You are on page 1of 15

10

Microstructure E¡ects on Transport in


Reverse Microemulsions
JOHN TEXTER Strider Research Corporation, Rochester, New York

I. INTRODUCTION

Isotropic single-phase microemulsions of oil and water can adopt a variety of microstruc-
tures, including water-in-oil droplets (reverse microemulsions), swollen oil-in-water
micelles, and irregular bicontinuous microstructures of low to zero mean curvature with
interpenetrating oil and water domains separated by a monolayer of surfactant. Various
field variables can drive transitions from one of these microstructures to another. The
transport of charge, ions, and molecular species as monitored by electrical conductivity or
faradaic electron transfer and by NMR self-diffusion measurements, respectively, can be
used to experimentally define an order parameter for quantitatively tracking the onset of
microstructure formation and the amount of a particular microstructure present at equili-
brium under a given set of field variables. Transitions from isolated reverse micelles in
reverse microemulsions through the formation of percolating clusters to the formation of
irregular bicontinuous microstructure are reviewed.
Electrochemistry [1], synthesis [2], and catalysis [3,4] in reverse microemulsions [5]
are significantly affected by microstructure and the effects of microstructure on ionic and
molecular transport. Such transport can occur by several distinct mechanisms. The first
involves diffusion of ions and molecules through the oil (pseudocontinuous) phase of the
reverse microemulsion. This kind of transport is encountered in simple solutions, except
that additional factors affecting viscosity and tortuosity come into play. A second mechan-
ism involves the diffusion of reverse micelles with the concomitant transport of solutes
contained therein. Such diffusive transport is generally more than an order of magnitude
slower than the first mechanism. A third mechanism of transport involves ion and mole-
cular transport through reverse micelles (droplets) and exchange between such droplets
[6,7]. This third mechanism has most often been probed by percolation in electrical con-
ductivity [8], in faradaic electron transfer [9], in dielectric susceptibility [10], and in self-
diffusion measured by NMR pulsed gradient spin-echo (PGSE) techniques [11,12].

A. Percolation
Percolation in microemulsions and concomitant microstructural changes are the focal
points of this review. A complete understanding of percolation phenomena in reverse
microemulsions will require an understanding of droplet interactions and the associated
thermodynamics of droplet fusion, fission, aggregation to form clusters of varying fractal
241

Copyright © 2001 Marcel Dekker, Inc.


242 Texter

dimension, and the transformation of droplets and clusters of droplets to irregular bicon-
tinuous [13] microstructures of low to zero mean curvature. This last transformation has
generated some controversy in the literature. The observation of percolation has at times
been equated with bicontinuous microstructure formation [14]. Alternatively, percolation
is equated with droplet aggregation and clustering [15,16]. A prime objective of this review
is to illustrate how to experimentally distinguish between these alternative microstructures,
each of which appears attendant to percolation-related transitions.

B. Disperse Phase Volume Fraction


Electrical conductivity is an easily measured transport property, and percolation in elec-
trical conductivity appears a sensitive probe for characterizing microstructural transfor-
mations. A variety of field (intensive) variables have been found to drive percolation in
reverse microemulsions. Disperse phase volume fraction has been often reported as a
driver of percolation in electrical conductivity in such microemulsions [17–20].
Lagües et al. [17] found that the percolation theory for hard spheres could be used to
describe dramatic increases in electrical conductivity in reverse microemulsions as the
volume fraction of water was increased. They also showed how certain scaling theoretical
tools were applicable to the analysis of such percolation phenomena. Cazabat et al. [18]
also examined percolation in reverse microemulsions with increasing disperse phase
volume fraction. They reasoned the percolation came about as a result of formation of
clusters of reverse microemulsion droplets. They envisioned increased transport as arising
from a transformation of linear droplet clusters to tubular microstructures, to form
wormlike reverse microemulsion tubules.

C. Temperature
Temperature is also an often-studied field variable in driving percolation and associated
microstructural transitions [21–25]. Jada et al. [21] found that the percolation induced by
temperature increases in a water, decane, AOT reverse microemulsion could be nicely
correlated with the bimolecular rate constant for solute exchange between colliding dro-
plets. Kim and Huang [22] found that temperature-driven percolation in reverse micro-
emulsions occurred when the effective droplet volume as defined by a hopping range
exceeded the hard-sphere close-packing limit of 0.65. Peyrelasse and Boned [23] showed
that temperature-driven percolation yielded similar critical exponents to those obtained in
scaling analyses of disperse phase volume-fraction-driven percolation. An extensive study
by Alexandridis et al. [24] suggests that temperature-driven percolation involves enthalpi-
cally disfavored clustering of droplets. The net driving force emanates from positive
entropic changes that increase with temperature and that are attributed to free volume
dissimilarity between the surfactant tails and the respective organic solvents.

