Vous êtes sur la page 1sur 18

Journal of Colloid and Interface Science 248, 203–220 (2002)

doi:10.1006/jcis.2001.8104, available online at http://www.idealibrary.com on

FEATURE ARTICLE
Dimeric (Gemini) Surfactants: Effect of the Spacer Group
on the Association Behavior in Aqueous Solution
Raoul Zana1
Institut Charles Sadron (CNRS-ULP), 6 rue Boussingault, 67000 Strasbourg, France

Received October 1, 2001; accepted November 8, 2001; published online February 15, 2002
the corresponding monomeric surfactant DTAB (dodecyltrime-
Dimeric (gemini) surfactants are made up of two amphiphilic thylammonium bromide) is 0.50 wt%.
moieties connected at the level of, or very close to, the head groups by 2. Dimeric surfactants are much more efficient than the cor-
a spacer group of varying nature: hydrophilic or hydrophobic, rigid responding monomeric surfactants at decreasing the surface ten-
or flexible. These surfactants represent a new class of surfactants sion of water. For instance, the concentration required for low-
that is finding its way into surfactant-based formulations. The na- ering the surface tension of water by 0.02 N/m is 0.21 wt% for
ture of the spacer group (length, flexibility, chemical structure) has
DTAB and only 0.0083 wt% for 12-2-12 (1, 2).
been shown to be of the utmost importance in determining the solu-
3. Aqueous solutions of dimeric surfactants with short spac-
tion properties of aqueous dimeric surfactants. This paper reviews
the effect of the nature of the spacer on some of these properties. The ers can have very high viscosities at relatively low concentrations
behavior of dimeric surfactants in the submicellar range of concen- whereas solutions of the corresponding monomeric surfactants
tration, at interfaces, in dilute solution (solubility in water, Krafft have low viscosities. For instance the viscosity of aqueous solu-
temperature, critical micellization concentration, thermodynamics tions of DTAB is only marginally larger than that of pure water
of micelle formation, micelle ionization degree, size, polydispersity, up to a surfactant concentration of at least 10 wt% whereas a
micropolarity and microviscosity, microstructure and rheology of 5 wt% solution of 12-2-12 has a viscosity of several hundred
the solutions, solubilization, micelle dynamics, and interaction with Pa/s and is viscoelastic (1, 2). Solutions of 12-2-12 can also
polymers) and in concentrated solution (phase behavior) are succes- display shear thickening at fairly low concentrations (3). These
sively reviewed. Selected results concerning trimeric and tetrameric properties reflect the ability of dimeric surfactants with a short
surfactants are also reviewed. C 2002 Elsevier Science (USA)
spacer, such as 12-2-12, to give rise to worm-like micelles at
Key Words: dimeric surfactants; oligomeric surfactants; gemini
fairly low surfactant concentrations, even in the absence of added
surfactants.
salt (4).
The fact that the properties of dimeric surfactants can differ
I. INTRODUCTION greatly from those of conventional surfactants has been related
to the distribution of distances between head groups in micelles
Dimeric surfactants represent a new class of surfactants. They formed by these two types of surfactants (5). For conventional
are made up of two amphiphilic moieties connected at the level surfactants, this distribution goes through a maximum at a ther-
of the head groups or very close to the head groups by a spacer modynamic equilibrium distance dT ≈ 0.7–0.9 nm. For dimeric
group, as schematically represented in Fig. 1A (1, 2). The current surfactants the distribution is bimodal, with a first maximum at
interest in such surfactants arises from three essential properties. the thermodynamic distance dT and another more narrow maxi-
1. Dimeric surfactants are characterized by cmc (critical mum at a distance ds that corresponds to the length of the spacer.
micellization concentrations) that are one to two orders of This length is determined by the bond lengths and bond angles
magnitude lower than those for the corresponding conventional between the atoms making up the spacer group. The bimodal
(monomeric) surfactants (1, 2). For instance, the cmc of the distribution of head group distances and the effect of the chemi-
dimeric surfactant 12-2-12 (dimethylene-1,2-bis(dodecyldime- cal link between head groups on the packing of surfactant alkyl
thylammonium bromide)) is about 0.055 wt%, whereas that of chains in the micelle core are expected to strongly affect the
curvature of surfactant layers and thus the micelle shape and
the properties of the solution. The distance ds can be adjusted
1 E-mail: Zana@ics.u-strasbg.fr. to be smaller than, equal to, or larger than dT by modifying the
203 0021-9797/02 $35.00

C 2002 Elsevier Science (USA)
All rights reserved.
204 RAOUL ZANA

that are subsequently referred to as oligomeric surfactants (8).


The schematic representation of a trimeric surfactant is given in
Fig. 1B. Results for oligomeric surfactants are reviewed below
whenever appropriate. At the outset it is pointed out that most
of the gain in properties is achieved by going from a monomeric
to a dimeric surfactant.
Several reviews on dimeric surfactants have been recently
A B published (1, 2, 6, 9–11). The present review specifically fo-
cuses on the effect of the spacer group. It does not attempt to be
FIG. 1. Schematic representation of a dimeric surfactant (A) and a trimeric thorough. Only the most significant results that refer to the effect
surfactant (B). of the nature and length of the spacer group on solution prop-
erties of dimeric surfactants are presented. Indeed, and this will
structure of the spacer. Those different situations are expected become clearer below, the nature and length of the spacer group
to give rise to a rich variety of behaviors. are the most important parameters in determining the proper-
In principle, it is possible to synthesize dimeric surfactants ties of dimeric surfactants. This paper successively reviews the
by using two of any type of the presently known amphiphiles, properties of dimeric surfactants at interfaces, in solutions at
identical or different, and connecting them with a spacer group concentrations below the cmc, and in solutions at concentra-
of varied chemical nature, hydrophilic or hydrophobic, rigid or tions above the cmc, in the dilute and concentrated ranges.
flexible. This possibility has led to the synthesis of dimeric sur- Most of the reviewed results concern cationic dimeric surfac-
factants with an enormous variety of structures (1, 2) and also tants of the bisquaternary ammonium type. Table 1 shows some
of properties that are difficult to obtain with conventional sur- of these surfactants (12–22). The abbreviations used to refer to
factants in the absence of additives. The potential applications the surfactants in Table 1 are retained throughout this review.
suggested by these properties led to a renewed interest of the The surfactants m-s-m (A1) and m-EOz -m (A3) are the ones
industrial community in dimeric surfactants. Many surfactant- that are referred to the most.
producing companies have ongoing research on dimeric surfac-
tants. The company Sasol (formerly Condea, Marl, Germany) II. AQUEOUS DIMERIC SURFACTANT SOLUTIONS AT
is offering formulations based on anionic dimeric surfactants CONCENTRATIONS BELOW THE CRITICAL
(Ceralution). MICELLIZATION CONCENTRATION
Dimeric surfactants have been also reported to have better
solubilizing, wetting, foaming, and lime-soap-dispersing prop- Many papers discussed the state of ionic dimeric surfactants
erties than conventional surfactants (6). These properties are at concentrations below the cmc. Several situations were consi
commonly used to evaluate surfactant performances. Besides, dered. First, the dimer may be completely dissociated, giving
the Krafft temperatures of dimeric surfactants with hydrophilic rise to one dimeric ion and two counterions, or not completely
spacers (see below) are generally very low (6), giving these sur- dissociated, with partial binding of one counterion to the dimeric
factants the capacity to be used in cold water. Finally, cationic ion. Second, depending on the conformation of the spacer group
dimeric surfactants have been shown to possess a strong bi- of the dimeric surfactant, its two alkyl chains may or may not in-
ological activity (7). The interesting properties displayed by teract. Third, premicellar aggregation may take place, giving rise
dimeric surfactants has led to the synthesis of longer homologs to small aggregates of dimeric ions of low aggregation number.
TABLE 1
Examples of Bisquaternary Ammonium Bromide Dimeric Surfactants

A1: R1 = R2 = Cm H2m+1; Y = CH2 ; x + y + 1 = s; m-s-m surfactants (12).


A2: R1 = R2 = Cm H2m+1 ; Y = CH2 , O, S, N(CH3 ), x = y = 2 (13, 14).
A2 : R1 = R2 = Cm H2m+1 ; Y = CHOH, (CHOH)2 ; x = y = 1 (14).
A3: R1 = R2 = Cm H2m+1 ; Y = (OCH2 CH2 )z , x = 2; y = 0; m-EOz -m surfactants (15).
A4: R1 = R2 = Cm H2m+1 ; Y = C ≡ C; x = y = 1 (16).
A5a : R1 = R2 = Cm H2m+1 ; Y = ; x = y = 1 (17).
A6: R1 = R2 = Cm H2m+1 OC(O)CH2 ; no Y ; x = y = 1; counterion = chloride (18, 19).
A7: R1 = R2 = Cm F2m C4 H8 ; no Y ; x = y = 1 (20).
A8: R1 = Cm H2m+1 ; R2 = Cm  H2m  +1 ; no Y ; x = y = 1; s; m-2-m  surfactants (21, 22).

a  represents a phenylene group.


