Anda di halaman 1dari 8

Journal of Colloid and Interface Science 256, 228235 (2002)

doi:10.1006/jcis.2001.8066

Polymer/Surfactant Interaction: Interfacial Aspects


E. D. Goddard1
9 Hatsawap Road, Cambridge, Maryland 21613
Received May 4, 2001; accepted October 29, 2001; published online February 5, 2002

that in a mixture of the surfactant sodium dodecyl sulfate (SDS)


and the relatively weakly surface-active polymer polyethyleneoxide (PEO), incorporation of the PEO led to two new transitions in the surface tension/logarithm of concentration (/ln c)
plot of the surfactant. Jones labeled the transitions T1 and
T2 , one below and one above the conventional critical micelle
concentration (cmc), respectively. His interpretation was that
T1 signaled the initiation of interaction between the surfactant
and the polymer (at a fixed level), and T2 , the saturation point
beyond which further addition of surfactant led to increasing
formation of regular SDS micelles. Many workers today refer to
T1 as the critical association concentration (cac). An idealized
depiction of this behavior is given in Fig. 1. A refinement is
the inclusion of a point T21 , which refers to saturation of the
polymer. Thereafter there is a buildup of monomer concentration, and hence surface tension reduction, in the T21 T2 region.
Jones own results, and subsequent work, especially by Lange
(3) on the polyvinylpyrrolidone (PVP)/SDS system, led to further modification of the model interaction scheme (see Fig. 2).
This version allows for the observation that in interacting systems in the dilute region (concentration < T1 ) the surface tension
is always below that of the surfactant solution by itself, implying that an interacting polymer is surface active in its own right.
Furthermore, there is no reason, a priori, that a surface tension
plateau exists in the T1 T21 region, and, in fact, a gradual drop
to the value at concentration T2 is most often observed. A recent
finding of great interest is that a maximum in surface tension can
exist in the T1 T21 region, for example, in the PVPSDS system
if the polymer concentration is 1% (4). (For illustration, see
the lower and upper dotted lines in Fig. 2, respectively.) These
behaviors are undoubtedly concerned with the changing composition (ratio and amounts of polymer and surfactant) of the
surface in the T1 T21 region. It should be mentioned that very
precise measurements in bulk phase using microcalorimetry are
providing evidence that the interaction patterns in comparable
systems in the T1 T2 region are themselves more complicated
than previously thought (5, 6). More work in this area can be
expected. Direct measurements on surface composition are referred to below.
A potentially important role of the surface activity of the polymer itself is implied by the above. Indeed, this may be inferred
from the listing by Breuer and Robb (7) of the reactivity

Surface methodsespecially surface tension measurements


have long been used in the study of interaction between water
soluble polymers and surfactants. Many factors control reactivity in such systems. For example, hydrophobicity (and surface
activity) of the polymer can be a dominant factor in the case of
uncharged polymer/ionic surfactant pairs, while oppositeness of
charge can be dominant in polyelectrolyte/ionic surfactant systems.
In the latter case, ideal surface behavior in the Gibbsian adsorption sense is often observed. Other surface methods used, and
briefly alluded to, include neutron reflection, surface rheology, radio
tagging, ellipsometry, X-ray reflectivity, and the study of foams. Interaction at the solid/water interface, traditionally studied in mineral dressing research, is receiving increasing attention. References
to the use of 20 different instrumental methods are given, together
with a brief discussion of the use of three of them, namely electron
spin resonance, photon correlation spectroscopy, and atomic force
microscopy. C 2002 Elsevier Science (USA)

I. INTRODUCTION

Recent years have witnessed an explosive growth in studies of


the interaction of water-soluble polymers with surfactants, both
in solution and at interfaces. It is with the latter category that
this review is concerned. To curtail length, the traditional and
still very active field of protein/surfactant interaction (1) is not
treated. The beginnings of research on the interfacial properties
of mixed synthetic water-soluble polymer, surfactant systems go
back to the 1960s: they include studies at both the air/water interface and the solid/water interface. The former category receives
emphasis here and only highlights of the rapidly expanding latter
category are presented.
II. AIR/WATER INTERFACE

A. Uncharged Polymer/Charged Surfactant


1. Surface Tension
Considerable impetus was provided to research in the field as
a whole by the surface tension studies of Jones (2). He showed

To whom correspondence should be addressed.

0021-9797/02 $35.00


C 2002 Elsevier Science (USA)

All rights reserved.

