Vous êtes sur la page 1sur 9

Fluid Phase Equilibria 322–323 (2012) 126–134

Contents lists available at SciVerse ScienceDirect

Fluid Phase Equilibria


journal homepage: www.elsevier.com/locate/fluid

Critical micelle concentration of some surfactants and thermodynamic


parameters of their micellization
Anna Zdziennicka, Katarzyna Szymczyk, Joanna Krawczyk, Bronisław Jańczuk ∗
Department of Interfacial Phenomena, Faculty of Chemistry, Maria Curie-Skłodowska University, Maria Curie-Skłodowska Sq. 3, 20-031 Lublin, Poland

a r t i c l e i n f o a b s t r a c t

Article history: Measurements of density, viscosity, conductivity and light scattering of aqueous solutions of
Received 1 February 2012 sodium dodecylsulfate (SDDS), sodium hexadecylsulfonate (SHS), sodium N-lauryl sarcosinate (SDSa),
Received in revised form 9 March 2012 cetyltrimethylammonium bromide (CTAB), cetylpyridinium bromide (CPyB), dodecyldimethyethylam-
Accepted 12 March 2012
monium bromide (DDEAB), tetradecyltrimethylammonium bromide (TTAB), benzyldimethyldodecylam-
Available online 21 March 2012
monium bromide (BDDAB) and Triton X-100 (TX-100), Triton X-114 (TX-114), Triton X-165 (TX-165)
were carried out at different temperatures. On the basis of the results obtained from these measure-
Keywords:
ments critical micelle concentration, aggregation number, apparent and partial molar volume, standard
Surfactant
Critical micelle concentration
Gibbs energy, enthalpy and entropy were calculated. In the case of ionic surfactants the degree of sur-
Standard Gibbs energy of micellization factant dissociation in micelle was taken into account. There was also determined the standard Gibbs
Enthalpy of micellization energy of micellization using hydrophobic tail-water, hydrophilic head-water interfacial free energy and
Entropy of micellization electrostatic intermolecular interactions. Then the results were compared with those obtained by other
methods and the literature data. The presence of micelles at the concentration of aqueous surfactant solu-
tions determined by the above mentioned methods was confirmed by the light scattering measurements.

© 2012 Elsevier B.V. All rights reserved.

1. Introduction surface tension, density, viscosity, conductivity, light scattering,


etc.) [1,5–21].
Micelle formation is an important phenomenon not only Many factors affect the value of CMC in aqueous media, for
because of a number of interfacial phenomena, such as detergency example, structure of surfactant, presence of the added electrolyte
and solubilization, depending on their existence in the solution but and various organic compounds in the solution as well as presence
because it affects other interfacial phenomena, such as surface or of second liquid phase and the temperature of solution [1,2,4–21].
interface tension reduction that do not directly involve micelles Therefore, many empirical and semi-empirical functions describ-
[1–3]. Recently, micelles have become a subject of great interest for ing the relationship between these factors and CMC are found in
organic chemists and biochemists because of their specific catalysis the literature.
of organic reactions [4]. Therefore, literature reports many studies Among others, Klevens [22] stated the existence of a linear
of the concentration at which the micelle formation takes place dependence between the logarithm of CMC and the number of
(called the critical micelle concentration – CMC), parameters influ- the carbon atoms in the tail of surfactant for homologous straight-
encing on CMC and changes of thermodynamic functions during chain ionic surfactants (soaps, alkane sulfonates, alkyl sulfates,
the micelle formation, which are good tools to explain this process alkyl ammonium chlorides, and alkyltrimethyloammonium bro-
[1–21]. mides), but the constants of this dependence are not associated
The CMC values are usually determined from the abrupt change with the contribution of van der Waals interactions, hydrogen bond
of a physical property in a very small concentration range of sur- and electrostatic interactions in the micellization process.
factants. However, in many cases there are some differences in the According to van Oss and Constanzo [23] the surface free energy
values for the same surfactant depending on the methods using of the surfactant depends on the orientation of its molecules toward
to study the behaviour of bulk solution of surfactants (such as the adherent medium and is different for tail and head of the
surfactant. In the case of tails, surface free energy results only
from the Lifshitz-van der Waals forces, but that of heads from
the Lifshitz-van der Waals and acid–base interactions. Of course,
∗ Corresponding author. Tel.: +48 81 537 56 70; fax: +48 81 533 3348. if the interactions between the surface active ions through the
E-mail address: bronislaw.janczuk@poczta.umcs.lublin.pl (B. Jańczuk). water phase are considered, then according to the extended DLVO