D. Chemical Potential
The last, and less extensively studied field variable driving percolation effects is chemical
potential. Salinity was examined in the seminal NMR self-diffusion paper of Clarkson et
al. [12] as a component in brine, toluene, and SDS (sodium dodecylsulfate) microemul-
sions. Decreasing levels of salinity were found to be sufficient to drive the microemulsion
microstructure from water-in-oil to irregular bicontinuous to oil-in-water. This paper was

Copyright © 2001 Marcel Dekker, Inc.


Microstructure Effects on Transport 243

one of the first to demonstrate the power of PGSE NMR determinations of self-diffusion
coefficients in resolving microstructure in surfactant systems.
The influence of alcohols as cosurfactants also falls in this chemical potential cate-
gory. Lang et al. [26] found that a variety of alkanols and benzyl alcohol tended to
suppress percolation in reverse microemulsions otherwise induced by disperse phase
volume fraction increases. They studied water, chlorobenzene, TBDAC (tetradecylbenzyl-
dimethylammonium chloride) reverse microemulsions and found that longer-chained alco-
hols retarded the onset of percolation to a greater extent. Somewhat similar retarding
effects of alcohols have more recently been reported by Nazário et al. [27]. At fixed
disperse phase volume fraction, increasing straight-chain alcohol (cosurfactant) concen-
tration suppressed percolation until higher temperatures were reached, relative to the
temperature at which percolation was driven in the absence of cosurfactant. It was sur-
mised that the physical effect of added alcohol was to decrease interfacial fluidity and
hypothesized that the alcohols intercalate between AOT tail groups, thereby increasing the
alkyl chain packing density. This decrease in fluidity then necessitates a higher temperature
for overcoming the enthalpic barrier to cluster formation and ensuing percolation.
Electrochemical redox studies of electroactive species solubilized in the water core of
reverse microemulsions of water, toluene, cosurfactant, and AOT [28,29] have illustrated a
percolation phenomenon in faradaic electron transfer. This phenomenon was observed
when the cosurfactant used was acrylamide or other primary amide [28,30]. The oxidation
or reduction chemistry appeared to switch on when cosurfactant chemical potential was
raised above a certain threshold value. This switching phenomenon was later confirmed to
coincide with percolation in electrical conductivity [31], as suggested by earlier work from
the group of Francoise Candau [32]. The explanations for this amide-cosurfactant-induced
percolation center around increases in interfacial flexibility [32] and increased disorder in
surfactant chain packing [33]. These increases in flexibility and disorder appear to lead to
increased interdroplet attraction, coalescence, and cluster formation.
Another example of chemical-potential-driven percolation is in the recent report on
the use of simple poly(oxyethylene)alkyl ethers, Ci Ej , as cosurfactants in reverse water,
alkane, and AOT microemulsions [27]. While studying temperature-driven percolation,
Nazário et al. also examined the effects of added Ci Ej as cosurfactants, and found that
these cosurfactants decreased the temperature threshold for percolation. Based on these
collective observations one can conclude that linear alcohols as cosurfactants tend to
stiffen the surfactant interface, and that amides and poly(oxyethylene) alkyl ethers as
cosurfactants tend to make this interface more flexible and enhance clustering, leading
to more facile percolation.

E. Microstructural Transitions
Self-diffusion NMR measurements have been shown to be very useful in characterizing
microstructural transitions in microemulsions [9,34]. Much uncertainty remains, however,
in describing transitions from oil-in-water to irregular bicontinuous to water-in-oil micro-
emulsions. Irregular bicontinuous microstructure of low to zero mean curvature has often
been inferred on the basis of intermediate to large magnitudes of self-diffusion coefficients
[35]. However, it has recently been shown that diffusion coefficients of such magnitude
arise naturally in reverse microemulsions as a result of a dynamic partitioning of various
solutes (e.g., water and cosurfactant) between continuous and disperse pseudophases [36].
It will be illustrated here that quantitative consideration of such partitioning provides
means to quantify effects on transport attendant to percolation processes.

Copyright © 2001 Marcel Dekker, Inc.


244 Texter

The remainder of this chapter develops this quantitative treatment of percolation as


it relates to self-diffusion data from NMR PGSE experiments and to an overarching
qualitative treatment also derived from self-diffusion data and from conductivity data.
This qualitative treatment focuses upon the coincidence of the upturn in water proton self-
diffusion with the onset of percolation in electrical conductivity. This treatment also
focuses upon observations that the upturn in surfactant self-diffusion usually occurs at
a higher field variable than do the water proton self-diffusion increase and the onset of
electrical conductivity percolation. The fact that these processes occur at different field
variables is used to qualitatively distinguish reverse microemulsion droplet aggregation
and clustering of percolating droplets from the onset of irregular bicontinuous structure
formation A more quantitative approach is based upon a two-state analysis of the NMR
water proton self-diffusion data. This two-state analysis is used to derive an order para-
meter for percolating cluster formation, and this order parameter is applied to the analysis
of percolation driven by disperse phase volume fraction, temperature, and cosurfactant
chemical potential.