DIMERIC (GEMINI) SURFACTANTS 205

The possibility of a partial binding of a counterion by a dimeric Thermodynamic studies suggested that the two alkyl chains
surfactant ion at concentrations C < cmc arose in the analysis in the surfactant A6 might be partly associated in the molecu-
of the plots of the surface tension γ vs C for dimeric surfactant larly dispersed state (19). This suggestion was not supported by
solutions. These plots permit the determination of the surface the values of the free energy change associated with the transfer
excess  and a = 1/NA (NA = Avogadro’s number), surface of one alkyl (dodecyl) chain from the aqueous phase to the mi-
area occupied by one surfactant at the air/water interface, on the celle, G ◦ (C12 ), for two series of oligomeric surfactants (34).
basis of the Gibbs equation: These values were calculated with the available cmc and ion-
ization data, using the appropriate equation (35). Self-coiling
of the alkyl chains was mentioned as a possible explanation for
 = −(dγ /d ln C)/n RT. [1]
the aging effect observed in measurements of surface tension of
solutions of dimeric surfactants with long alkyl chains, below
In Eq. [1] R is the gas constant and T the absolute temperature. the cmc (17). However such effects are not observed with con-
The constant n takes the value 2 for an ionic surfactant where ventional surfactants having the same long alkyl chains. A 13 C
the surfactant ion and the counterion are univalent and the value and 1 H NMR investigation of the conformation of dimeric sur-
3 for a dimeric surfactant made up of a divalent surfactant ion factant A5 with m = 8 below the cmc did not reveal interactions
and two univalent counterions, in the absence of a swamping between octyl chains (36).
electrolyte. The reported a values were calculated on the basis Premicellar aggregation occurs in solutions of conventional
of n = 2 (13, 23–25), on the assumption that one of the two surfactants that are sufficiently hydrophobic. The situation is
charged groups is neutralized by a bound counterion, or n = 3, somewhat similar with dimeric surfactants. Menger and Littau
on the assumption of a full dissociation of the dimeric surfac- (17) provided the first evidence for premicellar aggregation in
tants (16, 17, 26–29). Some studies reported two sets of values solutions of several series of dimeric surfactants, in particular
of a based on n = 2 and 3 (15, 30). The same problem arises for surfactant A5. The variation of the log cmc with the alkyl
for ionic trimeric surfactants for which the value n = 4 should chain carbon number m showed a strong upward curvature or
be used if the surfactant is fully ionized at C < cmc. The know- a minimum for m ≥ 16, instead of the usual linear variation.
ledge of the surface area occupied by a surfactant at an interface Premicellar aggregation was later reported for other series of
is very important in surfactant science. The problem of the value dimeric surfactants (14, 16, 37, 38). Electrical conductivity mea-
of n in the case of oligomeric surfactants therefore needed to be surements are particularly well suited for obtaining evidence of
solved. An attempt in this direction was performed by means of premicellar aggregation (33, 39). This is illustrated in Fig. 2A
neutron reflectivity, a technique that permits a direct determina- and 2B. The variations of the electrical conductivity K with
tion of the surface excess  (31). The comparison of the value the surfactant concentrations C and of the molar conductivity
of  from neutron reflectivity to that of (dγ /d ln C)/RT from with C 1/2 are represented for the surfactants 12-8-12 and 16-8-16
surface tension data made it possible to determine the value of n (33). The plots for the 12-8-12 surfactant show a normal behav-
at any surfactant concentrations. The comparison yielded n = 2 ior. For the 16-8-16 surfactant the K vs C plot shows a small
for the dimeric surfactants 12-2-12, 12-3-12, and 12-12-12 and upward curvature and the vs C 1/2 plot shows a pronounced
n = 3 for the surfactant A5 with m = 12 (31). For 12-6-12 the maximum in the submicellar concentrations range. The vs
value of n was found to decrease from 3 at low C to 2 at higher C C 1/2 plot for the dimers 14-8-14, 16-paraxylylene-16, and 18-
but still below the cmc. The authors concluded that one dimeric 8-18 also show a maximum, the amplitude of which increases
surfactant ion binds one bromide ion (ion pairing) in submicel- rapidly with m (33). This maximum is the signature of premi-
lar solutions of 12-s-12 surfactants. However it has been shown cellar association in surfactant solutions. It arises because the
that other effects may explain differences between neutron re- equivalent conductivity of a small aggregate of surfactant ions
flectivity and surface tension data (32). Besides, recent electrical (whether monomeric or dimeric) is larger than the sum of the
conductivity studies did not reveal any ion pairing in submicellar equivalent conductivities of the ions constituting it (33, 39). This
solutions of 12-s-12 surfactants (33). However, ion pairing was effect shows at different values of m for different dimeric sur-
evidenced in submicellar solutions of 8-s-8 and 10-s-10 surfac- factants (14, 16, 37, 38) but always at m ≥ 14–16. However, it
tants, owing to the larger cmc values of these surfactants with is already detected at m = 11 for more complex dimeric sur-
respect to 12-s-12 surfactants. Ion pairing is thus a real possibi- factants derived from arginine where the peptide segment sep-
lity for dimeric surfactants with high cmc values, i.e., those with arating the two charged groups may help in stabilizing small
an alkyl chain containing 8 or 10 carbon atoms but apparently premicellar aggregates (39). For the dimeric surfactants A1 the
not those with an alkyl chain containing 12 carbon atoms. Ad- spacer carbon number has a strong effect on the occurrence of
ditional studies are required to explain the different conclusions premicellization. Thus this effect does not occur for the 12-s-12
reached in neutron reflectivity and electrical conductivity stud- dimers up to s = 12 and for 16-3-16 and 16-4-16. However the
ies of 12-s-12 surfactants. Some consideration should be given vs C 1/2 plots for 12-14-12, 12-16-12, 12-20-12, 16-6-16, and
to a possible explanation of the neutron reflectivity results in 16-8-16 show a maximum (33). A study of two series of an-
terms of premicellar aggregation (see below). ionic trimeric surfactants suggests that premicellization occurs
206 RAOUL ZANA

at a value of m much lower than that for dimeric surfactants


A 0.5 0.5 (40, 41).
C (M )
0.00 0.01 0.02 0.03 0.04 0.05 III. DIMERIC SURFACTANTS AT INTERFACES
300 220
1. Air–Solution Interface
250 The studies of the adsorption of dimeric surfactant at the
200
air/solution interface aimed to assess the efficiency and effective-

Λ (S.cm /mol)
200 ness of these surfactants in reducing the surface tension of water.
µS/cm)

180 These measurements also aimed to measure the cmc and surface

2
150 area a occupied by one dimeric surfactant at the air/water inter-
Κ (µ

face on the basis of the Gibbs equation (Eq. [1]). The efficiency
160 and effectiveness are characterized by the value of the surfac-
100
tant concentration C20 at which the surface tension of water
140 is reduced by 0.02 N/m and by the value of the surface tension
50 at the cmc, γcmc , respectively.
Figure 3 shows the variation of a with the spacer carbon
0 120 number for 12-s-12 surfactants (26, 30) and with the number
0.0 0.5 1.0 1.5 2.0 n T = 3z + 2 of oxygen and carbon atoms in the spacer group
C (mM) in the case of 12-EOz -12 surfactants (15). For the 12-s-12 sur-
factants (hydrophobic spacer) a is a maximum at a value of s
around 10–12. A similar maximum in a at about the same value
of s appears to occur for the bolaform surfactants [N+ (CH3 )3 ,
Br− ]2 (CH2 )s (42). A maximum in a was observed for the dimeric
B 280
surfactants derived from arginine and which also contain a hy-
270
drophobic polymethylene spacer (43). However only a rather
Λ (S.cm /mol)

260 slow increase of a with n T is seen for the 12-EOz -12 surfac-
250 tants. A series of anionic surfactants having a poly(ethylene
2

240 oxide) spacer showed the same behavior as the 12-EOz -12 sur-
230 factants (44). Such a behavior is expected upon increasing the
volume and, thus, the surface of the poly(ethylene oxide) spacer,
220
which remains located on the water side of the interface.
210
200
0 2 4 6 8 10
3 0.5 0.5
10 C µM )

2
a (nm2)

16
µS/cm)

12
1
8
Κ (µ

0 0
0 20 40 60 80
4 8 12 16 20
µM)
C (µ
s, nT
FIG. 2. Variations of the electrical conductivity K with the surfactant con-
centration and of the molar conductivity with the square root of the surfactant FIG. 3. Surface area a occupied by one dimeric surfactant at the air–water
concentration for the dimeric surfactants 12-8-12 (A) and 16-8-16 (B). The interface. Variation with the spacer carbon number s at 25◦ C for 12-s-12 sur-
arrow in B (top) indicates the cmc as obtained from the K vs C plot in B factants (, data from Refs. 26 and 30) and with n T for 12-EOz -12 surfactants
(bottom). From Ref. (33). (, from Ref. 15).
DIMERIC (GEMINI) SURFACTANTS 207

For the 12-s-12 surfactants the maximum of a was explained 140


in terms of a change in location of the polymethylene spacer
120
upon increasing s. At s < 10, the spacer group is rather rigid

µM/g)
and lies flat with a fairly linear conformation at the air–solution 100
interface. This is suggested by the rapid initial increase of a 80