0378-3812/$ – see front matter © 2012 Elsevier B.V. All rights reserved.
doi:10.1016/j.fluid.2012.03.018
A. Zdziennicka et al. / Fluid Phase Equilibria 322–323 (2012) 126–134 127

theory, the electrostatic interactions must be taken into account. Table 1


Chemical sample specifications.
Thus, knowing the surface free energy of the head and tail of sur-
factants as well as electrostatic interactions between the heads of Chemical name Source Initial purity (mass Purification method
surfactants it should be possible to interpret the constants in the fraction)
Klevens equation [22] and to determine the CMC which can be CTAB Sigma–Aldrich 0.98 Crystallization
found in the literature. CPyB Sigma–Aldrich 0.98 Crystallization
Knowing the CMC it is possible to determine thermodynamic SDDS Merc 0.98 Crystallization
SHS Sigma–Aldrich 0.98 Crystallization
parameters of micellization, however, for many ionic surfactants
SDSa Sigma–Aldrich 0.94 Crystallization
type 1:1 electrolyte for the standard Gibbs energy of micelliza- DDEAB Fluka 0.99 None
tion determination it is sometimes unconsciously assumed that the TTAB Sigma–Aldrich 0.98 Crystallization
degree of counterion binding to the surface active agents in the BDDAB Fluka 0.99 None
TX-100 Fluka 0.99 None
micelle is equal to zero or unity [1,15–21,24]. Therefore, literature
TX-114 Fluka 0.99 None
reports considerably different values of the standard Gibbs energy TX-165 Fluka 0.99 None
of micellization of the same surfactants as a result of improper
equation used for its determination [1,15,17,21].
The above mentioned micellization process is also considered 2.2. Measurements
based on the extended DLVO theory [23,25] if the surface free
energy of tail and head of the surfactant as well as the contribu- The density of aqueous solution of surfactants was measured
tion of electrostatic forces to the surfactant ion interactions through with a U-tube densitometer (DMA 5000 Anton Paar) at the constant
the water phase are known. In the literature there are only a few temperatures 293, 303 and 313 K.
papers dealing with the correlation between the surface free energy The precision of the density and temperature measurements
of surfactants resulting from different intermolecular forces and given by the manufacturer is ±0.000001 g/m3 and ±0.001 K. Uncer-
their Gibbs energy of micellization [23,25,26]. tainty was calculated to be equal to 0.01%. The densitometer was
The purpose of our paper was to explain physicochemical signif- calibrated regularly with distilled and deionized water.
icance of constants in the Klevens equation and predict the Gibbs All viscosity measurements of the aqueous solution of stud-
energy of micellization on the basis of the interfacial free energy ied surfactants were performed with the Anton Paar viscosimeter
of tail-water and head-water as well as the electrostatic interac- (AMVn) at 293, 303 and 313 K ± 0.01 K with the precision of
tions between the heads of surfactants as well as to compare the 0.0001 mPa s and uncertainty 0.3%.
obtained results to those determined from CMC. As one can find in The conductivity measurements of ionic surfactant solutions
literature different values of CMC and the aggregation number for were made by the conductometer, Mettler Toledo, joined with
a given surfactant it seemed more proper to determine them from the thermostat LAUDA RE 415S with the temperature precision
the measurements of different physicochemical properties of the ±0.01 K. The uncertainty of the conductivity measurements was