II. VOLUME-FRACTION-INDUCED PERCOLATION

The influence of disperse phase volume fraction and of temperature on driving micro-
structure transitions in a brine (0.6% aqueous NaCl), decane, and AOT (12% w/w)
microemulsion system was studied extensively by Chen et al. [20]. They used small-
angle neutron scattering to measure the mean curvature over an extensive range of dis-
perse phase volume fraction and over a range in temperature. The Shinoda diagram for
this system is illustrated in Fig. 1, where  denotes the weight ratio of decane to decane
plus brine. The quantity 1   denotes the weight ratio of brine to brine plus decane, and

FIG. 1 Partial phase diagram of brine, decane, and AOT system as a function of temperature (T)
and decane-to–brine weight fraction (). The brine is aqueous 0.6% (w/w) NaCl; the AOT composi-
tion is constant at 12% (w/w). The double-ended arrow depicts the isothermal composition range
examined in this study at 45 C. The lamellar (L ), and two-phase regions (2; 2 ) are described in the
text. (Adapted from Fig. 5 of Ref. 20.)

Copyright © 2001 Marcel Dekker, Inc.


Microstructure Effects on Transport 245

the disperse phase (brine) volume fraction is proportional to 1  . This diagram illus-
trates a large single-phase isotropic microemulsion domain bounded by a lamellar region
(L ), an upper two-phase region (2) comprising a decane-rich phase and a decane-in-brine
microemulsion, and a lower two-phase region (2 ) comprising a brine-rich phase and a
brine-in-decane microemulsion. Chen et al. [20] found that the mean curvature changed
sign and went through zero as composition in the single-phase region changed with  from
brine-in-water droplets at high , to irregular bicontinuous microstructure at intermediate
, to oil-in-brine droplets at low  and higher temperatures (60–80 C). The measurements
described shortly were all obtained [37] at 45 C along the double arrow-ended isotherm
pictured in Fig. 1 in the isotropic single-phase microemulsion domain. Results for low-
frequency conductivity and self-diffusion over the composition range 1:0 ! 0:3 in  are
described along this isotherm.
Low-frequency conductivity data [37] obtained along this 45 C isotherm are illu-
strated in Fig 2. The initial oscillatory variation in the conductivity for  > 0:9 can be
assigned to variations in AOT partitioning among dimers and other low aggregates and
reverse micelles, as reverse micelles are nucleated by added water (brine). These variations
will be discussed in greater detail in another publication. The key behavior for the pur-
poses of this exposition is the onset of the electrical conductivity percolation at  ¼ 0:85.
The conductivity increases two orders as  decreases from 0.85 to 0.70, and as shown in the
inset, the conductivity increases another two orders as a  decreases from 0.7 to 0.3.
Self-diffusion data [37] derived from NMR PGSE measurements for decane, water,
and AOT are illustrated in Fig. 3. The self-diffusion of decane decreases gradually as 
decreases from 1.0 to 0.3. The magnitude of decane self-diffusion suggests that the micro-
structure remains substantially continuous in decane over this composition range. Both
water and AOT diffusion initially decrease as  decreases. One can readily see that in this

FIG. 2 Low-frequency conductivity at 45 C as a function of composition,  (weight fraction decane


relative to decane and brine) for brine, decane, and AOT microemulsions exhibiting the phase
behavior illustrated in Fig. 1. The breakpoint at  ¼ 0:85 corresponds to the onset of percolation.
This conductivity increases by two orders as  decreases from 0.85 to 0.7. (Reproduced by permis-
sion of the American Institute of Physics from Ref. 37.)

Copyright © 2001 Marcel Dekker, Inc.