Γmax (µ
with s in Fig. 3 and is supported by X-ray scattering studies of
60
the lamellar and hexagonal phases in water/12-s-12 surfactant
mixtures (45). At s > 10 the spacer becomes too hydrophobic 40
to remain in contact with water and moves to the air side of the 20
interface, where it adopts a looped (wicket-like) conformation
0
(30, 42). This results in an overall decrease of a. This effect
0 5 10
may be enhanced by a change of orientation of the alkyl chains
with respect to the interface as s increases. The absence of a s
maximum for dimeric surfactants with a hydrophilic spacer (15,
FIG. 4. Spacer carbon number dependence of the maximum amount of
44) also supports this explanation. The maximum of a observed adsorbed 12-s-12 surfactant on raw () and HCl-treated () silica (from the
for the 12-s-12 surfactants was accounted for theoretically by results in Refs. 51 and 52).
statistical-mechanical calculations (46). The main factors that
determine the variation of a with s were shown to be the spacer
conformational entropy and the attractive and repulsive inter- 2. Solid-Solution Interface
actions between surfactant molecules. At s > 12 the calculated The adsorption isotherms of dimeric 12-s-12 (50–52),
value of a decreases upon increasing s more slowly than ex- trimeric 12-s-12-s-12 (34, 53), and tetrameric 12-3-12-4-12-
perimentally observed (46). This discrepancy may be due to 3-12 cationic surfactants on macroporous amorphous silica
premicellar aggregation. Indeed, this effect occurs at s > 12 for showed that the adsorption involves two steps, as for conven-
12-s-12 surfactants (33). It was not taken into account in the cal- tional surfactants (34). The first step occurs at very low concen-
culations (46) and was not considered in the experimental study tration and corresponds to a binding of individual dimeric sur-
(30). Other studies do show that premicellar aggregation can factants to charged sites on the silica surface by an ion-exchange
cause a significant decrease in the apparent value of a (14, 37). mechanism. The second step occurs at a concentrations slightly
Monte Carlo simulations were also performed to understand the below the cmc and corresponds to the formation of surface ag-
behavior of dimeric surfactants at the air–solution interface (47). gregates. Figure 4 shows that in the case of the adsorption of
The results in Fig. 3 are of the utmost importance for ex- 12-s-12 surfactants on both raw and HCl-treated silica the max-
plaining the properties of m-s-m surfactants in aqueous solution imum amount of adsorbed surfactant, max , decreases as s in-
that are described below as well as the differences in behavior creases (50–52). This variation is in relation with the structure of
between the two closely related series of dimeric surfactants the surface aggregates. An atomic force microscopy study (54)
m-s-m and m-EOz -m. Indeed the large changes of a with s for showed that 12-2-12 adsorbs as a flat bilayer, whereas 12-4-12
the 12-s-12 surfactants indicate correspondingly large changes and 12-6-12 adsorb as parallel cylinders (see Fig. 5). A surfac-
of the surfactant packing parameter that results in changes of tant with still lower packing parameters was shown to adsorb
micelle shape. On the contrary the relatively slow change of a under the form of spherical surface aggregates (54). Obviously
with n T = 3z + 2 seen for 12-EOz -12 surfactants suggests that the maximum amount of adsorbed surfactant decreases as the
the micelle structure should vary slowly with the number of structure goes from a flat bilayer to parallel cylinders and to
ethylene oxide groups in the spacer. spheres. Since mica and silica surfaces differ mostly by their
For cationic oligomeric surfactants the surface area per am- charge density, it is likely that differences in structure of the
phiphilic moiety has been shown to decrease in going from the surface aggregates similar to those seen on mica occur on silica,
monomer (DTAB) to the dimer (12-3-12) and the trimer (12-3- thereby explaining the change of max .
12-3-12). It then levels off when going to the tetramer (12-3-12- In the case of the 12-s-12/silica systems the stoichiometry of
4-12-3-12) (34). the ion exchange taking place in the first step of adsorption de-
The kinetics of adsorption of dimeric surfactants at the air– pends strongly on the spacer carbon number (52). Thus, surfac-
water interface strongly depends on the nature of the surfactant. tants with a short spacer (s = 2 or 3) replace only one sodium
Two reports on anionic (48) and cationic (49) dimeric sur- ion at the silica surface, revealing that the surfactant binds to
factants, the latter of type A2 and A5, indicated a diffusion- the surface mainly by only one head group. Surfactants with
controlled adsorption of the investigated surfactants. On the a long spacer (s > 8) replace two sodium ions, indicating that
contrary, a large barrier to adsorption was reported to exist for the two head groups bind to two surface sites. Surfactants with
cationic surfactant dimers derived from disulfur betaine (27). intermediate s values replace a number of sodium ions between
Dimeric surfactants with a flexible spacer lowered the surface 1 and 2. That result suggests the use of 12-s-12 surfactants as
tension of water faster than dimers with a rigid spacer. rulers for the study of the distribution of adsorption sites on solid
208 RAOUL ZANA

FIG. 5. AFM images (150 × 150 nm) of dimeric surfactants on the cleavage plane of mica, showing the 12-4-12 parallel cylinders (left) and the 12-2-12
bilayer (indicated by the absence of structural features, right). Reproduced from Ref. (54) with permission of the American Chemical Society.

surfaces (52). The heat released upon adsorption of 12-s-12 sur- below 0◦ C (2, 6, 9, 44). Such low values of TK permit the use of
factants (s = 2 − 12) on raw and HCl-treated silica has been dimeric surfactants in cold water.
measured calorimetrically. The results clearly show the two ad- The variation of TK with the spacer carbon number s has been
sorption steps. The results are still being analyzed in order to determined for solutions of the cationic dimeric surfactants 12-
extract the values of the enthalpy of adsorption (52). s-12 and 16-s-16 (56 and 57) and compared to the variation
The nature of the surface greatly influences the shape of of the melting temperature, TM , of the solid surfactants. The
dimeric surfactant aggregates adsorbed at the surface. Thus 12- results are shown in Fig. 6. There appears to be no correlation
2-12 and 12-4-12 surfactants adsorb on the cleavage plane of between the variations of TK and TM with s. For homologous
graphite under the form of half-cylindrical aggregates whereas series of conventional surfactants it is usually observed that the
they adsorb on mica as a planar bilayer and parallel cylindrical variations of TK and TM are correlated. For instance both TK
micelles, respectively (54). and TM increase with the surfactant chain length. The maximum
Atomic force microscopy (AFM) and measurements of sur- of TM at s = 5 for the two series of surfactants in Fig. 6 is
face force (55) evidenced the different stages of aggregation and noteworthy. The minimum of TM at s = 10–12 for the 12-s-12
organization of 12-2-12 onto mica, upon increasing the bulk surfactants was mentioned in Ref. (45). Figure 6A also shows the
concentrations of 12-2-12. Separate monolayer patches, a full variation of the melting temperature of 12-EOz -12 surfactants
monolayer, a full monolayer with some surfactant adsorbed on with n T = 3z + 2, total number of oxygen and carbon atoms in
top of it, and the formation of a full bilayer at a bulk surfac- the spacer (58). The plot apparently goes through a minimum at
tant concentrations of 0.18 mM, i.e., well below the cmc, were n T = 11, a value close to that of s for which TM is a minimum
successively observed. The formation of a full bilayer required for the 12-s-12 series.
hours (55). The few instances where the Krafft temperatures of a dimeric
The efficiency of adsorption on silica at a given surfactant surfactant and of the corresponding monomeric surfactant could
concentration below the cmc increases much with the degree of be directly compared revealed no systematic trend. Thus, the
oligomerization of the surfactant (34, 53). However the max- values of TK for 16-2-16 and CTAB (cetyltrimethylammonium
imum amount adsorbed, expressed in grams of surfactant per bromide) are, respectively, 45 and 25◦ C (55, 56). Those for 12-
gram of silica, increases only slightly in going from the monomer 2-12 and DTAB are 15◦ C and below O◦ C, respectively (55, 56).
to the dimer and then levels out (34). On the contrary, dimeric surfactants derived from arginine have
TK values lower than those of the corresponding monomeric
IV. SOLUBILITY IN WATER, CMC, AND surfactants (59).
THERMODYNAMICS OF MICELLIZATION OF DIMERIC
AND OLIGOMERIC SURFACTANTS 2. Critical Micellization Concentration
One of the main reasons for the current interest in dimeric
1. Solubility in Water, Krafft Temperature, and Melting
surfactants is that their cmc are much lower than those of the cor-
Temperature of Dimeric Surfactants
responding monomeric surfactants, by about one order of mag-
Ionic dimeric surfactants with m ≤ 12 are generally highly nitude and more. For instance, the cmc of 12-2-12 and DTAB
soluble in water particularly those with a hydrophilic spacer. are 0.055 and 0.50 wt%, respectively (12). The low cmc values
The reported Krafft temperatures, TK , of several series of anionic of dimeric surfactants with respect to the corresponding conven-
dimeric surfactants with hydrophobic or hydrophilic spacers are tional surfactants arise mainly because two alkyl chains, rather
DIMERIC (GEMINI) SURFACTANTS 209

250 (iii) The cmc of surfactants m-s-m with a hydrophobic poly-


A methylene spacer is a maximum at s = 5–6, irrespective of the
200 value of m (see Fig. 7) (12, 62, 63). This value of s is also that
for which the melting temperature of the m-s-m surfactants is a
150 maximum. A similar maximum of cmc was observed for dimeric
TM, TK (˚C)

surfactants derived from arginine that also includes a hydropho-


100
bic spacer (43). On the contrary, for both anionic and cationic
30 surfactants with a hydrophilic poly(ethylene oxide) spacer the
cmc increased progressively with n T = 3z + 2, total number of
20 oxygen and carbon atoms in the spacer, as seen in Fig. 7 for 12-
EOz -12 surfactants (15, 44). The increase of cmc with s observed
10 for s ≤ 6 for m-s-m surfactants is probably due to a conforma-
tional change in the surfactant molecule. The two alkyl chains
0 would be in a gauche or trans position at low s values and in a
0 5 10 15 20 cis position at higher s (see also next paragraph). Monte Carlo
s simulations of dimeric surfactant solutions attempted to account
for the behavior of the cmc of the surfactant (47). The calcula-
250 tions correctly predicted the presence of a maximum in the cmc
B vs s plot for surfactants with a hydrophobic spacer. However, the
200 calculated overall variations of the cmc with s were much larger
than experimentally observed. Also, the calculations predicted a
150 decrease of cmc with increasing n T for hydrophilic spacers (47),
TM, TK (˚C)

whereas the experimental results show the opposite behavior.


100 (iv) The cmc of quaternary ammonium oligomeric surfac-
tants with m = 12 decreased in a somewhat hyperbolic manner
50
in going from the monomer to the dimer, trimer, and tetramer
40
(34). The largest part of the decrease occurred in going from the
30 monomer to the dimer.
20
10
3. Thermodynamics of Micellization
0
0 4 8 12 Several recent papers have reported on the thermodynamics
of micellization of 12-s-12 surfactants (64–67). Figure 8 shows
s the variation of the enthalpy of micellization, HM◦ , of 12-s-12
FIG. 6. Spacer carbon number dependence of the Krafft temperature TK
() and of the melting temperature TM (, ). (A) 12-s-12 () and 12-EOz -12
() surfactants; (B) 16-s-16 surfactants (from Refs. 56 and 58). The lines are 10-2
guides to the eye. The data points for the Krafft temperature of surfactants 12-
6-12, 12-8-12, and 12-10-12 have been set at 0◦ C, while in fact the TK values
for these surfactants could not be measured (TK < 0◦ C).