aqueous solution of surfactants under the same conditions. ±0.5%. All density, viscosity and conductivity measurements were
For this purpose the measurements of density, viscosity, made for 3 samples of four sets. Density, viscosity and conductivity
conductivity and light scattering of aqueous solutions of sodium for SHS were additionally measured at 323 K.
dodecylsulfate (SDDS), sodium hexadecylsulfonate (SHS), sodium The size of the surfactants aggregates was determined using
N-lauryl sarcosinate (SDSa), cetyltrimethylammonium bromide Zetasizer Nano (Malvern, UK) with the uncertainty equal to ±2%.
(CTAB), cetylpyridinium bromide (CPyB), dodecyldimethyethy-
lammonium bromide (DDEAB), tetradecyltrimethylammonium 3. Results and discussion
bromide (TTAB), benzyldimethyldodecylammonium bromide
(BDDAB) and Triton X-100 (TX-100), Triton X-114 (TX-114) and 3.1. Critical micelle concentration
Triton X-165 (TX-165) were performed at different temperatures.
In the investigations the literature data of surface tension of The determined values of CMC for each surfactant studied are
aqueous solutions and surface free energy of surfactants were also somewhat different depending on the property of solutions which
taken into account [27]. were taken into account (Table 2). However, it is difficult to find
the relationship between the method of CMC determination and its
value. The range of CMC values for a given surfactant determined
2. Experimental by using the surface tension [27], density (Fig. 1, Figs. S1 and S2),
viscosity (Fig. 2, Figs. S3 and S4) and conductivity data (Fig. 3,
2.1. Materials Figs. S5 and S6) is lower than the discrepancies of CMC values which
can be found in the literature [1,5,6,8,13,18,20,22–26,28–50]. It
Double distilled and deionized water (Destamat Bi18E), (p- is interesting that the light scattering data indicate that the
(1,1,3,3-tetramethylbutyl)phenoxypolyoxyethylene glycols, Tri- micelles are formed for each surfactant at the concentration close
ton’s: X-100 (TX-100), X-114 (TX-114) and X-165 (TX-165), to that determined from the surface tension isotherms of solu-
dodecyldimethylethylammonium bromide (DDEAB) and ben- tions [27]. Taking into account the structure of surfactants one
zyldimethyldodecylammonium bromide (BDDAB) (purity greater can state that for Triton’s the CMC increases with the increase
than 99%), were bought from FLUKA and used without fur- of the oxyethylene groups in their molecule, as expected. How-
ther purification. Cetyltrimethylammonium bromide (CTAB) ever, this increase is not proportional to the number of groups. The
(SIGMA–ALDRICH) (purity > 98%), tetradecyltrimethylammonium CMC for the surfactants decreases with the increase of tempera-
bromide (TTAB) (SIGMA–ALDRICH) (purity > 98%), cetylpyridinium ture in the range from 293 to 313 K which is probably connected
bromide (CPyB) (SIGMA–ALDRICH) (purity > 98%), sodium dode- with the dehydration of oxyethylene groups. It is known that each
cylsulfate (SDDS) (Merck) (purity > 98%) and sodium hexade- oxyethylene group can be joined with two molecules of water
cylsulfonate (SHS) (SIGMA–ALDRICH) (purity > 98%) and sodium [51]. The amount of water affects the Triton’s CMC more sig-
N-lauryl sarcosinate (SDSa) (SIGMA–ALDRICH) (purity > 94%) nificantly than the number of oxyethylene groups which can be
(Table 1) were purified by the method described in the literature seen from the comparison of CMC values determined at 293 and
[28]. 313 K for TX-114, TX-100 and TX-165. At 293 K these differences
128 A. Zdziennicka et al. / Fluid Phase Equilibria 322–323 (2012) 126–134