228

229

POLYMER/SURFACTANT INTERACTION

FIG. 1. Idealized surface tension/log concentration plot of a surfactant in


the presence of a complexing polymer.

(hydrophobicity) series of a group of water-soluble polymers


toward anionic surfactants, viz.,2
PVOH < PEO < MeC < PVAc PPO PVP,
the anomalously high reactivity of PVP being ascribed to its
residual positive charge.
A connection between the reactivity of the polymer and its
surface activity was proposed in the authors laboratory (8).
The effects are conveniently illustrated by data reproduced from
the first Mittal symposium. See Section B and Fig. 3, which
shows the /ln c behavior of SDS in the presence of a series
of cellulose-based water-soluble polymers. By surface tension
criteria it is seen that the feebly surface-active HEC shows
little interaction tendency with SDS, whereas with the much
more surface-active MeC, pronounced interaction is evident.
These trends were supported by bulk phase studies (dye solubilization and viscosity, reported in the same proceedings). One
should, however, report that more recent bulk phase methods,
viz., pyrene fluorescence (9) and NMR diffusion (L. Piculell,
personal communication, 1999), have shown definite premicellar interaction of HEC and SDS. This particular system needs
further investigation which should include details of the HEC
specimens employed (MW, MS, uniformity of substitution, and
so on). In passing, one should note the obviously high reactivity
toward SDS of the cationic HEC derivative, Polymer JR, seen
in Fig. 3. This aspect is elaborated on later.
Finally, mention is made of a system that apparently exhibits
the model behavior of Fig. 1 except for curvature in the T1 T21
2 PVOH, polyvinyl alcohol; PVAc, polyvinyl acetate; HEC, hydroxyethyl
cellulose; MeC, methyl cellulose; PPO, polypropylene oxide; PAAM, polyacrylamide; PNIPAM, poly(N-isopropylacrylamide).

FIG. 2.

More refined version of Fig. 1 (see text).

region (shown by the lower dotted line in Fig. 2). This behavior
was reported (10) for PNIPAM microgel/SDS systems: apparently the reactivity of the polymer toward SDS is retained in
bulk phase, leading to a clearly defined T1 transition, but not
at the air/water interface because of the limited surface activity
of the microgel particles. This behavior contrasts with a conventional response of molecular PNIPAM shown, for example, by
the work of Jean et al. (11) in which interaction of this highly
surface-active polymer and SDS was clearly evident from the
surface behavior.
In closing this section, it is noted that the feebly surface-active
polymer PAAM has been found to display virtually no reactivity

FIG. 3. Surface tension/log concentration plots of SDS alone and in the


presence of 0.1% added polymers. Inset: Surface tension plots for polymers
alone. Reproduced, with permission, from Goddard and Hannan (21).

230

E. D. GODDARD

toward SDS, just the opposite of hydrophobically modified polymers which are markedly surface active (8).
2. Other Surface Analyses, Films, Foams
In a series of papers, Thomas, Penfold, and co-workers
(1214), using specular neutron reflection techniques on PEO/
SDS and PVP/SDS systems, demonstrated in general terms that
below T1 , mixed surface films of polymer and surfactant are
present, but above T1 the polymer is displaced. Increases in surfactant concentration in the surface occur beyond T1 and can
occur to some extent beyond T2 : the actual concentration of
polymer affects the particular behavior observed. In the case of
PEO/SDS the presence of polymer tends to depress somewhat
the surface concentration of SDS: with PVP/SDS up to T1 there
is an increase and beyond it a slight decrease. A discordant note
remains on comparing these results with the radiotracer work of
Chari and Hossein in which a 30% reduction in surface concentration of SDS was observed in the presence of PVP (15), but,
in general terms, the above trends are in accord with neutron reflectance data on PNIPAM/SDS systems, although differences
in degree result from the much higher surface activity of this
polymer versus PVP or PEO (11).
The case of adsorption in a mixed surfactant uncharged polymer system has been analyzed thermodynamically by de Gennes
(16). If the polymer adsorbs one can expect a change in adsorption of the surfactant and of its foaming and thin-film properties,
including drainage rate, when the layers of polymer overlap and
the surface viscosity changes (increases). The result would be
more stable films and foams (see Fig. 4). Testing to date of these
concepts has been limited. In film drainage studies on a series
of polymer/surfactant systems Cohen-Addad and diMeglio (17)
found rather small effects of the polymer even in the case of
the strongly interacting PEO/SDS system. It should be noted,
however, that the concentration of SDS was well above the cmc.
On the other hand, in very recent work, involving foam and
single-film studies by Folmer and Kronberg (4) on the PVP/SDS
system, definite improvements in stability were found up to a
surfactant concentration approximating T21 ; beyond this, progressive removal of polymer from the interface rendered any
improvement negligible (or even negative). More work in this
important area can be expected (18).
B. Charged Polymer/Oppositely Charged Surfactant
1. Surface Tension
It was shown, evidently for the first time, by the author and
co-workers in 19751976 (1921) and reported at the 1977
Mittal symposium, that the surface tension behavior of the title
systems can be quite different from those involving uncharged
polymers (see Fig. 3). As observed in bulk phase, interaction
at the surfaceevident from pronounced lowering of surface
tensioncan be detected at very low concentrations of surfactant (104 M or less) in the cationic cellulosic Polymer
JR/SDS system. Even in the precipitation zone, when most of the

