 30, 372 (1962).
32. M. M. Breuer and I. D. Robb, Chem. Ind., 530 (1972).
33. F. M. Winnik, M. A. Winnik and S. Tazuke, J. Phys. Chem., 91, 594
(1987).
34. G. Kar1strom , A. Carlsson and B. Lindman, J. Phys. Chem., 94, 5005
(1990).
35. K. E. Lewis and C. P. Robinson, J. Colloid Interface Sci., 32, 539
(1970).
36. G. H. Schild and D. A. Tirrell, Polym. Prepr., Am. Chem. Soc., 30
(2), 350 (1989).
37. N. Murai, S. Makino and S. Sugai, J. Colloid Interface Sci., 41, 399
(1972) .
38. K. Hayakawa, J. P. Santerre and J. C. T. Kwak, Biophys. Chem., 17, 175
(1983).
39. K. Hayakawa and J. C. T. Kwak, J. Phys. Chem., 86, 3866 (1982).
40. R. Nagaraj an, Colloids and Surfaces, 13, 1 (1985).
41. T. Gi1anyi and E. Wolfram, Colloids and Surfaces, 1, 181 (1981).
42. D. M. Bloor and E. Wyn-Jones, J. Chem. Soc. Faraday Trans., 2, 78,657
(1982).
43. J. Gettings, C. Gould, D. G. Hall, P. L. Job1ing, J. Rassing and E.
Wyn-Jones, J. Chem. Soc. Faraday Trans., 2, 76, 1535 (1980).
44. D. M. Bloor, W. Knoche and E. Wyn-Jones, in "Techniques and Ap-
plications of Fast Reactions in Solution," NATO Adv. Study Inst. Ser.,
Ser. C. (NASCDS), VC50, p.265, D. Reidel, Dordrecht, 1979.
45. E. Ruckenstein, G. Huber and H. Hoffmann, Langmuir, 1, 382 (1987).
46. R. Nagaraj an, J. Chem. Phys., 90, 1980 (1989).
47. D. G. Hall, J. Chem. Soc. Faraday Trans., 1, 81, 885 (1985).
48. K. Shirahama, K. Tsujii and T. Takagi, J. Biochem. (Tokyo), 75, 309
(1974).
49. L. M. Lando11, Private Communication, 1979.
50. I. Satake and J. T. Yang, Biopo1ymers, 15, 2263 (1976).
51. K. Shirahama, H. Yuasa and S. Sugimoto, Bull. Chem. Soc. Jpn., 54, 375
(1981).
52. K. Shirahama and M. Tashiro, Bull. Chem. Soc. Jpn., 57, 377 (1984).
53. A. Delvil1e, Chem. Phys. Lett., 118, 617 (1985).
54. J. Francois, J. Dayantis and J. Sabbadin, Eur. Po1ym. J., 21, 165
(1985).
55. P. S. Leung, E. D. Goddard, C. Han and C. J. Glinka, Colloids and
Surfaces, 13, 47 (1985).
56. S. Saito, Kol1oid Z., 154, 19 (1957).
57. M. N. Jones, J. Colloid Interface Sci., 26, 532 (1968).
58. N. J. Turro, B. H. Baretz and P. L. Kuo, Macromolecules, 17, 321
(1984).
59. T. Isemura and A. Imanishi, J. Po1ym. Sci., 33, 337 (1958).
60. M. J. Schwuger and H. Lange, Proc. 5th Int. Congr. of Surface
Active Agents, Barcelona, (1968) Ediciones Unidas S.A., Barcelona,
Vol. 2, p.955 (1969).
61. P. L. Dubin and D. Davis, Colloids and Surfaces, 13, 113 (1985).
62. P. L. Dubin, M. E. Y. Vea, M. A. Fallon, S. S. The, D. R. Rigsbee and
L. M. Gan, Langmuir, Q, 1422 (1990).