Table 2
The values of the critical micelle concentration (CMC) of surfactant determined from the surface tension (LV ), density (), viscosity () and conductivity () measurements
at 293 (1), 303 (2) and 313 K (3), respectively, calculated from Eqs. (4) and (10a)–(10g) at 293 K and taken from literature.

Substrate CMC [mol/dm3 ] CMC [mol/dm3 ] CMC [mol/dm3 ] CMC [mol/dm3 ] CMC [mol/dm3 ] CMC [mol/dm3 ] References
(LV )a ()b ()c ()d theoret. literature data

1 9.15 × 10−4 7.65 × 10(4 9.33 × 10(4 9.77 × 10(4


CTAB 2 1.14 × 10(3 8.58 × 10(4 1.01 × 10(3 1.11 × 10(3 9.0 × 10(4 7.0 × 10(4 –1.0 × 10(3 [1,5,6,8,29,30]
3 1.16 × 10(3 9.92 × 10(4 1.07 × 10(3 1.22 × 10(3

1 4.48 × 10(4 3.91 × 10(4 3.43 × 10(4 4.62 × 10(4


CPyB 2 6.19 × 10(4 5.24 × 10(4 4.73 × 10(4 6.41 × 10(4 5.23 × 10(4 6.40 × 10(4 –7.0 × 10(4 [13,31–33]
3 7.46 × 10(4 6.32 × 10(4 5.63 × 10(4 7.35 × 10(4

1 8.20 × 10(3 6.92 × 10(3 8.10 × 10(3 8.14 × 10(3


SDDS 2 8.87 × 10(3 7.28 × 10(3 8.72 × 10(3 9.12 × 10(3 7.69 × 10(3 5.40 × 10(3 –8.40 × 10(3 [1,6,8,30,34–36]
3 9.45 × 10(3 7.62 × 10(3 9.08 × 10(3 9.26 × 10(3

SHS 4 7.12 × 10(4 6.99 × 10(4 7.18 × 10(4 6.76 × 10(4 7.75 × 10(4 7.0 × 10(4 [1,22]

1 3.65 × 10(3 2.40 × 10(3 2.35 × 10(3 3.15 × 10(3


SDSa 2 4.39 × 10(3 2.50 × 10(3 2.13 × 10(3 2.77 × 10(3
3 4.37 × 10(3 2.55 × 10(3 1.76 × 10(3 2.67 × 10(3

1 9.90 × 10(3 1.38 × 10(2 1.00 × 10(2 1.01 × 10(2


DDEAB 2 1.14 × 10(3 1.53 × 10(2 1.11 × 10(2 1.19 × 10(2 9.95 × 10(3 1.40 × 10(2 –1.52 × 10(2 [38,39]
3 1.29 × 10(2 1.71 × 10(2 1.27 × 10(3 1.39 × 10(2

1 3.53 × 10(2 1.69 × 10(3 3.41 × 10(3 3.30 × 10(3


TTAB 2 6.86 × 10(3 3.30 × 10(3 6.21 × 10(3 4.05 × 10(3 3.37 × 10(3 3.60 × 10(3 –3.75 × 10(3 [6,29,40,41]
3 7.79 × 10(3 3.97 × 10(3 7.04 × 10(3 4.31 × 10(3

1 5.25 × 10(3 6.25 × 10(3 3.46 × 10(3 5.55 × 10(3


BDDAB 2 6.71 × 10(3 7.80 × 10(3 4.15 × 10(3 6.26 × 10(3 6.10 × 10(3 5.20 × 10(3 –5.60 × 10(3 [42,43]
3 6.76 × 10(3 8.00 × 10(2 4.72 × 10(3 7.08 × 10(3

1 2.90 × 10(4 2.51 × 10(4 2.58 × 10(4


TX-100 2 2.69 × 10(4 2.52 × 10(4 2.44 × 10(4 2.78 × 10(4 2.20 × 10(4 –3.0 × 10(4 [44–47]
3 2.50 × 10(4 2.25 × 10(4 2.32 × 10(4

1 5.41 × 10(4 5.00 × 10(4 5.00 × 10(4


TX-165 2 4.87 × 10(4 4.55 × 10(4 4.56 × 10(4 5.30 × 10(4 4.50 × 10(4 –5.59 × 10(4 [44,48]
3 4.41 × 10(4 2.84 × 10(4 2.85 × 10(4

1 1.68 × 10(4 1.60 × 10(4 2.00 × 10(4


TX-114 2 1.53 × 10(4 1.46 × 10(4 1.81 × 10(4 3.51 × 10(4 2.10 × 10(4 –2.80 × 10(4 [49,50]
3 1.41 × 10(4 1.35 × 10(4 1.65 × 10(4