246 Texter

FIG. 3 Self-diffusion coefficients of decane (~), water (&), and AOT (*) in brine, decane, and
AOT microemulsions at 45 C as a function of decane weight fraction,  (relative to decane and
brine). Breakpoints in the self-diffusion data for both water and AOT are observed at  ¼ 0:85 and
at 0.7. (Reproduced by permission of the American Institute of Physics from Ref. 37.)

system, at least, water is necessary for the existence of reverse micelles. In the absence of
added water (brine), the AOT diffusion approaches that of monomeric water, suggesting
that very, very small aggregates (e.g., dimers) of AOT form at very low water contents.
These reverse micelles grow as added brine (water) increases ( decreases), and the self-
diffusion of water and of AOT concomitantly slow.
Water proton self-diffusion exhibits a break point and begins to increase at  ¼ 0:85.
In the case of AOT self-diffusion, a breakpoint also occurs, but AOT self-diffusion con-
tinues to slow as  decreases further. These breakpoints in both water and AOT self-
diffusion behavior at  ¼ 0:85 coincide with the breakpoint in electrical conductivity illu-
strated in Fig. 1, where the onset of electrical conductivity percolation occurs. At  ¼ 0:7
two more breakpoints in the water proton and AOT self-diffusion are seen. Water proton
self-diffusion increases more markedly and AOT self-diffusion beings to increase mark-
edly.

A. Two-State Model
A simple two-state model for the observed water proton self-diffusion may be put forward
in the form
Dobs ¼ xDc þ ð1  xÞDmic ð1Þ
where Dobs is the observed water proton self-diffusion coefficient, Dc is the self-diffusion
coefficient of molecular water in decane, Dmic is the self-diffusion coefficient of the reverse
micelles, and x is the mole fraction of water dispersed molecularly in the decane. The
coefficient Dc was taken as 0:93  105 cm2 s1 from independent measurements [37] of
molecular water diffusion in decane. The coefficient Dmic was approximated by the
observed self-diffusion coefficient for AOT for   0:7 and by the dotted curve illustrated
in Fig. 3 for  < 0:7. With these assumptions and measurements, the mole fraction x of

Copyright © 2001 Marcel Dekker, Inc.


Microstructure Effects on Transport 247

FIG. 4 Apparent mole fraction (x) water in continuous phase of brine, decane, and AOT micro-
emulsion system derived from the water self-diffusion data of Fig. 3 using the two-state model of
Eq. (1).

water in the continuous phase may be derived from these self-diffusion data. This deriva-
tion is illustrated in Fig. 4, where it appears that the amount of water in the continuous
decane pseudophase increases steadily as  decreases below 0.85.

B. Order Parameter
The apparent increase in water solubilized in the continuous decane pseudophase that is
illustrated in Fig. 4 goes counter to chemical intuition. The amount of water in the decane
pseudophase should partition between micelles and the decane relatively quickly as brine is
added to the microemulsion. This in fact appears to occur by the time  has decreased to
0.95, and this partitioning may be expected to remain constant over some interval of
further brine increase. The mole fraction of water in decane should hold constant and
then decrease as total water (brine) in the system increases. The horizontal asymptote
illustrated in Fig. 4 and extrapolated to the right in the direction of decreasing  represents
this partitioning approximation. The mole fraction of water in decane should not go above
this level, no matter how much total brine is added to the system. We define x^ as the
anomalous mole fraction of water in the continuous phase and as the quantity represented
by the vertical arrow in Fig. 4. We further assign this quantity as the amount of water in
percolating clusters and irregular bicontinuous microstructure. With this definition, we
can define the following order parameter S for the amount of water in such clusters and
microstructures,
x^
S¼ ð2Þ
1  x þ x^
where 1  x is simply the amount of water in nonpercolating clusters and isolated reverse
microemulsion droplet and where 0 S 1. This order parameter describes quantita-

Copyright © 2001 Marcel Dekker, Inc.


248 Texter

FIG. 5 Order parameter for disperse pseudophase water (percolating clusters versus isolated swol-
len micelles and nonpercolating clusters) derived from self-diffusion data for brine, decane, and AOT
microemulsion system of single-phase region illustrated in Fig. 1. The a and arrow denote the onset
of percolation in low-frequency conductivity and a breakpoint in water self-diffusion increase. The
other arrow (b) indicates where AOT self-diffusion begins to increase.

tively how water is distributed among disperse phase microstructures: (1) isolated reverse
micelles and nonpercolating micellar clusters; (2) percolating micellar clusters and irregu-
lar bicontinuous microstructures.
The order parameter values calculated from the data of Fig. 4 are illustrated in Fig.
5. The data there suggest the existence of two continuous transitions, one at  ¼ 0:85 and
another at  ¼ 0:7. The first transition at  ¼ 0:85, denoted by the arrow labeled ‘‘a’’ in
Fig. 5, is assigned to the formation of percolating clusters and aggregates of reverse
micelles. The onset of electrical percolation and the onset of water proton self-diffusion
increase at this same value of  (0.85) as illustrated in Figs. 2 and 3, respectively, are
qualitative markers for this transition. This order parameter allows one to quantify how
much water is in these percolating clusters. As  decreases from 0.85 to 0.7, this quantity
increases to about 2–3% of the water.
Another more abrupt transition in this order parameter occurs at  ¼ 0:7 under the
arrow labeled ‘‘b.’’ This transition is assigned to the onset of irregular bicontinuous
microstructure formation, and is indicated qualitatively by the marker illustrated in Fig.
3, where the onset in AOT self-diffusion increase occurs.