FIG. 4. Drawing of a draining lamella formed from a mixed polymer/


surfactant solution.

surfactant and polymer are out of solution, the surface tension


remains low. This indicates that the complexes that form are
extremely surface active. The situation in solution and at the
interface is depicted in Fig. 5 (20, 22). Once micelles of surfactant abound in solution, polymer is stripped from the interface
and the surface tension becomes essentially equal to that of a
micellar, polymer-free surfactant solution. In broad terms this
picture has gained support from recent studies on the neutron
reflectance characteristics of gelatin/SDS systems (23). Further
support for the above depiction (Fig. 5) of the mixed component
surface was obtained from a monolayer penetration experiment in which the alkyl sulfate was confined to the surface, viz.,
by employing an insoluble homolog (docosyl sulfate) (20). See
Fig. 6. In this experiment (apparently the first reported for a
charged insoluble surfactant monolayer spread on a solution
of an oppositely charged polymeric electrolyte), although the
subphase contained only a very low concentration (10 ppm) of
the polymer its presence totally changed the surface pressure
area curve. While it is likely that the observed expansion would
have been found at an even lower concentration of the polyelectrolyte, e.g., 1 ppm, time effects due to diffusion of the polyion
would probably have been much more serious.3 What can be
3 Monolayer expansion effects by oppositely charged polyelectrolyte similar
to that described above have recently been reported by several other workers
(24, 25).

231

POLYMER/SURFACTANT INTERACTION

FIG. 5. Conditions in bulk and surface of a solution of a polycation (fixed


concentration) and an anionic surfactant. Solid line: hypothetical surface tension curve of the surfactant alone; dashed line: in mixture with the polycation.
Reproduced, with permission, from Goddard (22).

deduced is that the strongly adsorbed polyion imposes a new


packing arrangement (expansion) of the monolayer, lending direct support to the surface scheme sketched in Fig. 5. A further
possibility is that the transition noted at an area/molecule value
2 represents the limiting area of the surfactant ion in the
of 60 A
unperturbed surface complex. The high stability of the monolayer, evident from the high collapse pressure, is noteworthy,
as were indications of the development of considerable surface
viscosity.
An important development is represented in the surface tension study of SDS and a range of polylysine homopolymers
(MW 500070,000) by Buckingham et al. (26). Working at SDS
concentrations below the precipitation range they observed linear characteristics of the /ln c plots which were essentially
independent of polymer concentration and molecular weight
(above low limiting values). Furthermore, they showed that the
Gibbs adsorption equation for their systems reduces to the very
simple form

Returning to the Polymer JR/SDS system, Goddard and


Hannan found that it was not possible to apply the above equation to their data because of (a) severe time effects noted in
some of the surface tension points (requiring more than 12 day
for equilibrium and (b) clear indications that changes of polymer concentration had a definite effect on the observed /ln c
plots (20). In contrast, model behavior for the system LAS
(tridecylbenzenesulfonate)/Polymer JR is suggested in the observed linear /ln c plot, preceding a surface tension plateau
(19): calculated areas of these wing-shaped surfactant molecules
2 (alone) and 150 A
2 (in the complex) (see below).
are 100 A
In a major program (2731) on the surface behavior of
mixtures of various polyanions and cationic surfactants (notably, dodecyltrimethylammonium bromide (DTAB)) Langevin
and co-workers have emphasized four systems: (a) polyacrylamide, polyacrylamidesulfonate (PAMPS) containing 10
and 25% charged monomer; (b) polystyrenesulfonate (PSS);
(c) DNA; (d) xanthan. In all cases the presence of the polyanion
led to a substantial reduction in surface tension of the surfactant.
The PAMPS/DTAB system has been studied in greatest detail
(27). For PAMPS with 25% charge they obtained data similar
to those of Buckingham et al. inasmuch as, over a range of
concentrations (75750 ppm) of added polyanion, the observed
/ln c plot was linear and independent of polyanion level. Also,
a preprecipitation surface tension plateau similar to that found
by Goddard and Hannan (20) was observed. Application of the
equation
d =  RT d ln[DTAB]
2 /mol of DTA+ ,
to the linear part of the plot led to an area of 78 A
2