10-3
cmc (M)

than one, are transferred at a time from water to the micelle


pseudo-phase (35, 60).
The results concerning the effects of various parameters on
the cmc of dimeric surfactants are briefly summarized. 10-4
(i) The cmc of surfactant A2 with the flexible hydrophobic
spacer (CH2 )2 Y (CH2 )2 depends little on the chemical nature of
Y . Thus, the cmc of surfactant A2 with m = 12 were found to
10-5
be (7, 13) 1.2, 1.1, 1.0 and 0.84 mM for Y ≡N(CH3 ), O, CH2 ,
0 4 8 12 16 20
and S, respectively.
(ii) The cmc of anionic or cationic dimeric surfactants with s, nT
a poly(ethylene oxide) spacer increase with the number z of
FIG. 7. Dependence of the cmc on the spacer carbon number s for the
ethylene oxide groups (15, 44, 61) whereas the cmc of conven- surfactants 10-s-10 (, from Ref. 62), 12-s-12 (∇, from Ref. 12), and 16-s-16
tional surfactants decreases upon intercalation of ethylene oxide (, from Ref. 12; , from Ref. 63) and on n T = 3z + 2 for the 12-EOz -12
groups between the alkyl chain and the charged group. surfactants (, from Ref. 15).
210 RAOUL ZANA

20 Eq. [3], valid for oligomeric surfactants (35):


28
19
G ◦M (C12 ) = RT (1 + 1/x − α) ln cmc. [3]
24 18

∆V˚M (cm /mol)


−∆H˚M (kJ/mol)

17 In Eq. [3], x is the number of dodecyl chains in the oligomeric


20
surfactant. G ◦M (C12 ) was found to be equal to −18.3, −20.8,
16
−21.5, and −22.8 kJ/mol, respectively (34). This variation is

3
16
15 rather small, within the experimental error on α. A similar result
12 14 was found for the three surfactants 12-3 (monomer), 12-6-12,
and 12-6-12-6-12 (34). Such results do not support the exis-
−∆

13 tence of an intramolecular association of the alkyl chains of the


8
12 oligomeric surfactant below the cmc postulated in Ref. (19).
4
11
V. PROPERTIES OF MICELLES OF
0 10 DIMERIC SURFACTANTS
0 4 8 12
1. Micelle Ionization Degree
s
Data concerning the ionization degree α of dimeric surfactant
FIG. 8. Variation of the enthalpy of micellization HM◦ (, from Refs. 52
micelles have been mostly obtained from the analysis of small-
and 66; , from Ref. 65) and of the volume change upon micellization VM◦
(, from Ref. 67) for the 12-s-12 surfactants at 25◦ C.
angle neutron scattering (SANS) and electrical conductivity data
(12, 15, 63, 67, 69–74). The electrical conductivity data have
been recently reinterpreted (66) and this resulted in values of α
surfactants with the spacer carbon number, as obtained from di- lower and more precise than those in other reports (12, 67).
rect calorimetric measurements (64–66). The two sets of data Both SANS and electrical conductivity yielded α values that
represented show important differences in numerical values that increased with the spacer carbon number for three series of
may be due to the calibration of the measuring devices. Never- cationic dimeric surfactants: 10-s-10 (70, 71), 12-s-12 (12), and
theless the trends are similar, with −HM◦ going through a rather 16-s-16 (12, 63, 72), and one series of anionic dimeric surfac-
shallow minimum at around s = 5–6, i.e., at about the same s tants with phosphate head groups (74). Figure 9 shows the results
value as some of the properties reviewed above. A very large de- for the 12-s-12 surfactants. The chemical trapping method was
crease of −HM◦ is observed in going from 12-2-12 to 12-4-12; used to determine the concentrations of bromide ions at the sur-
−HM◦ depends only weakly on s at s > 4. The large decrease face of 12-s-12 micelles (75). The results showed a decrease
of −HM◦ at low s has been attributed to the conformational in this concentrations upon increasing s, in agreement with the
change discussed in the preceding section (66). results of SANS and conductivity.
The value of the free energy of micellization of a dimeric SANS studies of the cationic dimeric surfactants 16-EOz -16
surfactant, G ◦M , can be obtained from Eq. [2], where α is the showed little variation of α with the spacer length (73, 74). How-
micelle degree of ionization (35): ever a SANS study of 12-EOz -12 showed a large increase of α

G ◦M = 2RT (1.5 − α) ln cmc. [2]


0.7
0.6
A recently reported set of α values (see Ref. (65) and below)
has been used together with available cmc data (12) to calculate 0.5
the values of G ◦M and of the entropy of micellization SM ◦
0.4
α

of 12-s-12 surfactants. The data show that the free energy of 0.3
micellization is nearly independent of s and that most of the

variation of SM and G ◦M with s occurs in going from s = 2 0.2
to s = 4. 0.1
The volume change upon micellization VM◦ of m-s-m sur- 0.0
factants has been determined (67, 68). For 12-s-12 surfactants 0 4 8 12
VM◦ goes through a shallow minimum at s around 5–6 (see
Fig. 8). s
The free energy of micellization per dodecyl chain G ◦ (C12 ) FIG. 9. Variation of the micelle ionization degree α at the cmc with the
for DTAB, 12-3-12, 12-3-12-3-12, and 12-3-12-4-12-3-12 was spacer carbon number for the dimeric surfactants 12-s-12 at 25◦ C (, from
calculated by inserting the available cmc and α values into Ref. 66; , from Ref. 67; , from Ref. 12).
DIMERIC (GEMINI) SURFACTANTS 211

with the spacer length (15). The reason for the difference be-
tween the two studies is unknown. Nevertheless it is noteworthy
that for both the dimeric surfactants A1 with a polymethylene
hydrophobic spacer and A3 with a hydrophilic poly(ethylene
oxide) spacer the increase of α parallels the increase of the sur-
face area occupied by one surfactant at the air/water interface
seen in Fig. 3.
The ionization degree of micelles of cationic oligomeric sur-
factants at the cmc appeared to depend only little on the de-
gree of oligomerization (34). However the conductivity data
that led to this conclusion were analyzed using a not so ac-
curate method (see above). The data should be reanalyzed using
the same method as in Ref. (65) in order to assess the validity
of the above conclusion.

2. Micelle Size and Polydispersity


Fairly complete sets of aggregation numbers (N ) have been FIG. 11. Variation of G 0SC with the spacer carbon number s for 10-s-10
obtained for the 10-s-10 surfactants by SANS (70, 71) and 12-s- surfactants at 23◦ C. Reproduced from Ref. (71) with permission of Springer-
Verlag.
12 surfactants by time-resolved fluorescence quenching (TRFQ)
(76). Figure 10 shows the variation of N with the concentrations
of micellar 10-s-10 surfactant. The different plots are seen to
converge to about the same value of N at the cmc. This value the corresponding monomeric surfactant decyltrimethylammo-
is close to that for the maximum spherical micelle formed by nium bromide, indicating that the 10-s-10 micelles are nearly
spherical at concentrations close to the cmc. The increase of N
with concentration becomes steeper as s is decreased, indicating
an increased tendency toward micelle growth and a change of
micelle shape. The 12-s-12 surfactants showed a very similar
behavior, with an even stronger tendency toward micelle growth
(76). Decreases of N upon increasing s were also evidenced for
other series of dimeric surfactants (63, 72, 74). A decrease of
N upon increasing length of the hydrophilic spacer was also
reported for 12-EOz -12 surfactants (15, 73).
The results in Fig. 10 for the 10-s-10 surfactants were ana-
lyzed in terms of the ladder model for micelle growth (70, 71).
This yielded the values of G 0SC , free energy difference between
N0 dimeric surfactants packed in a part of a cylindrical micelle
and in the maximum spherical micelle of aggregation number
N0 , represented in Fig. 11. The quantity −G 0SC has the largest
value for s = 2, then decreases, is a minimum at s = 10, and
increases again. The minimum nearly corresponds to the s value
for which the micelle aggregation number has the smallest value
at C close to the cmc. This behavior is clearly in relation to the s
dependence of the packing parameter P of 10-s-10 surfactants
(77). Indeed the value s = 10–12 is that for which the surface
area a occupied by one surfactant at the air/water interface is
a maximum (see Fig. 3), and thus P is a minimum. Spherical
micelles are then expected, as experimentally observed.
The effect of the spacer conformation on the aggregation
of dimeric surfactants was investigated for surfactant A5 with
m = 8, and with the phenylenedimethylene spacer having the or-
FIG. 10. Variation of the micelle aggregation number N with the square root tho, meta, or para structure (69). At a concentrations of 2.5 wt%
of the concentration of 10-s-10 micellar surfactant, expressed in mole fraction,
at 23◦ C: 10-2-10 (), 10-3-10 (), 10-4-10 (), 10-6-10 (), 10-8-10 (),
the ortho surfactant formed micelles whereas the meta and para
10-10-10 (), and 10-12-10 (). Reproduced from Ref. (71) with permission surfactants gave rise to only small aggregates of aggregation
of Springer-Verlag. number 3–4.
212 RAOUL ZANA

The trimeric surfactant 12-3-12-3-12 showed a very strong m-s-m surfactant A1 when comparing surfactant having spacers
tendency toward micelle growth, similar to that for 12-2-12 and of nearly equal length, i.e., m-s-m and m-EOz -m with s = 3z +
much stronger than that for 12-3-12 (78). Thus, for these qua- 2. These different behaviors of A1 and A3 micelles emphasize
ternary ammonium oligomeric surfactants, an increase of the once more the importance of the spacer conformation and nature
degree of oligomerization has the same effect as a decrease of on the packing of the surfactant alkyl chains.
the carbon number of the polymethylene spacer.
TRFQ studies of 12-s-12 surfactants clearly showed that the 4. Microstructure of Aqueous Solutions of Dimeric and
polydispersity decreases with the micelle size, i.e., as the mi- Oligomeric Surfactants
celles are less and less elongated (76). Such behavior is in The microstructure of aqueous solutions of dimeric surfac-
agreement with the theoretical prediction that spherical mi- tants A1 and A3 and oligomeric surfactants has been extensively
celles are nearly monodisperse (77). Thus, as dimeric surfac- investigated by means of transmission electronic microscopy
tant micelles grow their shape changes and their polydispersity at cryogenic temperature (cryo-TEM) (4, 15, 34, 76, 80–82).
increases. These studies showed that dimeric and oligomeric surfactants
3. Micropolarity and Microviscosity of Oligomeric Surfactant with short spacers always form aggregates that are less curved
Micelles than those formed by the corresponding monomeric surfactants.
This is expected from the fact that for oligomeric surfactants
Figure 12 shows that the micropolarity of 12-s-12 micelles, with short spacers the surface area per amphiphilic moiety is
as measured by the pyrene polarity ratio I1 /I3 , goes through a lower for the dimer than for the monomer and lower for the
maximum at s around 5 (79). However, the micropolarity was trimer than for the dimer (34).
nearly independent of the degree of oligomerization in the series Cryo-TEM investigations showed that 12-2-12 forms thread-
DTAB, 12-3-12, 12-3-12, and 12-3-12-4-12-3-12 (79). like micelles that can be several micrometers long and that are
Figure 12 also shows that the microviscosity of 12-s-12 mi- already entangled at a concentration of 2 wt% (4, 76, 80). How-
celles, as determined using the fluorescent probe dipyrenyl- ever this surfactant forms spherical micelles at 0.2 wt%. Shorter
propane, decreases rather rapidly at s > 4 (34, 79). A similar but still elongated micelles were also visualized in a 7 wt% solu-
result was reported for the surfactants 16-s-16 (63). These stud- tion of 12-3-12 but only spheroidal micelles were visualized in
ies reported that the microviscosity of m-s-m surfactant micelles a 2 wt% solution of this surfactant (4, 76). By comparison, the
is always larger than that of the micelles of the corresponding micelles of DTAB, which can be considered as the monomer of
monomeric surfactants, m-s/2. The microviscosity increased 12-2-12 and 12-3-12, remain spherical micelles at fairly high
strongly and nearly linearly in going from the monomer (DTAB) concentrations and ionic strength. The solutions of 12-4-12,
to the dimer (12-3-12), trimer (12-3-12-3-12), and tetramer (12- 12-8-12, and 12-12-12 showed only densely packed spheroidal
3-12-4-12-3-12) (34). micelles at 5–7 wt% (4, 76). Such changes of microstructure
The microviscosity of micelles of 16-EOz -16 surfactant A3 with s are in agreement with the results concerning the variation
showed no clear trend upon increasing number of ethylene oxide, of the aggregation number of 12-s-12 micelles with concentra-
z (73). Nevertheless the microviscosity of micelles of m-EOz -m tions (76). Vesicles were visualized in solutions of 12-16-12 and
surfactant A3 was found to be larger than that of micelles of 12-20-12 (76). The surfactants 12-8 and 12-10 (quaternary am-
monium surfactants with two unequal alkyl chains), which are
the corresponding monomers of 12-16-12 and 12-20-12, also
formed vesicles but at higher concentrations (76). Thus the in-
crease of s for the 12-s-12 series results in the unusual sequence
of structures:

Elongated micelles → spheroidal micelles → vesicles. [4]

The cryo-TEM micrographs in Fig. 13 show that the 12-EOz -


12 surfactants form spheroidal micelles when z = 4 (n T = 14)
and worm-like micelles when z = 1 (n T = 5) (15). Recall that
12-s-12 surfactants form spherical micelles for s = 5. The dif-
ference between 12-EOz -12 and 12-s-12 surfactants can be ex-
plained on the basis of the results in Fig. 3. At a given value of s
or n T the surface area occupied by one 12-EOz -12 surfactant is
FIG. 12. Variations of the micropolarity () and of the relative microvis- lower than that for 12-s-12 surfactants. Thus aggregates of lower
cosity () of 12-s-12 micelles with the spacer carbon number at 25◦ C. The
micelle polarity is represented by the pyrene polarity ratio I1 /I3 ; the values of
curvature are expected with the former on the basis of packing
the micelle microviscosity are referred to the microviscosity of the micelles of parameter considerations, in agreement with the experimental
the conventional surfactant DTAB. Data from Ref. (79). observations.
DIMERIC (GEMINI) SURFACTANTS 213

FIG. 13. Cryo-TEM images of 10 wt% solutions of 12-EOz -12 surfactants quenched from 25◦ C: (left) only spherical micelles are seen with the 12-EO4 -12
surfactant (long spacer); (right) worm-like micelles are seen with the 12-EO-12 surfactant (short spacer). Reproduced from Ref. (15) with permission of the
American Chemical Society.

Cryo-TEM investigation of the 16-s-16 surfactants revealed the structures because the two alkyl chains of a dimeric surfactant
following sequence of structures upon increasing s (76): can take different relative orientations in the micelles.

Vesicles + elongated micelles → elongated micelles 5. Rheological Behavior of Solutions of Oligomeric Surfactant
→ spheroidal micelles. [5] Studies of the effect of the spacer on the rheological behavior
of solutions of dimeric and oligomeric surfactants are still scarce.
Elongated micelles were observed for s = 3 and 4. Spheroidal Most studies used the surfactant 12-2-12 because its solution
micelles persisted up to s = 12 (72), and 16-s-16 surfactants shows a very pronounced increase of viscosity at concentrations
with s > 12 are expected to form vesicles but have not been above 2 wt% as can be seen in Fig. 14 (84) and shear-thickening
investigated yet. in the concentration range between 0.7 and 1.5 wt% (3).
Early cryo-TEM investigations of the trimeric surfactant Figure 14 illustrates the peculiar rheological behavior of
12-3-12-3-12 (81) revealed the first branched worm-like mi- oligomeric surfactants with short spacers. The steep and very
celles ever observed with a surfactant in the absence of any large increase of the zero shear viscosity η with the surfactant
additive. A later investigation of 12-2-12 solution using an im- volume fraction  at above some volume fraction ∗ was
proved cryo-TEM showed branched worm-like micelles (80). interpreted for 12-2-12 as the onset of fast micellar growth
The trimeric surfactant 12-6-12-6-12 showed only spherical mi- and of the semidilute regime (84). Above ∗ the worm-like
celles (34), just like the corresponding dimeric surfactant 12-6- micelles are entangled. Several explanations were proposed for
12. Such is not the case for the shorter spacer s = 3, with long and the presence of a maximum in the η vs  plot: (i) true decrease
branched worm-like micelles for 12-3-12-3-12 and only slightly of micelle length associated with complex electrostatic effects
elongated micelles for 12-3-12 at 3 wt% (see above). This shows involving the counterions (84), (ii) strong correlations of orien-
the complex relationship between spacer carbon number and tation between worm-like micelles (84), and (iii) branching of
degree of oligomerization. Cryo-TEM showed that solutions of the worm-like micelles (83). The third explanation comes from
the tetrameric surfactant 12-3-12-4-12-3-12 contain a significant molecular dynamics simulations. Self-diffusion studies of vari-
fraction of surfactant under the form of closed ring micelles (34, ous dimeric surfactants showed that the electrostatic effects must
82). A similar structure was also later observed with 12-2-12 be taken into account to give a correct description of micelle
solutions (80). growth (85).
Molecular dynamics simulations of aqueous solutions of m- The results in Fig. 14 (34, 84) clearly show the importance
s-m surfactants predicted the experimentally observed change of the value of the spacer carbon number (compare the plots for
of micelle shape from spheroidal to elongated and the formation 12-2-12 and 12-3-12) and of the surfactant degree of oligomer-
of branched thread-like micelles upon decreasing spacer carbon ization (compare the plots for 12-3-12, 12-3-12-3-12, and
number (83). Dimeric surfactants are capable of forming such 12-3-12-4-12-3-12) on the zero-shear viscosity. These results
214 RAOUL ZANA

104 surfactant at a time into/from its micelles, similarly to conven-


tional monomeric surfactants at low concentrations (87). The
103 results showed that the micelle formation/breakup process be-
comes faster as the spacer carbon number is increased. Never-
102
theless, the lifetime of dimeric surfactant micelles can be very
long, in the range of tens of seconds.
η0 ((Pa.s)

101
The rapid slowing down of micellar equilibria with increasing
100 hydrophobicity of the alkyl chains of the dimeric surfactants was
confirmed by recent NMR studies of surfactants such as 14-2-14,
10-1 surfactant A7, and surfactant A8 with s = 2 and m + m  > 28
10-2 (20, 89). Separate signals were observed for the free and micellar
surfactants indicating slow surfactant exchanges on the NMR
10-3 time scale.
The time for the breakup of a worm-like micelle into two
10-1 100 101 102 daughter micelles has been evaluated from rheological studies
of 12-2-12 micelles. It has been found to be between 0.1 and
C (wt %) 10 s, under the experimental conditions used (84).
FIG. 14. Variation of the zero shear viscosity with the surfactant weight
percentage at 25◦ C for aqueous solutions of 12-2-12 (), 12-3-12 (), 7. Solubilization by Solutions of Dimeric Surfactants
12-3-12-3-12 (), and 12-3-12-4-12-3-12 (). Reproduced from Refs. (34,
84) with permission of the American Chemical Society. Few studies of the solubilization of water-insoluble com-
pounds in solutions of dimeric surfactants and of the correspond-
ing monomeric surfactants have been reported thus far despite
the obvious interest of such studies for future uses of dimeric
correlate very well with those concerning the micelle aggre-
surfactants.
gation number of the same surfactants (See Section V.2). The
Dam et al. (90) compared the solubilization capacity (number
surfactants that formed branched micelles (for instance 12-3-12-
of moles of solubilized compound per mole of micellar surfac-
3-12) were found to show a peculiar concentration dependence
tant) of a few conventional and A1 dimeric surfactants. They
of the elastic modulus (86).
reported that dimeric surfactants have a larger solubilizing ca-
pacity than conventional surfactants for oils such as toluene and
6. Dynamics of Dimeric Surfactant Micelles
n-hexane on both a molar basis and a weight basis. They also
Micelles are dynamic objects that are constantly exchanging showed that the m-s-m cationic surfactants prefer toluene to
surfactants with the intermicellar solution (surfactant exchange n-hexane. The solubilizing capacity of 12-s-12 surfactants de-
process) and that are forming and breaking up (87). This of creased progressively as s was increased from 2 to 6 and 10.
course also applies to dimeric surfactants. Intermediate surfactants were not investigated.
An ultrasonic relaxation study of the short chain surfactants Devinsky et al. (91) performed the most comprehensive study
8-3-8 and 8-6-8 showed that the entry of the surfactant into of solubilization by solutions of m-s-m dimeric surfactants at
its micelles occurs in one step and is diffusion-controlled (68), 25◦ C. The solubilizing capacity for trans-azobenzene increased
as for conventional monomeric surfactants (87). The exit rate linearly with m for a series of dimeric surfactants of given s
constants of these two surfactants from their micelles showed value, a behavior similar to that for conventional surfactants.
little dependence on s. More interestingly, the solubilization capacity was a maximum
The kinetics of micellar equilibria for the longer chain 12-s-12 at about s = 6 for the 10-s-10 surfactants (see Fig. 15). This
surfactants, with s = 2, 3, 4, and 6, were investigated by the value of s is close to that where the cmc is a maximum, possibly
pressure-jump technique, in the concentration range between as a result of conformational change and increased hydropho-
the cmc and the 3cmc (88). The results showed that the entry bicity of the spacer (See Section V.2).
rate constant of a dimeric surfactant into its micelles may be A study of the solubilization of styrene by micellar solution
up to a 100 times smaller than that for a diffusion process. The of 12-s-12 did not confirm this effect of the spacer length. The
results also revealed an increase of the entry rate constant with solubilizing capacity increased progressively by a factor of about
the spacer carbon number. This rather unexpected result is prob- 5 as s increased from 2 to 12 and temperature had a strong effect
ably in relation to the conformational change of the surfactant (92). This effect of the spacer length is opposite to that reported
molecule that occurs as the spacer carbon number is increased by Dam et al. (90) for the solubilization of toluene and n-hexane
from 2 to 4. The kinetics of the micelle formation/breakup pro- also in 12-s-12 micellar solutions.
cess were also investigated. The variation of the corresponding The solubilizing capacity of 12-EOz -12 surfactants for styrene
relaxation time with the surfactant concentrations indicated that showed a maximum at n T = 5; i.e., z = 1 (15). This value
this process takes place via stepwise entry/exit of one dimeric of n T corresponds to a spacer of length nearly equal to that
DIMERIC (GEMINI) SURFACTANTS 215