4 – at 323 K.
a
Uncertainties of CMC determination are from 0.8 to 2% depending whether the CMC was calculated from the surface tension (LV ) measurements.
b
Uncertainties of CMC determination are from 0.8 to 2% depending whether the CMC was calculated from the density () measurements.
c
Uncertainties of CMC determination are from 0.8 to 2% depending whether the CMC was calculated from the viscosity () measurements.
d
Uncertainties of CMC determination are from 0.8 to 2% depending whether the CMC was calculated from the conductivity () measurements.

between the CMC values for Triton’s are higher than those at with a larger number of water molecules oriented in a specific
313 K. way at the hydrophilic head including a longer alkyl group. The
It means that the influence of one hydrated oxyethylene group change of hydrogen in the methyl group into the aryl one causes
on CMC is larger than that of the dehydrated one. There is no cor- the decrease of CMC (Table 2). It is known that the interactions
relation between the CMC of Triton’s and the efficiency of their between benzene and alkane molecules through the water are
adsorption at the water–air interface, which is almost the same for equal to the double water-benzene and water-alkane interfacial
all surfactants of this type [27,52–54]. This probably results from tension (numerically equal to the interfacial free energy) [55].
the fact that during the adsorption process changes of oxyethylene The work of benzene molecules interactions through the water
group dehydration at a given temperature do not occur and Gibbs phase is equal to about 67 mJ/m2 and that of alkane is close to
energy of adsorption results only from the transfer of hydrophobic 102 mJ/m2 [54]. On the other hand, the contactable area of benzene
group from the solution phase to the air one. As all Triton’s have the molecule is equal to 25 Å2 and CH3 group close to 10.48 Å2 . Thus,
same hydrophobic group, the Gibbs energy of adsorption should be the so-called hydrophobic interactions between the aryl groups
the same as reported in the literature [52–54]. through the water phase are about 1.6 times larger than those of
Comparing the CMC values for the studied cationic surfactants the methyl group. It should be stressed that for the cationic sur-
it can be stated that the repulsive interactions of the pyridinium factants the discrepancies of CMC values at a given temperature
hydrophilic group are lower than those of trimethylammonium obtained from the surface tension, density, viscosity and conduc-
because the CMC value at each studied temperature is lower for tivity data are larger than those for nonionic Triton’s. This may
this surfactant than that of CTAB (Table 2). However, both these be associated with a wider range of concentration in which the
hydrophilic groups are hydrated to a lower degree than that of micellization process occurs than in the case of nonionic surfactants
oxyethylene group because for the CTAB and CPyB CMC increases [12].
with the increasing temperature. The change of the methyl group The changes of the CMC values of surfactants including oxygen
in the head joined with nitrogen in the ammonium group into in their hydrophilic group as a function of temperature are lower
the ethyl one causes the minimal influence on CMC of the surfac- than those having pyridinium and ammonium groups. Probably the
tants at the temperature from 293 to 313 K. This may be associated greater increase of CMC for the studied cationic surfactants in the
A. Zdziennicka et al. / Fluid Phase Equilibria 322–323 (2012) 126–134 129

Fig. 1. A plot of the density () of aqueous solutions of SDDS (curve 1), SHS (curve
2), SDSa (curve 3), CTAB (curve 4), CPyB (curve 5), DDEAB (curve 6), TTAB (curve 7), Fig. 3. A plot of the conductivity () of aqueous solutions of SDDS (curve 1), SHS
BDDAB (curve 8), TX-100 (curve 9), TX-165 (curve 10) and TX-114 (curve 11) vs. the (curve 2 – at 323 K), SDSa (curve 3), CTAB (curve 4), CPyB (curve 5), DDEAB (curve
surfactant concentration, C, at 293 K, and at 323 K for SHS (curve 2 ). 6), TTAB (curve 7), BDDAB (curve 8) vs. the surfactant concentration, C, at 293 K.