far in excess of the value of 45 A /mol obtained from the Gibbs

d =  RT d ln[SDS].
The net aggregate area/molecule (1/ ) of the surfactant ion
computed from the above equation, for all molecular weights of
2 /mol, to be almost the same as
polylysine, turned out, at 43 A
that of SDS in simple aqueous solution, possibly reflecting the
flexibility and high charge density of this polymer. It should be
noted that the assumptions made in deriving the above equation
are that the surface excess values of all the simple ions present
(Na+ , Br ) approximate zero and that variations in the actual
concentration of the polyion present are, for practical purposes,
without effect.

FIG. 6. Surface pressure/area isotherms of sodium docosyl sulfate (SDCS)


spread on 0.1 M NaCl and on 0.1 M NaCl plus 0.001% Polymer JR 400. Inset:
2 on latter
versus square root of time after spreading monolayer at A = 100 A
solution. Reproduced, with permission, from Goddard and Hannan (20).

232

E. D. GODDARD

FIG. 7. Effect of PAMPS on DTAB surface tension. Open circles, DTAB;


diamonds, DTAB + 75 ppm PAMPS; crosses, DTAB + 350 ppm PAMPS;
closed circles, DTAB + 750 ppm PAMPS. Reproduced, with permission, from
Asnacios et al. (27).

plot for DTAB alone, again suggesting that the macrocounterion


imposes a new spacing on the adsorbed surfactant ions. The
point at which the linear portion of the / ln c plot ends has
been labeled the cac by these authors (27). See Fig. 7.
On the other hand, for the PSS/DTAB system no such independence of the /ln c plots of polymer concentration, no obvious
linearity of the plots, and no constancy of apparent cac values
were found. An explanation of this behavior (28) was offered
that invokes the hydrophobicity of the PSS macromolecule but it
is pointed out that similar lack of conformity had been obtained
by Goddard and Hannan for the very hydrophilic polycation,
Polymer JR, in combination with SDS. While this equation is
strictly valid only if it is demonstrated that the linear portion of
the /ln c plot is independent of polymer concentration,4 the
data led to the following areas of the DTA ion in the presence
2 (DNA), 80 A
2 (xanthan).
of other polyanions: 70 A
It should be mentioned that extensive studies showing the
marked surface tension depression of a charged surfactant
by an oppositely charged polyelectrolyte have been reported
by Stenius and co-workers who have employed both anionic
(32) and cationic (33) polysaccharide derivatives. Work by
Manuszak Guerrini and coinvestigators (34) on cationic polysaccharide/SDS systems and by Desbri`eres and Rinaudo (35) on
chitosan carboxylate/TAB systems should also be cited.
2. Other Surface Analysis, Films, Foams
The surface tension data referred to above, and other data
given below, provide clear evidence that in mixed solutions of
polyion and oppositely charged surfactant the macroion is intimately associated with the surfactant at the interface (except
4 This independence has recently been shown for the xanthan/DTAB system
in the polymer concentration range 60170 ppm, but irregularities appear at
350 ppm (D. Langevin, personal communication, 2000).