0.25 effect of the nature of the spacer on the interaction. Also struc-
tural studies of dimeric surfactant/polyelectrolyte complexes in
the solid state should be performed in view of the extremely
0.20 interesting results obtained with conventional surfactant/poly-
electrolyte complexes.
SC (M/M)

0.15
VI. PHASE BEHAVIOR

0.10 Most studies of phase behavior concerned m-s-m surfactants


(45, 97, 98). No thermotropism was observed with the pure 12-s-
0.05
12 surfactants, contrary to the corresponding monomeric surfac-
tants (45) and to the 12-EOz -12 surfactants (58). This behavior
was attributed to geometric constraints on the head group ar-
0.00 rangement associated with the presence of the spacer. The longer
0 2 4 6 8 10 12 14 16-m-16 surfactants showed thermotropism (97).
s The effect of the spacer length on the occurrence of ly-
otropic mesophases was investigated for the 12-s-12/water mix-
FIG. 15. Dependence of the solubilizing capacity of micellar solutions of tures by X-ray scattering. Cylindrical (hexagonal) and lamellar
10-s-10 surfactants for trans-azobenzene as a function of the spacer carbon mesophases were found to occur in a concentration range that
number s. Surfactant concentration, 20 mM. Prepared from the results reported
narrowed as s increased and completely disappeared for s = 10
in Ref. (91).
and 12 (45). For these surfactants the micellar range extends
to concentrations as high as 90 wt% as for the corresponding
of a polymethylene spacer with s = 5 where the solubiliz- monomeric surfactants. This behavior was attributed to a maxi-
ing capacity of 10-s-10 surfactants for trans-azobenzene is a mum mismatch between the length of the spacer and that of the
maximum. dodecyl chains. It is readily explained on the basis of packing
Obviously the results just reviewed show diverging behaviors parameter considerations. Indeed, the value of the surface area
when the spacer length and nature are changed. More studies are occupied by one 12-s-12 surfactant at the air–water interface
required in order to get a better understanding of the solubiliza- is a maximum at s = 10–12 (see Fig. 3). Thus, this value of
tion by solutions of surfactant dimers. Also studies of solubiliza- s corresponds to the lowest value of the packing parameter for
tion of representative oils by a series of oligomeric surfactants 12-s-12 surfactants and to a maximum stability of the spheroidal
of increasing oligomerization degree are still lacking. micellar structure. Lyotropic mesophases were again observed
for the 12-16-12 surfactant with the texture of the conventional
8. Interaction of Dimeric and Oligomeric Surfactants
lamellar and hexagonal phases. The alkyl chains are located in-
with Water-Soluble Polymers
side the cylinders of the hexagonal phase. The spacer group was
The study of the interaction of dimeric and oligomeric surfac- shown to lie nearly flat and in a fairly extended conformation
tants with water-soluble polymers is important in view of a future at the core/water interface in both the lamellar and hexagonal
use of these compounds in formulations. Indeed, surfactant- phases, up to s = 8 (45). Raman spectroscopy confirmed this
based formulations often contain water-soluble polymers intro- result for m-6-m surfactants (98). Intermediate and/or bicontin-
duced to improve the formulation properties. uous cubic phases, in addition to lamellar and hexagonal phases,
The reported studies involved m-s-m dimeric surfactants and were evidenced in a study of 16-s-16 surfactants (97).
several water-soluble polymers (93–96). The 12-s-12 surfac- The phase behavior of 12-EOz -12 surfactants has been re-
tants were shown to interact weakly with hydroxypropylguar ported (58). The mesophases were generally easier to observe
and strongly with hydrophobically modified hydroxypropylguar than were the mesophases with the 12-s-12 surfactants. The mi-
(93). The corresponding monomer, DTAB, hardly interacted cellar range was found to become wider as the number z of
with these two polymers. The trimeric surfactant 12-3-12-3-12 ethylene oxide groups was increased.
interacted more strongly with these polymers than the dimeric The partial phase diagrams of the ternary systems water/12-
surfactant 12-3-12 (93). EOz -12/styrene (15) and water/12-s-12/styrene have been re-
The study of the interaction between m-s-m surfactants and ported (92). The studies focused on the effect of the spacer length
oppositely charged polyelectrolytes, poly(disodium maleate- and nature on the extent of the one-phase region (microemul-
alt-alkylvinylether) (94) and hyaluronic acid (95, 96), showed sion range) in the systems. Figure 16 shows that the extent of the
that, as for neutral polymers, dimeric surfactants interact more microemulsion range is a maximum at s = 10 for the water/12-
strongly than monomeric surfactants with the polymers. s-12/styrene system (92) and that it strongly depends on tem-
More studies of the interaction of dimeric and oligomeric perature. The one-phase region for the water/12-EOz -12/styrene
surfactants are required in order to get a precise picture of the systems has a maximum extension at n T = 5, i.e., with the
216 RAOUL ZANA

FIG. 16. Partial phase diagrams of the systems water/12-s-12/styrene (a–f) and of the system water/12-EO2 -12/styrene (g) at 25 and 60◦ C. 1 and 2 denote
one-phase and two-phase regions. Reproduced from Refs. (15) and (92) with permission of the American Chemical Society.
DIMERIC (GEMINI) SURFACTANTS 217

12-EOz -12 surfactant, and depends only little on temperature intermediate structures, such as bilayer membrane fragments
(15). The dependence on temperature for the two types of sys- and/or giant thread-like micelles usually seen during such a
tems can be seen in Fig. 16 by comparing the partial phase transformation, were observed. Addition of the thread-like form-
diagrams of the water/12-8-12/styrene (Fig. 16d) and water/12- ing surfactant 12-2-12 to vesicular suspensions of 12-20-12
EO2 -12/styrene (Fig. 16g) systems. Note that the surfactants (103) had a very different effect. The vesicle size first increased
12-8-12 and 12-EO2 -12 have spacers of about equal length with at very low 12-2-12 content. This was followed by vesicle break-
both s and n T equal to 8. age into smaller vesicles and then the formation of disk-like mi-
Theoretical calculations indicated that an increased rigidity celles, and then ring-like micelles and short thread-like micelles.
of the spacer induces a transition from a three-phase system to At still higher 12-2-12 concentration the thread-like micelles
a two-phase system (99). became longer and the final structure was that of a network of
connected thread-like micelles containing a few isolated closed
VII. MISCELLANEOUS STUDIES ring micelles. Metastable ribbon-like structures were evidenced
at intermediate 12-2-12 content. The equilibrium structure in
1. Mixed Micellization of Dimeric Surfactants with this range was reached only after several weeks.
Conventional Surfactants
2. Polymerization of Styrene in Microemulsions Based on
Such studies were performed with the hope of observing syn- Dimeric Surfactants
ergism that would make more attractive the use of dimeric sur-
factants in formulation and also for the sake of understanding Styrene was polymerized in microemulsions based on 12-s-12
mixed micellization in such systems. and 12-EOz -12 dimeric surfactants (15, 92). The molecular
A SANS study of mixtures of CTAB and 16-s-16 surfac- weight of the polystyrene synthesized in the two types of mi-
tants showed that the micelle size and the extent of micellar croemulsions was a maximum at s = 6 (92) or z = 2–3 (15),
growth could be modulated by the fraction of dimeric surfactant in the absence of a cross-linker, at 25◦ C (see Fig. 17). In the
in the mixture (100). A cryo-TEM study of the comicelliza- presence of a cross-linker the maximum molecular weight was
tion of DTAB and 12-2-12 revealed that the thread-like micelles obtained at s = 10 (92) or n T = 2 (15). The maximum disap-
formed by 12-2-12 are rapidly transformed into spheroidal mi- peared at 60◦ C. The experimental results were discussed in terms
celles, already at a DTAB mole fraction of 0.3 (4). A SANS study of the spontaneous curvature of the surfactant assemblies.
showed a rather uniform distribution of DTAB in the thread-like
3. Biological Effects
micelles and no detectable accumulation of DTAB in their end-
caps, thereby explaining the cryo-TEM observation (101). Bisquaternary ammonium surfactants can induce various bi-
Additions of the spherical micelle-forming surfactants DTAB ological effects. The inhibition of bacterial activity particularly
and 12-10-12 to vesicular suspensions of 12-20-12 were shown has been investigated (7, 59, 104–106). In general the spacer
by the cryo-TEM study to result in the progressive transforma- length was found to have little effect on the minimum inhibitory
tion of the vesicles into mixed spheroidal micelles (102). No concentration for several bacteria.

FIG. 17. Variation of the weight average molecular weight of the polystyrene obtained by polymerization of styrene in the microemulsion systems (a) water/12-
s-12/styrene as a function of s and (b) water/12-EOz -12/styrene as a function of z at 25◦ C in the absence of cross-linker. WR is surfactant-to-monomer weight
ratio. Reproduced from Refs. (15) and (92) with permission of the American Chemical Society.
218 RAOUL ZANA

TABLE 2
Summary of the Properties Reviewed and of the Observed Behaviors for Surfactants A and A3

Property reviewed Observed behavior

Premicellar association Observed at s ≥ 12 for 12-s-12 and at s ≥ 4 for 16-s-16 (33).