range of temperature from 293 to 313 K is caused by the higher attain the size which can be detected by the Zetasizer Nano. On
degree of their hydration. the other hand, it is known that the CMC is not the one strictly
defined value but it is a range of concentrations at which aggre-
3.2. Aggregation number and degree of counterion dissociation gates are formed. Because of that in the case of SDSa aggregates
are noticeable from the light scattering measurements with the
The light scattering studies indicate that all ionic surfac- concentration 10−2 mol/dm3 , but its CMC determined from the
tants taken into account, except for SHS, form spherical micelles density (Fig. 1), viscosity (Fig. 2), conductivity (Fig. 3) and sur-
(TableS1). For SDDS and SDSa there is excellent agreement between face tension measurements [27] is in the range from 2.35 × 10−3
the experimental values of the aggregation number (Nagg ) obtained to 3.65 × 10−3 mol/dm3 (Table 2). For this surfactant it is difficult
from the light scattering measurements and those calculated from to find reasonable values of its aggregation number in the literature
simple geometrical considerations and for the SDDS the Nagg val- .
ues are in the range of the aggregation number presented in the The aggregation number for TTAB and DDEAB obtained from the
literature [1,56,57]. In the case of SHS according to the literature light scattering measurements is the same as that determined by
data CMC at 323 K is equal to 7 × 10−4 mol/dm3 [1]. This value Kwon and Lii [58], Junquera and Aicart [59] and Lianos and Zana [60]
was confirmed by the density (Fig. 1, curve 2 ), viscosity (Fig. 2, but in the case of BDDAB this value is somewhat higher than that
curve 2 ), conductivity (Fig. 3, curve 2) and surface tension measure- in the literature [61]. For DDEAB and BDDAB the values of the Nagg
ments; however, the Zetasizer Nano does not indicate the presence are somewhat lower than those determined from the geometrical
of the micelle at this temperature. It is possible that under the considerations. In the case of CTAB and CPyB there is no correlation
mentioned conditions the aggregates start to form but they do not between the experimental, theoretical and literature values of Nagg
[62–64].
Formation of the micelles by the ionic surfactants is connected
with the change of the dissociation degree (˛). It is assumed that in
the monomer state the ionic surfactants are practically completely
dissociated, but in the aggregated form the degree of dissociation
decreases and depends on the temperature (Table S2).
At 293 K SDDS and SDSa have lower degree of dissociation than
the cationic surfactants, but the increase of temperature causes
greater increase of dissociation degree than that for the cationic
surfactants. As mentioned above, it should be effected by stronger
dehydration of anionic than cationic surfactants as a function of
temperature. However, it is difficult to find directly the correla-
tion between the degree of counterions dissociation and the CMC
value. The lowest degree of counterion dissociation is observed
for SDSa and the largest for CTAB. For all ionic surfactants studied
the increase of the degree of counterion dissociation is observed.
It is difficult to compare the obtained values to the literature
data because they were determined at 298 K [6,28,29,38–40,61,65].
However, it can be stated that these values are somewhat different.
It is interesting to note that the Nagg determined from the light
scattering data indicates that the radius of the spherical micelles
Fig. 2. A plot of the viscosity () of aqueous solutions of SDDS (curve 1), SHS (curve
is proportional rather to the total length of surfactant than to that
2), SDSa (curve 3), CTAB (curve 4), CPyB (curve 5), DDEAB (curve 6), TTAB (curve 7),
BDDAB (curve 8), TX-100 (curve 9), TX-165 (curve 10) and TX-114 (curve 11) vs. the of the hydrophobic group and the size of hydrophilic heads plays
surfactant concentration, C, at 293 K, and at 323 K for SHS (curve 2 ). rather a smaller role than repulsive forces between them.
130 A. Zdziennicka et al. / Fluid Phase Equilibria 322–323 (2012) 126–134

In the case of Triton’s the lamellar micelles are formed


[44,45,66,67], however, if we assume that spherical micelles can
be formed in the same range of concentrations, then the Nagg val-
ues obtained from the light scattering measurements are close to
those determined by the other investigators [44,45,47,66,67]. Our
assumption is confirmed by simple geometrical calculations and
that, at the first approximation, the hydrophilic groups can form a
random coil [68]. In such a case it is possible to assume that the radii
of sphere for free Triton’s differ only slightly. Based on the Robson
and Dennis data [68] and the length of bonds between the atoms in
the molecules, the radius of the spheres was calculated as equal to
2.828 nm (Table S1). Taking into account this radius and apparent
or partial molar volumes (Fig. 4a and 4b) at the same concentrations
as in the case of the light scattering measurements the calculated
values of aggregation numbers are close to those obtained from the
experimental data.