at high supramicellar concentrations. See Fig. 5.) Hence, it is


desirable to obtain more information concerning the polymer
at the surface. Several methods have been employed. For example, Asnacios et al. (27) determined by ellipsometry that the
optical thickness of the surface of a simple DTAB solution is
and those of DTAB/PAMPS systems before the cac are
10 A,
(40 A)
and 200 A
(50 A)
for the 30 and 10% charged
80 A
polymer, respectively. The authors suggest, in the latter case, a
more loopy configuration in which the polymer may not be
replacing all of the Na counterions at the interface.
Another method employed by Langevin and co-workers (28)
to determine the thickness of the surface layer is X-ray reflectivity. In all cases this technique, which measures the electron density perpendicular to the surface, led to measured thicknesses of
for the polyion/surfactant systems; the surface reflectiv20 A
ity of the surfactant (DTAB) alone was the same as that of water.
The authors conclude that this technique sees only the polymer train segments and short loops. It should be mentioned
here that surface X-ray diffraction measurements on monolayers of (negatively charged) dipalmitoyl phosphatidic acid
(DPPA) spread on a subsolution with and without the polyelectrolyte polydiallyldimethylammonium chloride (PDADMAC)
have been reported (25). Clear evidence of binding of this polycation to the monolayer molecules was seen through the observed chain tilting and surface lattice expansion effects.
In the early work of Goddard and Hannan (20) it was reported that substantial enhancement of the foaming of SDS solutions could be brought about by the addition of Polymer JR in
amounts sufficient to bring the system close to or into the precipitation zone of the phase diagram. In all cases the conditions
were such that low surface tension values (45 mN/m) were
realized. These results were essentially confirmed for this and
two related cationic cellulosic polymers (36).
Though foaming is recognized to be a complex phenomenon,
influenced by a variety of factors, early studies of the SDS/water
system showed that the addition of trace quantities of highly
surface-active dodecanol to the systems could cause the formation of viscous surface layers, and promote slowness in drainage
of generated soap films together with increased foam stability
(37). The observation (20) of viscoelasticity in monolayers of
C22 SO4 caused by a small concentration (10 ppm) of polycation
in the subphase suggested that there might be a parallel mechanism in which the high surface activity of the polycation, due
to strong electrostatic forces, could lead to network or lattice
formation (25) in an adsorbed layer of surfactant such as SDS.
Subsequent qualitative testing of surface rheology using gentle air current displacement of talc particles sprinkled on the surface (38, 39) confirmed these expectations. (The test is based on
procedures developed by Lord Rayleigh (40).) The surface phase
map showed wide areas of viscoelasticity development including
much of the preprecipitation and precipitation zones in Polymer
JR/C12,10,8 SO4 systems. Furthermore, in limited testing, singlefilm reflectance tests showed the formation of slow-draining
vertical films formed from mixed Polymer JR/SDS systems

POLYMER/SURFACTANT INTERACTION

233

FIG. 9. Coadsorption of a surfactant and a polymer at the air/water,


oil/water, or solid/water interface.

FIG. 8. Drainage films (pictorial) for SDS (1%) observed after 5 s (a)
and after 30 s (b) and for SDS (0.01%)/Polymer JR 400 (0.01%) after 30 s
(c). Reproduced, with permission, from Regismond et al. (39).

versus fast-draining films observed from SDS solutions alone


(see Fig. 8). Because of the high molecular weight ( 0.5 106 )
of this polycation one is aware that its extended molecular length
could approximate the thickness of such a draining film. As regards foams themselves, it was found that wider compositional
areas of foam stability in the shorter-chain-length alkyl sulfate
(C8 , C10 )/Polymer JR systems than in the C12 SO4 systems exist,
possibly indicating a type of optimized HLB requirement in
these systems.
Recently a variety of approaches have been employed by
Langevin and co-workers to examine the surface and thin-film
properties of several anionic polyelectrolytes in combination
with a cationic (usually DTAB). The techniques included disjoining pressure measurements (29, 31) on thin films stabilized
by PAMPS/DTAB and PSS/DTAB combinations. (Interestingly,
stable films could not be obtained with the DNA and xanthan
systems.) The authors propose a polymer crosslinking process in
the stabilizing mechanism. Film thicknesses in the range 300
were observed, together with time-dependent jumps in
800 A
thickness and heterogeneities in the films themselves.
Another approach employed by this group involves the measurement of surface viscoelasticity by a capillary wave technique
(27). The presence of the polyion (PAMPS and xanthan) causes
a substantial increase in the surface viscoelasticity of DTAB
solutions, peaking below the cmc of DTAB by itself (41). The
authors also carried out foam stability tests on these systems.
Though increases in foam height were found in certain zones
in the mixed systems over that of the surfactant alone, no simple correlations with the measured surface viscoelasticity were
found, confirming the complexity of the foaming phenomenon
found by all previous workers.
III. SOLID/WATER INTERFACE

Figure 9 represents one-half of the thin film shown in Fig. 4.