Behavior at air/solution interface Surface occupied by one surfactant: is a maximum at s = 10–12 for 12-s-12 (26, 30); increases progressively with z
for 12-EOz -12 (15).
Behavior at silica/solution interface Maximum amount of adsorbed 12-s-12 decreases as s increases (50–52).
Structure of the surface aggregates: bilayer at s = 2; cylinders at s = 4 and 6; unknown at higher s (54).
Krafft temperature No clear trend for 12-s-12; may be a minimum at s around 5 for 16-s-16 (56, 57).
Melting temperature Maximum at s = 5, irrespective of alkyl chain length for m-s-m surfactants (56).
cmc Maximum at s = 5, irrespective of alkyl chain length for m-s-m surfactants (1, 2).
Continuous increase with z for 12-EOz -12 (15).
Enthalpy of micellization H ◦ M −H ◦ M is a minimum at s around 6 for 12-s-12 (64–66).
Volume of micellization V ◦ M Minimum at s = 5–6 for 12-s-12 (67).
Micelle ionization degree Progressive increase with s for 12-s-12 and with z for 12-EOz -12 (15, 66).
Micelle size Decreases as s increases up to s = 10–12 and then increases for 12-s-12 (76). Continuous decrease with z up to
z = 5 for 12-EOz -12 (15).
Micelle shape Worm-like micelles for s = 2 and 3 and spheroidal micelles for s ≥ 4 for 12-s-12 (76). Worm-like micelles for
z = 1 and spheroidal micelles for z = 4 for 12-EOz -12 (15).
Tendency toward micelle growth Decreases as s increases up to 10 and then increases for 10-s-10 surfactants (70, 71).
Micelle micropolarity Maximum at s around 5 for 12-s-12 (79).
Micelle microviscosity Inflection at s around 5 for 12-s-12 (79).
Rheology Non-Newtonian behavior at C > 2 wt% for 12-2-12 and >5 wt% for 12-3-12 (34, 84).
Dynamics Entry rate constant of 12-s-12 surfactants in micelles increases with s (88).
Solubilization Variation of solubilizing power with s very dependent on the solubilizate. Maximum observed for 10-s-10 at s = 5
(91), not for 12-s-12 or 12-EOz -12 (15, 92).
Phase behavior Extension of the micellar range: maximum at s = 10–12 for 12-s-12 (45); increases with z for 12-EOz -12 (15).
Microemulsion with styrene Extension of the microemulsion range: maximum at s = 10 for 12-s-12 (92) and at z = 1 for 12-EOz -12 (15).
Polymerization in microemulsion Polystyrene of largest molecular weight and smallest radius obtained at s = 6 for 12-s-12 (92) and at z = 2–3 for
12-EOz -12 (15).

SUMMARY AND GENERAL REMARKS and the alkyl chain length that lowers the stability of the surfac-
tant in the solid state. The value s = 10–12 also corresponds to
Table 2 summarizes many of the reviewed results, focusing the maximum in the plot of the surface area occupied by one
on the surfactant series A1 and A3. Obviously several proper- dimeric surfactant at the air/water interface (see Fig. 3). On the
ties of the m-s-m surfactant A1 show a special behavior for the basis of surfactant packing parameter we noted that this value of
values s = 5–6 and s = 10–12. The first value of s corresponds s corresponds to the maximum stability of the micellar phase
to a maximum of the melting temperature of the solid surfac- for m-s-m surfactants, as experimentally observed. For the
tant, a maximum of cmc, a minimum of some thermodynamic same reason it corresponds to the minimum tendency to-
quantities associated with micelle formation, and a maximum ward micelle growth as also experimentally observed. Last,
of micelle micropolarity. The maximum of melting temperature it is also responsible for the maximum extension of the mi-
indicates that the stability of the m-s-m surfactants in the liquid croemulsion range for the ternary system water/12-s-12/styrene
state is the lowest for s = 5. Likewise, the maximum of cmc (92).
at s = 5–6 indicates that owing to different effects the stability In several respects the 12-EOz -12 surfactants show a behavior
of the surfactant in the micellar state is the lowest. The simi- different from that of the 12-s-12 surfactants. This is so because
larity between surfactants in the micellar state and in the liquid already at z = 1 the corresponding spacer, CH2 CH2 OCH2 CH2 ,
(melted) state has been often emphasized, particularly when dis- is already long enough to allow the two alkyl chains to be in
cussing the state of the alkyl chain within the micelle core. This a cis or a gauche conformation. In other words the change in
similarity provides a qualitative explanation for the correlation conformation of the 12-s-12 molecule that probably occurs as s
between the variations of cmc and melting temperature with s for is increased from 2 to 5–6 does not take place upon increasing z
m-s-m surfactants. The thermodynamic quantities are all deriva- with 12-EOz -12 surfactants. This explains that the cmc increases
tives of the cmc with respect to some variables. It is therefore not progressively with the spacer length. Nevertheless, the two se-
surprising that they show a peculiar behavior at s = 5. Turning ries of surfactants present some similarities. For instance, the
to the value s = 10–12, the minimum of melting temperature is micelle size decreases with increasing spacer length, reflecting
associated with a maximum mismatch between the spacer length the decrease of the packing parameter.
DIMERIC (GEMINI) SURFACTANTS 219

CONCLUSIONS 18. Rozycka-Roszak, B., Witek, S., and Przestalski, S., J. Colloid Interface
Sci. 131, 181 (1989).
This paper reviewed the effect of the length, nature, and flex- 19. Rozycka-Roszak, B., Fisicaro, E., and Ghiozzi, A., J. Colloid Interface
ibility of the spacer group on selected properties of dimeric and Sci. 184, 209 (1996).
20. Huc, I., and Oda, R., Chem. Commun. 2025 (1999).
oligomeric surfactants. These surfactants display properties that 21. Oda, R., Huc, I., and Candau, S. J., Chem. Commun. 2105 (1997).
are often superior to those of the corresponding conventional 22. Oda, R., Huc, I., Homo, J.-C., Heinrich, B., Schmutz, M., and Candau,
monomeric surfactants. However most of the gain in property S. J., Langmuir 15, 2384 (1999).
is achieved in going from the monomer to the dimer. The prop- 23. Devinsky, F., and Lacko, I., Tenside Deterg. 27, 344 (1990).
erties still improve in going from the dimer to the trimer, but 24. Pinazo, A., Diz, M., Solans, C., Pés, M. A., Era, P., and Infante, M. R.,
J. Am. Oil Chem. Soc. 70, 37 (1993).
the synthesis of oligomeric surfactants is much more difficult
25. Devinsky, F., Masarova, L., and Lacko, I., J. Colloid Interface Sci. 105,
and costly than that of dimeric surfactants. Oligomeric surfac- 235 (1985), and references therein.
tants may thus remain topics of interest for only the academic 26. Espert, A., v. Klitzing, R., Poulin, P., Colin, A., Zana, R., and
community. In the field of applications the most interesting prop- Langevin, D., Langmuir 14, 1140 (1998).
erties of dimeric surfactants are their much lower cmc values, 27. Pinazo, A., Infante, M. R., Chang, C.-H., and Franses, E. I., Colloids Surf.
their stronger efficiency at decreasing the surface tension of wa- A 87, 117 (1994).
28. Takemura, T., Shiina, M., Izumi, M., Nakamura, K., Miyazaki, M.,
ter and the interfacial tension of oil/solution interfaces, and their Torigoe, K., and Esumi, K., Langmuir 15, 646 (1999).
stronger adsorption at solid/solution interfaces than for the corre- 29. Esumi, K., Taguma, K., and Koide, Y., Langmuir 12, 4039 (1996).
sponding monomeric surfactants. Also dimeric surfactants with 30. Alami, E., Beinert, G., Marie, P., and Zana, R., Langmuir 9, 1465 (1993).
short spacers can give rise to linear and branched thread-like 31. Li, Z. X., Dong, C. C., and Thomas, R. K., Langmuir 15, 4392 (1999).
micelles as well as closed ring micelles at fairly low concentra- 32. Eastoe, J., Nave, S., Downer, A., Paul, A., Rankin, A., Tribe, K., and
Penfold, J., Langmuir 16, 4511 (2000).
tions. These structures bring about interesting rheological prop- 33. Zana, R., J. Colloid Interface Sci. 246, 182 (2002).
erties. Changing the length, nature (hydrophilic, hydrophobic), 34. In, M., Bec, V., Aguerre-Chariol, O., and Zana, R., Langmuir 16, 141
and flexibility of the spacer offers the possibility to modulate (2000).
very efficiently the properties of solutions of dimeric surfac- 35. Zana, R., Langmuir 12, 1208 (1996).
tants. The enormous variety in the structures of dimeric surfac- 36. Hattori, N., Yoshino, A., Okabayashi, H., and O’Connor, C. J., J. Phys.
Chem. B 102, 8965 (1998).
tants that can be synthesized will no doubt result in the synthesis 37. Song, L., and Rosen, M. J., Langmuir 12, 1149 (1996).
of low cost dimeric surfactants with improved properties and in 38. Rosen, M. J., and Liu, L., J. Am. Oil Chem. Soc. 73, 885 (1996).
their increased uses. New regulations that are enacted in the 39. Pinazo, A., Wen, X., Pérez, L., and Infante, M. R., Langmuir 15, 3134
developed world in relation to toxicity of surfactants and ecol- (1999).
ogy may accelerate a move toward an increased use of dimeric 40. Masuyama, A., Yokota, M., Zhu, Y.-P., Kida, T., and Nakatsuji, Y.,
J. Chem. Soc. Chem. Commun. 1435 (1994).
surfactants. 41. Sumida, Y., Oki, T., Mayusama, A., Maekawa, H., Nishiura, M., Kida, T.,
Nakatsuji, Y., Ikeda, I., and Nojima, M., Langmuir 14, 7450 (1998).
REFERENCES 42. Menger, F. M., and Wrenn, S., J. Phys. Chem. 78, 1387 (1974).
43. Pérez, L., Pinazzo, A., Rosen, M. J., and Infante, M. R., Langmuir 14,
1. Zana, R., in “Novel Surfactants: Preparation, Applications and Biodegrad- 2307 (1998).
ability” (C. Holmberg, Ed.), Chap. 8, p. 241. Dekker, New York, 1998. 44. Zhu, Y.-P., Masuyama, A., Kobata, Y., Nakatsuji, Y., Okahara, M., and
2. Zana, R., Adv. Colloid Interface Sci., in press (2002). Rosen, M. J., J. Colloid Interface Sci. 158, 40 (1993), and references
3. Schmitt, V., Schosseler, F., and Lequeux, F., Europhys. Lett. 30, 31 therein.
(1995). 45. Alami, E., Lévy, H., Zana, R., and Skoulios, A., Langmuir 9, 940
4. Zana, R., and Talmon, Y., Nature 362, 228 (1993). (1993).
5. Danino, D., Kaplun, A., Talmon, Y., and Zana, R., in “Structure and Flow 46. Diamant, H., and Andelman, D., Langmuir 10, 2910 (1994); 11, 3605
in Surfactant Solutions” (C. A. Herb and R. K. Prud’homme, Eds.), ACS (1995).
Symp. Ser. 578, Chap. 6, p. 105. Am. Chem. Soc., Washington, DC, 1994. 47. Maiti, P. K., and Chowdhury, D., Europhys. Lett. 41, 183 (1998); J. Chem.
6. Rosen, M. J., Chem. Technol. 30 (1993). Phys. 109, 5126 (1998).
7. Devinsky, F., Lacko, I., Mlynarcik, D., Racansky, V., and Krasnec, L., 48. Gao, T., and Rosen, M. J., J. Am. Oil Chem. Soc. 71, 771 (1994).
Tenside Deterg. 22, 1 (1985). 49. Rosen, M., and Song, L. D., J. Colloid Interface Sci. 179, 261 (1996).
8. Zana, R., and In, M., Uzbek. J. Phys. 1, 24 (1999). 50. Chorro, C., Chorro, M., Dolladille, O., Partyka, S., and Zana, R., J. Colloid
9. Rosen, M. J., and Tracy, D. J., J. Surf. Deterg. 1, 547 (1998). Interface Sci. 199, 169 (1998).
10. Menger, F. M., and Keiper, J. S., Angew. Chem. Int. Ed. 39, 1906 (2000). 51. Chorro, C., Chorro, M., Dolladille, O., Partyka, S., and Zana, R., J. Colloid
11. Fisicaro, E., Compari, C., Rozycka-Roszak, B., Viscardi, G., and Interface Sci. 210, 134 (1999).
Quagliotto, P. L., Curr. Top. Colloid Interface Sci. 2, 53 (1997). 52. Grosmaire, L., Doctorat de l’Université Montpellier 2, University of Mont-
12. Zana, R., Benrraou, M., and Rueff, R., Langmuir 7, 1072 (1991). pellier 2, Montpellier, France, July 2001, and manuscript in preparation.
13. Devinsky, F., Lacko, I., Bittererova, F., and Tomeckova, L., J. Colloid 53. Esumi, K., Goino, M., and Koide, Y., J. Colloid Interface Sci. 183, 539
Interface Sci. 114, 314 (1986), and references therein. (1996).
14. Rosen, M. J., Mathias, J. H., and Davenport, L., Langmuir 15, 7340 (1999). 54. Manne, S., Schäffer, T. E., Huo, Q., Hansma, P. K., Morse, D. E., Stucky,
15. Dreja, M., Pyckhout-Hintzen, W., Mays, H., and Tiecke, B., Langmuir 15, G. D., and Aksay, I. A., Langmuir 13, 6382 (1993).
391 (1999). 55. Fielden, M. L., Claesson, P. M., and Verrall, R. E., Langmuir 15, 3924
16. Menger, F. M., Keiper, J. S., and Azov, V., Langmuir 16, 2062 (2000). (1999).
17. Menger, F. M., and Littau, C. A., J. Am. Chem. Soc. 115, 10083 (1993). 56. Zana, R. 246, 175 (2002).
220 RAOUL ZANA