240
63. M. Yu. Pletnev and A. A. Trapeznikov, Kolloidn. Zh., 40, 948 (1978).
64. L. D. Rhein, C. R. Robbins, K. Fernee and R. Cantore, J. Soc. Cosmet.
Chem., 37, 125 (1986).
65. F. J. Prescott, E. Hahnel and D. Day, Drug and Cosmet. Industry, 93,
443 (1963).
66. J. B. Ward and G. J. Sperandio, J. Soc. Cosmet. Chem., 15, 327
(1964).
67. J. A. Faucher, E. D. Goddard, R. B. Hannan and A. M. Kligman,
Cosmetics and Toiletries, 92, 39 (1977).
68. E. D. Goddard and R. B. Hannan, J. Colloid Interface Sci., 55, 73
(1976).
69. K. P. Clark and R. A. Falk, "Polysaccharide/Perfluoroalkyl Complexes",
Eur. Pat. Appl. E. P. 311, 570, 12 Apr. 1989; U.S. Appl. 107, 434, 9
Oct. 1987.
70. C. W. Glancy and D. R. Bassett, Polymeric Material Science and
Engineering Proc., American Chemical Society, 51, 348 (1984).
7l. J. E. Glass, Editor, "Polymers in Aqueous Media," Advances in
Chemistry, No. 223, American Chemical Society, Washington, D.C., 1989.
72. E. D. Goddard and D. B. Braun, Cosmetics and Toiletries, 100, 41
(1985).
73. A. Carlsson, G. Karlstrom and B. Lindman, Langmuir, 1, 536 (1986).
74. A. Carlsson, G. Karlstrom, B. Lindman and o. Stenberg, Colloid
Polym. Sci., 266, 1031 (1988).
75. F. Xie, C. Ma and T. Gu, Colloids and Surfaces, 36, 39 (1989).
76. K. Tha1berg, B. Lindman and G. Kar1strom, J. Phys. Chem.,95, 3370
(1991).
77. E. D. Goddard, K. P. A. Padmanabhan and P. S. Leung, Proc. 16th
Int. Fed. Soc. Cosmetic Chem., pp. 103-109, New York, 1990.
78. P. S. Leung and E. D. Goddard, Langmuir, In Press.
79. A. Carlsson, G. Karlstrom and B. Lindman, Colloids and Surfaces,
47, 147 (1990).
80. J. C. Brackman and J.B.F.N. Engberts, J. Am. Chem. Soc., 112, 872
(1990).
81. E. D. Goddard, T. S. Phillips and R. D. Hannan, J. Soc. Cosmet.
Chem., 26, 461 (1975).
82. P. Somasundaran and B. Moudgil, in "Macromolecular Solutions,"
R. B. Seymour and G. A. Stahl, Editors, p.151, Pergamon Press, New
York, 1982.
83. J. C. Brackman, N. M. van Os and J.B.F.N. Engberts,
Langmuir, ~, 1266 (1988).
84. F. M. Witte and J.B.F.N. Engberts, Colloids and Surfaces, 36, 47
(1989).
85. F. M. Winnik, Langmuir, £, 522 (1990).
86. F. M. Winnik and M. A. Winnik, Polymer J., 22, 482 (1990).
87. Y-Z. Hu, C-L. Zhao, M. A. Winnik and P. R. Sundarajan, Langmuir, £,
880 (1990).
88. A. C. Balazs and J. Y. Hu, Langmuir, ~, 1230, 1253 (1989).
89. J-L. Hsu and U. P. Strauss, J. Phys. Chem., 91, 6238 (1987).
90. D-Y. Chu and J. K. Thomas, Macromolecules, 20, 2133 (1988).
91. L-B. Shih, D. H. Mauer, C. J. Verbrugge, C.F. Wu, S. L. Chang and S.
H. Chen, Macromolecules, 21, 3235 (1988).
92. L-B. Shih, E. Y. Shen and S. H. Chen., Macromolecules, 21, 1387
(1988).
93. H. T. Oyama, W. T. Tang and C. W. Frank, Macromolecules, 20, 474
(1987) .
94. K. Char, C. W. Frank, A. P. Gast and W. T. Tang, Macromolecules, 20,
1833 (1987).
95. I. Yamazaki, F. M. Winnik, M. A. Winnik and S. Tazuke, J. Phys.
Chem., 91, 4213 (1987).
96. Y. Wang and M. A. Winnik, Langmuir, £, 1437 (1990).

241
97. L. M. Landoll, J.Polym. Sci., Polymer Chem. Ed., 20,443 (1982).
98. R. A. Gelman and H. G. Barth, in "Water Soluble Polymers," J. E.
Glass, Editor, Advances in Chemistry, No. 213, p. 101, American
Chemical Society, 1986.
99. J. W. Goodwin, R. W. Hughes, C. K. Lam, J. A. Miles and B. C.
H. Warren, Reference 71, p. 365.
100. E. Hamad and S. Qutubuddin, Macromolecules, 23, 4185 (1990).
101. D. R. Bassett, Private Communication, 1989.
102. A. C. Sau and L. M. Landoll, Reference 71, p. 343.
103. A. J. Dualeh and C. A. Steiner, Macromolecules, 23, 251 (1990).
104. K. Sivadasan and P. Somasundaran, Colloids and Surfaces, 49,
229 (1990).
105. K. P. Ananthapadmanabhan, P. S. Leung and E. D. Goddard, in "Polymer
Association Structures," M. A. El-Nokaly, Editor, ACS Symposium
Series No. 384, p. 297, American Chemical Society, Washington, D.C.,
1989.

242

You might also like