It depicts an uncharged water-soluble polymer coadsorbed with

a surfactant at the interface between water and a low-polarity


medium (air, oil, or a solid). The depiction in the last case (S/W)
is obviously overrestrictive and simplistic since coadsorption
can also take place at hydrophilic solid surfaces, in which case
the orientation (up or down) of the surfactant molecules would
depend on factors such as charge effects.
An illustration of the coadsorption of a cationic polymer
on (negatively charged) quartz in the presence of a cationic
surfactant and an anionic surfactant is given in Fig. 10 (49).
Coadsorption of a polycation in case a depresses the floatability
(hydrophobicity) conferred by the cationic surfactant alone. On

FIG. 10. Polycation adsorption on a negatively charged solid. Top (a): with
a cationic surfactant (depressed flotation); bottom (c): with an anionic surfactant (elevated flotation). Reproduced, with permission, from Somasundaran
et al. (49).

234

E. D. GODDARD

TABLE 1
Experimental Techniques (Solid/Liquid)
Adsorption analysis (43)
Atomic force microscopy (44)
Calorimetry (45)
Ellipsometry (46)
Electron spin resonance (47)
Fluorescence (48)
Flotation/wetting (42, 49)
Infrared (50)
Microscopy (51)
Neutron reflection (52)

Neutron scattering (53)


Nuclear magnetic resonance (54)
Photon correlation spectroscopy (53)
Radiotracer (55)
Rheology (56, 58)
Solubilization (Ad-) (57)
Suspension stability (56, 58)
Surface forces balance (46)
Ultrafiltration (59)
Potential (49)

the other hand, coadsorption of a polycation and a (normally


nonadsorbing) surfactant of opposite charge can activate flotation (case c). A practical example of Fig. 9-type behavior has
been given by Kilau and Voltz (42) in which wetting of a hydrophobic surface (coal) by an anionic surfactant is promoted
by the addition of the uncharged polymer PEO.
Early studies on polymer/surfactant interaction on solid surfaces were concerned largely with analytical determination of
the extent of adsorption (8). By choice of conditions (order
of addition, absolute and relative concentrations, etc.) it was
frequently found possible to obtain elevated uptake of one or
other of the components. The upsurge in interest in this field
in recent years has led to increasing variety and sophistication in the instrumental methods used for the studies and, in
turn, to a substantial increase in literature on the subject. For
space economy we list here 20 techniques that have been employed together with a single reference (in most cases) for each
category, as an opening to the wider literature on the subject
(see Table 1).
By way of a conclusion, three papers are chosen for a brief
elaboration of representative results that have been reported using a few of the above techniques. The first by Otsuka and Esumi
(47), dealing with a coadsorption study, shows a remarkable increase in uptake of PVP on alumina in the presence of LiDS. A
maximum in PVP adsorption occurs near the T1 concentration
of surfactant (see Fig. 11). ESR measurements revealed that the
conformation of adsorbed PVP molecules changed from loops
and tails increasingly to trains as the LiDS concentration increased toward T2 . The polymer was found to have little effect
on the adsorption of LiDS.
A second illustration, by Cosgrove et al. (54), concerns the
effect of SDS on the adsorption of PEO by silica. PEO adsorbs strongly on silica, and SDS hardly at all. In this case
PCS measurements were used to determine the thickness of the
adsorbed layer. SDS is seen to bring about a progressive thinning of the PEO layer, reaching a limit around the cmc of the
surfactant (Fig. 12). What is surprising is that beyond the cmc
the layer thickness increases again, suggesting that SDS micelles
are bound by the residually attached PEO chains. In general
terms these results were confirmed by similar measurements
on polystyrene latex particles on which the PEO was either

FIG. 11. Adsorption of LiDS and PVP on alumina from PVPLiDS mixed
solution. Reproduced, with permission, from Otsuka and Esumi (47).

adsorbed or grafted (53, 54), and also by SANS measurements


(54).
Finally, recent AFM studies on the graphite/PVP/SDS system are cited (44). This system had previously been shown
(61) to conform to the pattern shown in Fig. 11 in which addition of surfactant leads to an initial increase and then to a
decrease in PVP adsorption. The AFM data suggest that the
system may be different and more complicated than originally
supposed. The presence of the polymer seems to cause disruption of part of the array of hemicylindrical SDS admicelles;
this disordered array persists above the cac (T1 ) and substantial time effects are observed as have been noted by many other
workers for solid/water systems (47, 50, 62). These results can
be expected to stimulate even more vigorous research in this
field.

FIG. 12. Hydrodynamic thickness (nm) of adsorbed layers of PEOSDS on


silica as a function of SDS concentration. Reproduced, with permission, from
Cosgrove et al. (60).