III. TEMPERATURE-INDUCED PERCOLATION

A somewhat different water, decane, and AOT microemulsion system has been studied by
Feldman and coworkers [25] where temperature was used as the field variable in driving
microstructural transitions. This system had a composition (volume percent) of 21.30%
water, 61.15% decane, and 17.55% AOT. Counterions (sodium ions) were assigned as the
dominant charge transport carriers below and above the percolation threshold in electrical

Copyright © 2001 Marcel Dekker, Inc.


Microstructure Effects on Transport 249

conductivity. A charge-hopping mechanism above the percolation threshold was also


invoked. This mechanism was outlined as emanating from the hopping of AOT molecules
between droplets in percolating clusters.
Low-frequency conductivity data and NMR PGSE self-diffusion results for water
protons and for AOT were obtained [25] over 6–40 C. These data are illustrated in Fig. 6.
The water self-diffusion data exhibit a breakpoint at about 18–19 C in close proximity to
the onset of percolation in electrical conductivity. This ‘‘onset’’ is highlighted in Fig. 6 by
the arrow labeled p at about 18 C. A separate breakpoint in the AOT self-diffusion data
appears at about 28 C. This coincidence in the onset of percolation in electrical conduc-
tivity with a breakpoint in water self-diffusion increase appears similar to the qualitative
markers seen for the onset of formation of percolating clusters in the discussion of the
previous section, where disperse phase volume fraction was the field variable driving
percolation. Additionally, the separate breakpoint for increasing AOT self-diffusion at
higher field variable (temperature 28 C) is similar to the breakpoint seen in Fig. 3 at
 ¼ 0:7. Again, this surfactant self-diffusion increase marker can be assigned to the
onset of irregular bicontinuous microstructure formation. These qualitative similarities
to the system analyzed in the previous section suggest a similar order parameter analysis
may be warranted.
Independent self-diffusion measurements [38] of molecularly dispersed water in
decane over the 8–50 C interval were used, in conjunction with the self-diffusion data
of Fig. 6, to calculate the apparent mole fraction of water in the pseudocontinuous phase
from the two-state model of Eq. (1). In these calculations, the micellar diffusion coeffi-
cient, Dmic , was approximated by the measured self-diffusion coefficient for AOT below
28 C, and by the linear extrapolation of these AOT data above 28 C. This apparent mole
fraction x was then used to graphically derive the anomalous mole fraction x^ of water in
the pseudocontinuous phase. These mole fractions were then used to calculate values for

FIG. 6 Self-diffusion and conductivity data reported by Feldman et al. [25] for reverse water,
decane, and AOT microemulsion as a function of temperature. The p and arrow between 18 and
19 C shows the approximate onset of percolation in low-frequency conductivity and a breakpoint in
water self-diffusion increase. Another breakpoint, at about 28 C, occurs in the AOT self-diffusion
data where AOT self-diffusion begins to markedly increase.

Copyright © 2001 Marcel Dekker, Inc.


250 Texter

FIG. 7 Order parameter for disperse pseudophase water derived from self-diffusion data for water,
decane, and AOT reverse microemulsion illustrated in Fig. 6. The p and arrow denote the approx-
imate onset of percolation in low-frequency conductivity and a breakpoint in water self-diffusion
increase. The arrow labeled AOT shows a second continuous transition corresponding to the onset
of AOT self-diffusion increase.

the order parameter of Eq. (2). This order parameter is illustrated in Fig. 7, where break-
points suggest the occurrence of two continuous transitions. The first of these at about
18 C and marked with an arrow labeled p occurs at the temperature at which the onset in
electrical conductivity percolation begins and at the temperature at which the water self-
diffusion data exhibit a breakpoint. This first continuous transition corresponds to the
onset of percolating cluster formation. The second breakpoint in the order parameter
occurs at about 28–29 C where the AOT self-diffusion coefficients begin to increase.
This second transition is assigned to the onset of irregular bicontinuous microstructure
formation.
These qualitative microstructure transition assignments may be extended to other
reverse microemulsion systems reported in the literature. Geiger and Eicke [39] applied
both electrical conductivity and NMR self-diffusion measurements to the study of a water,
isooctane, and AOT microemulsion driven into percolation by increasing temperature.
Inspection of the data published in their figures indicates a coincidence in the temperature
at which electrical conductivity percolation commences and at which water proton self-
diffusion markedly begins to increase. In addition, the self-diffusion of AOT begins to
increase markedly at a higher temperature, indicating the existence of two distinguishable
microstructural transitions, such as those discussed here. Similar qualitative markers may
be seen by inspection of the data published by Jonströmer et al. [40], where reverse
microemulsions of water (HDO), N-methylformamide (NMF), isooctane, and AOT
were examined. Their NMR self-diffusion data showed that the HDO and NMF self-
diffusion exhibited a breakpoint in the neighborhood of 19–20 C where they markedly
began to increase. These breakpoints correspond to the onset of electrical conductivity
percolation. At higher temperature, at about 26 C, the self-diffusion of AOT begins to
increase significantly. Again, these qualitative markers strongly suggest that the processes
of percolating cluster formation and irregular bicontinuous microstructure formation are
experimentally distinguishable, at least for the microemulsion system reviewed herein.