57. Zhao, J., Christian, S. D., and Fung, B. M., J. Phys. Chem. B 102, 761 82. In, M., Aguerre-Chariol, O., and Zana, R., J. Phys. Chem. B. 103, 7747
(1998). (1999).
58. Dreja, M., Gramberg, S., and Tieke, B., Chem. Commun. 1371 (1998). 83. Karaborni, S. K., Esselink, P. A., Hilbers, B., Smit, J., Karthauser, van Os,
59. Pérez, L., Torres, J. L., Manresa, A., Solans, C., and Infante, M. R., Lang- N. M., and Zana, R., Science 266, 254 (1994).
muir 12, 5296 (1996). 84. Kern, F., Lequeux, F., Zana, R., and Candau, S. J., Langmuir 10, 1714
60. Camesano, T. A., and Nagarajan, R., Colloids Surf. 167, 165 (2000). (1994).
61. Zhu, Y.-P., Ishahara, K., Masuyama, A., Nakatsuji, Y., and Okahara, M., 85. Narayanan, J., Urbach, W., Langevin, D., Manohar, C., and Zana, R., Phys.
J. Jpn. Oil Chem. Soc. 42, 1611 (1993). Rev. Lett. 81, 228 (1998).
62. Devinsky, F., Lacko, I., and Imam, T., Acta Fac. Pharm. 44, 103 (1990). 86. In, M., Warr, G. W., and Zana, R., Phys. Rev. Lett. 83, 2278 (1999).
63. De, S., Aswal, V. K., Goyal, P. S., and Bhattacharya, S., J. Phys. Chem. 87. Zana, R., in “Encyclopedia of Surface and Colloid Science” (A. Hubbard,
100, 11664 (1996). Ed.). Dekker, New York, 2002.
64. Bai, B., Yan, H., and Thomas, R. K., Langmuir 17, 4501 (2001). 88. Zana, R., and Ulbricht, W., Colloids Surf. A 183–185, 487 (2001).
65. Bai, B., Wang, J., Yan, H., Li, Z., and Thomas, R. K., J. Phys. Chem. 89. Oda, R., Huc, I., Danino, D., and Talmon, Y., Langmuir 16, 9759 (2000).
B 105, 3105 (2001). 90. Dam, Th., Engberts, J. B. F. N., Karthauser, J., Karaborni, S., and van Os,
66. Grosmaire, L., Chorro, M., Chorro, C., Partyka, S., and Zana, R., J. Colloid N. M., Colloids Surf. A 118, 41 (1996).
Interface Sci., in press. 91. Devinsky, F., Lacko, I., and Imam, T., J. Colloid Interface Sci. 143, 336
67. Verrall, R. E., and Wettig, S. D., J. Colloid Interface Sci. 235, 301 (1991).
(2001). 92. Dreja, M., and Tieke, B., Langmuir 14, 800 (1998).
68. Frindi, M., Michels, B., Lévy, H., and Zana, R., Langmuir 10, 1140 (1994). 93. Kastner, Ü., and Zana, R., J. Colloid Interface Sci. 218, 468 (1999).
69. Hattori, N., Hirata, H., Okabayashi, H., and O’Connor, C. J., Colloid 94. Zana, R., and Benrraou, M., J. Colloid Interface Sci. 226, 286 (2000).
Polym. Sci. 277, 361 (1999). 95. Pisarcik, M., Soldan, M., Bakos, D., Devinsky, F., and Lacko, I., Colloids
70. Hirata, H., Hattori, N., Ishida, M., Okabayashi, H., Furusaka, M., and Surf. A 150, 207 (1999).
Zana, R., J. Phys. Chem. 99, 17778 (1995). 96. Pisarcik, M., Imae, T., Devinsky, F., Lacko, I., and Bakos, D., J. Colloid
71. Hattori, N., Hirata, H., Okabayashi, H., Furusaka, M., O’Connor, C. J., Interface Sci. 228, 207 (2000).
and Zana, R., Colloid Polym. Sci. 277, 95 (1999). 97. Fuller, S., Shinde, N., Tiddy, G. J., Attard, G. S., and Howell, O., Langmuir
72. Aswal, V. K., De, S., Goyal, P. S., Bhattacharya, S., and Heenan, R. K., 12, 1117 (1996).
Phys. Rev. E 57, 776 (1998). 98. Hattori, N., Hara, M., Okabayashi, H., and O’Connor, C. J., Colloid Polym.
73. De, S., Aswal, V. K., Goyal, P. S., and Bhattacharya, S., J. Phys. Chem. Sci. 277, 306 (1999).
B 102, 6152 (1998). 99. Layn, K. L., Debenedetti, P. G., and Prud’homme, R. K., J. Chem. Phys.
74. Aswal, V. K., De, S., Goyal, P. S., Bhattacharya, S., and Heenan, R. K., 109, 5651 (1998).
Phys. Rev. E 59, 3116 (1999). 100. De, S., Aswal, V. K., Goyal, P. S., and Bhattacharya, S., J. Phys. Chem.
75. Menger, F. M., Keiper, J. S., Mbadugha, B. N. A., Caran, K. L., and B 101, 5639 (1997).
Romsted, L. S., Langmuir 16, 9092 (2000). 101. Schosseler, F., Anthony, O., Beinert, G., and Zana, R., Langmuir 11, 3347
76. Danino, D., Talmon, Y., and Zana, R., Langmuir 11, 1448 (1995). (1995).
77. Israelachvili, J., Mitchell, D. J., and Ninham, B. W., J. Chem. Soc. Faraday 102. Danino, D., Talmon, Y., and Zana, R., J. Colloid Interface Sci. 185, 84
Trans. 72, 1525 (1976). (1997).
78. Zana, R., Lévy, H., Papoutsi, D., and Beinert, G., Langmuir 11, 3694 103. Bernheim-Groswasser, A., Zana, R., and Talmon, Y., J. Phys. Chem.
(1995). B 104, 12192 (2000).
79. Zana, R., In, M., Lévy, H., and Duportail, G., Langmuir 13, 5552 104. Infante, M. R., Pérez, L., and Pinazo, A., in “Novel Surfactants: Prepa-
(1997). ration, Applications and Biodegradability” (K. Holmberg, Ed.), Chap. 3,
80. Bernheim-Groswasser, A., Zana, R., and Talmon, Y., J. Phys. Chem. p. 87. Dekker, New York, 1998.
B 104, 4005 (2000). 105. Imam, T., Devinsky, F., Lacko, I., Mlynarcik, D., and Krasnec, L., Phar-
81. Danino, D., Talmon, Y., Lévy, H., Beinert, G, and Zana, R., Science 269, mazie H.5 38, 308 (1983).
1420 (1995). 106. Kralova, K., and Sersen, F., Tenside Surf. Deterg. 31, 192 (1994).

Vous aimerez peut-être aussi