POLYMER/SURFACTANT INTERACTION

IV. CONCLUSIONS

Though somewhat idealistic in its original application to these


systems, the surface tension method retains its usefulness as an
indicator of the interaction behavior of uncharged water-soluble
polymers and ionic surfactants. The reactivity of the polymer
correlates with its own surface activity. Applied to combinations of oppositely charged surfactants and polyelectrolytes, the
method is a very sensitive indicator of interaction. Ideal behavior, in the Gibbs adsorption equation sense, is found in many
systems: failure to conform in others may be due to the formation of viscous/thick surface films.
Important information on polymer/surfactant interaction at
the air/water interface is currently being provided by a number of
other techniques, such as neutron and X-ray reflection, ellipsometry, disjoining pressure measurements, and surface rheology.
Polymer/surfactant interaction at solid/water interfaces,
where each component can profoundly influence the adsorption of the other, is now a burgeoning field of research. Twenty
different methods of study are referenced. Attention is drawn
to the increasing awareness of slow equilibrium effects in these
systems.
REFERENCES
1. Jones, M. N., in Surface Activity of Proteins (S. Magdassi, Ed.),
pp. 237284. Marcel Dekker, New York, 1996.
2. Jones, M. N., J. Colloid Interface Sci. 23, 36 (1967).
3. Lange, H., Kolloid - Z. Z. Polym. 243, 101 (1971).
4. Folmer, B. M., and Kronberg, B., Langmuir 16, 5987 (2000).
5. Torn, L. H., de Keizer, A., Koopal, L. K., and Lyklema, J., Colloids Surf.
160, 237 (1999).
6. Ghoreishi, S. M., Fox, G. A., Bloor, D. M., Holzwarth, J. F., and
Wyn-Jones, E., Langmuir 15, 5474, (1999).
7. Breuer, M. M., and Robb, I. D., Chem. Ind. 530535 (1972).
8. Goddard, E. D., and Ananthapadmanabhan, K. P., Interactions of Surfactants with Polymers and Proteins, CRC Press, Boca Raton, FL, 1993.
9. Sivadasan, K., and Somasundaran, P., Colloids Surf. 49, 229 (1990).
10. Abuin, E., Leon, A., Lissi, E., and Varas, J. M., Colloids Surf. 147, 55
(1999).
11. Jean, B., Lee, L.-T., and Cabane, B., Langmuir 15, 7585 (1999).
12. Cooke, D. J., Dong, C. C., Lu, J. R., Thomas, R. K., Simister, E. A., and
Penfold, J., J. Phys. Chem. B 102, 4912 (1998).
13. Purcell, I. P., Lu, J. R., Thomas, R. K., Howe, A. M., and Penfold, J.,
Langmuir 14, 1637 (1998).
14. Purcell, I. P., Thomas, R. K., Penfold, J., and Howe, A. M., Colloids Surf.
94, 125 (1995).
15. Chari, K., and Hossain, T. Z., J. Phys. Chem. 95, 3302 (1991).
16. de Gennes, P. G., J. Phys. Chem. 94, 8407 (1990).
17. Cohen-Addad, S., and di Meglio, J.-M., Langmuir 10, 773 (1994).
18. La Mesa, C., Colloids Surf. 160, 37 (1999).
19. Goddard, E. D., Phillips, T. S., and Hannan, R. B., J. Soc. Cosm. Chem. 26,
461 (1975).
20. Goddard, E. D., and Hannan, R. B., J. Colloid Interface Sci. 55, 73
(1976).
21. Goddard, E. D., and Hannan, R. B., in Micellization, Solubilization,
and Microemulsions (K. L. Mittal, Ed.), Vol. 2, pp. 835845. Plenum,
New York, 1977.
22. Goddard, E. D., Colloids Surf. 19, 301 (1986).