Copyright © 2001 Marcel Dekker, Inc.


Microstructure Effects on Transport 251

IV. CHEMICAL-POTENTIAL-DRIVEN PERCOLATION

As described in the introduction, certain cosurfactants appear able to drive percolation


transitions. Variations in the cosurfactant chemical potential, RT ln (where is cosur-
factant concentration or activity), holding other compositional features constant, provide
the driving force for these percolation transitions. A water, toluene, and AOT microemul-
sion system using acrylamide as cosurfactant exhibited percolation type behavior for
a variety of redox electron-transfer processes. The corresponding low-frequency electrical
conductivity data for such a system is illustrated in Fig. 8, where the water, toluene,
and AOT mole ratio (11.2 : 19.2 : 1.00) is held approximately constant, and the acrylamide
concentration, , is varied from 0 to 6% (w/w). At about ¼ 1:2%, the arrow labeled p in
Fig. 8 indicates the onset of percolation in electrical conductivity.
NMR PGSE self-diffusion coefficients obtained [36] for toluene, water, acrylamide,
and AOT in this microemulsion system are illustrated in Fig. 9 as a function of , the
acrylamide concentration. The self-diffusion coefficients for water, acrylamide, and AOT
exhibit breakpoints at about ¼ 1:2%, as marked by the arrow labeled p . Again, the
onset of percolation in electrical conductivity coincides with the onset of water proton self-
diffusion increase, and these markers indicate that percolating cluster formation (alterna-
tively, cluster percolation) commences at about ¼ 1:2%. While water and acrylamide
self-diffusion coefficients coincide for 1:2%, they diverge for  1:2%. The AOT and
acrylamide self-diffusion coefficients run parallel to one another, but the AOT values are
much lower. The higher values observed for acrylamide may be assigned to the partition-
ing of acrylamide into the continuous toluene pseudophase, resulting from the nontrivial
solubility of acrylamide in toluene. A major qualitative difference observed in this system,
compared to the two systems described, respectively, in the previous two sections, is that
the self-diffusion of AOT does not exhibit a significant increase at any observed field
variable. There is no qualitative marker, therefore, for the formation of irregular bicon-
tinuous microstructure in this system.
A quantitative analysis of these self-diffusion data according to the two-state model
of Eq. (1) to generate the order parameter of Eq. (2) is straightforward. Dc was found to be

FIG. 8 Low-frequency conductivity () of water, toluene, and AOT reverse microemulsions at 25 C
as a function of acrylamide (cosurfactant) concentration, (wt%). The p and arrow at ¼ 1:2%
shows the approximate onset of percolation in low-frequency conductivity.

Copyright © 2001 Marcel Dekker, Inc.


252 Texter

FIG. 9 Measured self-diffusion coefficients at 25 C for toluene (~), water (*), acrylamide (&),
and AOT (^) in water, toluene, and AOT reverse microemulsions as a function of cosurfactant
(acrylamide) concentration, (wt%). The breakpoint at about 1.2% acrylamide approximately
denotes, the onset of percolation in electrical conductivity.

5:41  105 cm2 s1 by measuring molecularly dispersed water in toluene and by correcting
for local viscosity differences between toluene and these microemulsions [36]. Values for
Dmic were taken as the observed self-diffusion coefficient for AOT. The apparent mole
fraction of water in the continuous toluene pseudophases was then calculated from Eq. (1)
and the observed water proton self-diffusion data of Fig. 9. These apparent mole fractions
are illustrated in Fig. 10 (top) as a function of .
The apparent mole fraction in Fig. 10 is approximately constant at x ¼ 0:013 for
1:2%. This constancy corresponds to equilibrium partitioning of water between
toluene and the reverse micelles. An extrapolation of this partitioning to higher is
illustrated in Fig. 10 (top) by the dashed line. In analogy to the process illustrated in
Fig. 4, the anomalous mole fraction of water in the pseudocontinuous phase, x^ , is defined
as the difference between this dashed line and x. The corresponding order parameter is
then calculated according to Eq. (2) and is illustrated in Fig. 10 (bottom). A single break-
point is depicted by the arrow labeled p at about ¼ 1:2%. This breakpoint correspond-
ing to the single continuous transition is assigned to the onset of percolating cluster
formation. Since no additional continuous transition is revealed in this order parameter
with increasing , it can be concluded that there is no indication of irregular bicontinuous
microstructure formation in this system, at least over the composition range investigated.