235

23. Cooke, D. J., Dong, C. C., Thomas, R. K., Howe, A. M., Simister, E. A.,
and Penfold, J., Langmuir 16, 6546 (2000).
24. Hwang, M.-J., and Kim, K., Langmuir 15, 3563 (1999).
25. de Meijere, K., Brezesinski, G., and Mohwald, H., Macromolecules 30,
2337 (1997).
26. Buckingham, J. H., Lucassen, J., and Hollway, F., J. Colloid Interface Sci.
67, 423 (1978).
27. Asnacios, A., Klitzing, R., and Langevin, D., Colloids Surf. 167, 189 (2000).
28. Stubenrauch, C., Albouy, P.-A., von Klitzing, R., and Langevin, D.,
Langmuir 16, 3206 (2000).
29. Von Klitzing, R., Espert, A., Asnacios, A., Hellweg, T., Colin, A., and
Langevin, D., Colloids Surf. 149, 131 (1999).
30. Asnacios, A., Langevin, D., and Argillier, J.-F., Macromolecules 29, 7412
(1996).
31. Bergeron, V., Langevin, D., and Asnacios, A., Langmuir 12, 1550 (1996).
32. Barck, M., and Stenius, P., Colloids Surf. 89, 59 (1994).
33. Merta, J., and Stenius, P., Colloids Surf. 149, 367 (1999).
34. Manuszak Guerrini, M., Negulescu, I. I., and Daly, W. H., J. Appl. Polym.
Sci. 68, 1091 (1998).
35. Desbri`eres, J., and Rinaudo, M., in Polysaccharide Applications (M. A.
El-Nokaly and H. A. Soini, Eds.), ACS Symp. Ser. 737, p. 199. Am. Chem.
Soc., Washington, DC, 1999.
36. Manuszak Guerrini, M., Lochhead, R. Y., and Daly, W. H., Colloids Surf.
147, 67 (1999).
37. Goddard, E. D., and Ananthapadmanabhan, K. P., in Polymer Surfactant
Systems (J. C. T. Kwak, Ed.), pp. 2164. Marcel Dekker, New York, 1998.
38. Regismond, S. T. A., Winnik, F. M., and Goddard, E. D., Colloids Surf.
119, 221 (1996).
39. Regismond, S. T. A., Winnik, F. M., and Goddard, E. D., Colloids Surf.
141, 165 (1998).
40. Lord Rayleigh, Proc. Roy. Soc. London 48, 127 (1890).
41. Bhattacharyya, A., Monroy, F., Langevin, D., and Argillier, J.-F., Langmuir,
in press.
42. Kilau, H. W., and Voltz, J. I., Colloids Surf. 57, 17 (1991).
43. Ma, C., Colloids Surf. 16, 185 (1985).
44. Fleming, B. D., Wanless, E. J., and Biggs, S., Langmuir 15, 8719 (1999).
45. Bury, R., Desmazi`eres, B., and Treiner, C., Colloids Surf. 127, 113 (1997).
46. Shubin, V., Langmuir, 10, 1093 (1994).
47. Otsuka, H., and Esumi, K., Langmuir 10, 45 (1994).
48. Esumi, K., and Meguro, K., J. Colloid Interface Sci. 129, 217 (1989).
49. Somasundaran, P., and Cleverdon, J., Colloids Surf. 13, 73 (1985).
50. Duffy, D. C., Davies, P. B., and Creeth, A. M., Langmuir 11, 2931
(1995).
51. Sjoberg, M., Bergstrom, L., Larsson, A., and Sjostrom, E., Colloids Surf.
159, 197 (1999).
52. Lu, J. R., Su, T. J., Thomas, R. K., and Penfold, J., Langmuir 14, 6261
(1998).
53. Mears, S. J., Cosgrove, T., Obey, T., Thompson, L., and Howell, I., Langmuir
14, 4997 (1998).
54. Cosgrove, T., Mears, S. J., Obey, T., Thompson, L., and Wesley, R. D.,
Colloids Surf. 149, 329 (1999).
55. Arnold, G. B., and Breuer, M. M., Colloids Surf. 13, 103 (1985).
56. Shimabayashi, S., Uno, T., and Nakagaki, M., Colloids Surf. 123/124, 283
(1997).
57. Esumi, K., Mizuno, K., and Yamanaka, Y., Langmuir 11, 1571 (1995).
58. Otsuba, Y., Langmuir 10, 1018 (1994).
59. Guo, W., Uchiyama, H., Tucker, E. E., and Scamehorn, J. F., Colloids Surf.
123/124, 695 (1997).
60. Cosgrove, T., Mears, S. J., Thompson, L., and Howell, I., in Surfactant
Adsorption and Surface Solubilization (R. Sharma, Ed.), ACS Symp. Ser.
615, p. 196. Am. Chem. Soc., Washington, DC, 1995.
61. Otsuka, H., and Esumi, K., J. Colloid Interface Sci. 170, 113 (1995).
62. Dedinaite, A., Claesson, P. M., and Bergstrom, M., Langmuir 16, 5257
(2000).

Anda mungkin juga menyukai