V. EQUILIBRIUM MODEL

Several unifying conclusions may be based upon the order parameter results illustrated
here for microstructural transitions driven by three different field variables, (1) disperse
phase volume fraction, (2) temperature, and (3) chemical potential. It appears that the
onset of percolating cluster formation may be experimentally and quantitatively distin-
guished from the onset of irregular bicontinuous structure formation. It also appears that

Copyright © 2001 Marcel Dekker, Inc.


Microstructure Effects on Transport 253

FIG. 10 (Top) Mole fraction (x) of water in the continuous (toluene) phase as a function of
acrylamide content, (wt%), for water, toluene, and AOT reverse microemulsions. (Bottom)
Order parameter for disperse pseudophase water derived from mole faction data above. The p
and arrow at about 1.2% (w/w) acrylamide indicate the approximate onset of percolation in low-
frequency conductivity

the formation of percolating clusters does not necessarily lead to formation of irregular
bicontinuous microstructure, as exemplified in the system driven by acrylamide (cosurfac-
tant) chemical potential.
The order parameter approach reviewed here provides explicit support for the clus-
ter-bicontinuous cartoon presented in 1990 by Chen et al. [20] in their neutron scattering
study of curvature evolution. Furthermore, this order parameter approach provides a
quantitative measure of the amount of water present in each disperse pseudophase. Such
a quantitative measure should provide a basis for testing phenomenological and statistical
mechanical theories that advance to the point of being able to discern such microstructural
transitions. The quantitative estimates for the distribution of water among the different
microstructures provide a basis for new thinking about these isotropic microemulsion
domains. It appears that irregular bicontinuous microstructure evolves gradually from
reverse microemulsion droplets and/or from percolating clusters of such droplets. The
formation of such irregular bicontinuous microstructure has heretofore only been con-
sidered in the context of first-order transitions between multiple phase domains [41].
While the order parameters derived from the self-diffusion data provide quantitative
estimates of the distribution of water among the competing chemical equilibria for the
various pseudophase microstructures, the onset of electrical percolation, the onset of water
self-diffusion increase, and the onset of surfactant self-diffusion increase provide experi-
mental markers of the continuous transitions discussed here. The formation of irregular
bicontinuous microstructures of low mean curvature occurs after the onset of conductivity
increase and coincides with the onset of increase in surfactant self-diffusion. This onset of
surfactant diffusion increase is not observed in the acrylamide-driven percolation. This
combination of conductivity and self-diffusion yields the possibility of mapping pseudo-
phase transitions within isotropic microemulsions domains.

Copyright © 2001 Marcel Dekker, Inc.


254 Texter

ACKNOWLEDGMENTS

The collaboration of Edwin Garcı́a of Eastman Kodak Company in the early stages of this
work is gratefully acknowledged. The skillful collaboration of Brian Antalek of Eastman
Kodak and of Tony Williams while at Kodak has been essential and invaluable in gen-
erating the extensive NMR PGSE data upon which the order parameter approach
reviewed here was based. Many fruitful ideas and results were also generated in collabora-
tion with Nissim Garti and Yuri Feldman, and their students, at the Hebrew University of
Jerusalem.

REFERENCES

1. R. A. Mackay and J. Texter (eds.), Electrochemistry in Colloids and Dispersions, VCH


Publishers, New York, 1992.
2. M.-P. Pileni (ed.), Structure and Reactivity in Reverse Micelles, Elsevier, Amsterdam, 1989.
3. P. Douzou, E. Keh, and C. Balny. Proc. Natl. Acad. Sci. U.S.A. 76:681 (1979).
4. N. Miyoshi and G. Z. Tomita. Z. Naturforsch. 35b:107 (1980).
5. S.-H. Chen and R. Rajagopalan (eds.), Micellar Solutions and Microemulsions, Springer-Verlag,
New York, 1990.
6. R. Zana and J. Lang, in Microemulsions: Structure and Dynamics (S. E. Friberg and P.
Bothorel, eds.), CRC Press, Boca Raton, pp. 153–172.
7. H.-F. Eicke, J. C. W. Shepherd, and A. Steinemann. J. Colloid Interface Sci. 56:168 (1976).
8. J. Peyrelasse and C. Boned. Phys. Rev. A 41:938 (1990).
9. J. Texter, B. Antalek, E. Garcı́a, and A. J. Williams, in Amphiphiles at Interfaces (J. Texter,
ed.), Steinkopff Verlag, Darmstadt, 1997, pp. 160–169; Prog. Colloid Polym. Sci. 106:160
(1997).
10. Y. Feldman, N. Kozlovich, I. Nir, and N. Garti. Phys. Rev. E 51:478 (1995).
11. B. Lindman and U. Olsson. Ber. Bunsenges. Phys. Chem. 100:344 (1996).
12. M. T. Clarkson, D. Beaglehole, and P. T. Callaghan. Phys. Rev. Lett. 54:1722 (1985).
13. L. E. Scriven. Nature 263:123 (1976).
14. B. Lagourette, J. Peyrelasse, C. Boned, and M. Clausse. Nature 281:60 (1979).
15. M. Lagües. J. Phys. (Paris) Lett. 40:L331 (1979).
16. R. Jóhannsson, and M. Almgren. Langmuir 9:2879 (1993).
17. M. Lagües, R. Ober, and C. Taupin. J. Phys. 39:L487 (1978).
18. A.-M. Cazabat, D. Chatenay, D. Langevin, and J. Meunnier. Faraday Discuss. Chem. Soc.
76:291 (1982).
19. D. Chatenay, W. Urbach, A.-M. Cazabat, and D. Langevin. Phys. Rev. Lett. 54:2253 (1985).
20. S.-H. Chen, S.-L. Chang, and R. Strey. J. Chem. Phys. 93:1907 (1990).
21. A. Jada, J. Lang, and R. Zana. J. Phys. Chem. 93:10 (1989).
22. M. W. Kim and J. S. Huang. Phys. Rev. A 34:719 (1986).
23. J. Peyrelasse and C. Boned, in Structure and Dynamics of Strongly Interacting Colloids and
Supramolecular Aggregates in Solution (S.-H. Chen, J. S. Huang, and P. Tartaglia, eds.),
Kluwer Academic Publishers, Dordrecht, 1992, pp. 801–806.
24. P. Alexandridis, J. F. Holzwarth, and T. A. Hatton. J. Phys. Chem. 99:8222 (1995).
25. Y. Feldman, N. Kozlovich, 1. Nir, N. Garti, V. Archipov, Z. Idiyatullin, Y. Zuev, and V.
Fedotov. J. Phys. Chem. 100:3745 (1996).
26. J. Lang, N. Lalem, and R. Zana. J. Phys. Chem. 95:9533 (1991).
27. L. M. M. Nazário, T. A. Hatton, and J. P. S. G. Crespo. Langmuir 12:6326 (1996).
28. E. Garcı́a and J. Texter. Proc. Electrochem. Soc. 93(1):2166 (1993).
29. E. Garcı́a, S. Song, L. E. Oppenheimer, B. Antalek, A. J. Williams, and J. Texter. Langmuir
9:2782 (1993).

Copyright © 2001 Marcel Dekker, Inc.


Microstructure Effects on Transport 255

30. E. Garcı́a and J. Texter. J. Colloid Interface Sci. 162:262 (1994).


31. B. Antalek, A. Williams, E. Garcı́a, D. H. Wall, S. Song, and J. Texter, in Dynamic Properties
of Interfaces and Association Structures (V. Pillai and D. O. Shah, eds.), Champaign, IL: AOCS
Press, 1996, pp. 183–196.
32. M. T. Carver, E. Hirsch, J. C. Whittmann, R. M. Fitch, and F. Candau. J. Phys. Chem.
93:4867 (1989).
33. B. Antalek, A. J. Williams, E. Garcı́a, and J. Texter. Langmuir 10:4459 (1994).
34. P. Stilbs, B. Lindman. Progr. Colloid Polymer Sci. 69:39 (1984).
35. P. Guéring and B. Lindman. Langmuir 1:464 (1985).
36. B. Antalek, A. J. Williams, and J. Texter. Phys. Rev. E 52:R5913 (1996).
37. J. Texter, B. Antalek, and A. J. Williams. J. Chem. Phys. 106:7869 (1997).
38. B. Antalek, A. J. Williams, J. Texter, Y. Feldman, and N. Garti. Colloids Surf. A. Physiochem.
Eng. Aspects 128:1 (1997).
39. S. Geiger and H.-F. Eicke. J. Colloid Interface Sci. 110:181 (1986).
40. M. Jonströmer, U. Olsson, and W. O. Parker, Jr. J. Phys. Chem. 11:61 (1995).
41. Y. Talmon and S. Prager. J. Chem. Phys. 69:2984 (1978).

Copyright © 2001 Marcel Dekker, Inc.

You might also like