3.3. Apparent and partial molar volume

Formation of aggregates of surfactants in aqueous solutions


should be reflected by the changes of their density (Fig. 1,
Figs. S1 and S2) and viscosity (Fig. 2, Figs. S3 and S4). As follows
from Figs. 1 and 2 the slope of the relationships between the den-
sity or viscosity and the concentration of surfactants before and
after CMC is different, and therefore it is possible to determine the
CMC values for a given surfactant.
The influence of surfactants concentration on the structure of
the solution can be more visible from the apparent (v ) and partial
(V̄M ) molar volumes.
The apparent molar volume was determined from the following
equation [69,70]:
MS 1000(0 − )
V = + (1)
0 C
where MS is the molecular weight of the surface active agent, C is
the concentration of the surface active agent in mol/dm3 and 0 is
Fig. 4. (a) A plot of the apparent and partial molar volumes of CTAB (curves 1 and
the density of “pure” solvent.
1 ), CPyB (curves 2 and 2 ), DDEAB (curves 3 and 3 ), TTAB (curves 4 and 4 ), BDDAB
The partial molar volume V̄M was calculated from the equation (curves 5 and 5 ) vs. the surfactant concentration, C, at 293 K, as the example. (b)
[71]: A plot of the apparent and partial molar volumes of SDDS (curves 1 and 1 ), SHS
  (curves 2 and 2 ), SDSa (curves 3and 3 ), TX-100 (curves 4 and 4 ), TX-165 (curves 5
MS (100 − Cp ) d and 5 ) and TX-114 (curves 6 and 6 ) vs. of the surfactant concentration, C, at 293 K,
V̄M = 1− (2) as the example.
  dCp

It appeared that the  data fit a polynomial of Cp (the mass percent-


agreement between the apparent and partial molar volumes of
age concentration) given by:
these surfactants [39,72].
 = A + BCp + DCp2 (3)
3.4. Gibbs energy of interactions and CMC
where A, B and D are the constants.
The calculated values of the molar partial volumes for all stud- According to the extended DLVO theory [23,25,26] the values of
ied surfactants and apparent molar volume for CTAB and SDDS the CMC depend mainly on the three kinds of intermolecular inter-
increase as a function of surfactants concentration and no char- actions between the surfactant molecules or ions through the water
acteristic points corresponding to the CMC are observed (Fig. 4a phase. Gibbs energy (G) of these interactions between the ions of
and b). surfactants of 1:1 type electrolytes results from the Lifshitz-van der
For all surfactants except for SDDS and CTAB the drop of Waals, Lewis acid–base and electrostatic forces and is connected
apparent molar volume is observed in the range of surfactant con- with the CMC based on the relation:
centration lower than that of CMC and corresponds to almost linear
dependence of the surface tension as a function of surfactant con- G = 2kT ln CMC (4)
centration [27]. This probably indicates that dimmers or trimmers
are formed. This drop of apparent molar volume indicates that the where k is the Boltzman constant and T is the absolute temperature.
density of the micelles is higher than that of the surfactants in On the other hand, the total Gibbs energy of interactions
the monomeric form. The light scattering measurements do not between the surfactant molecules through the water phase fulfills
show clearly whether the size of micelles grows as a function of the equation [23,25,26]:
surfactant concentration. Unfortunately, in most cases the changes GTot = G1W
LW AB EL
1 + G1W 1 + G1W 1 (5)
of the partial and apparent molar volumes of the studied surfac-
tants are somewhat different from those in the literature. However, where G1WLW , GAB and GEL are the Gibbs energy of the inter-
1 1W 1 1W 1
it should be stressed that the literature does not report the actions resulting from the Lifshitz-van der Waals, Lewis acid–base
A. Zdziennicka et al. / Fluid Phase Equilibria 322–323 (2012) 126–134 131

and electrostatic forces, respectively. Subscripts 1 and W refer to dependence between the CMC and the number of the carbon atoms
the surface active ions of the surfactant and water, respectively. in the hydrophobic tails of surfactants. It has the following form:
It is known that the surfactants oriented by the tails (T) and
heads (H) toward the air phase have a different surface free energy log CMC = a − bn (11)
[23,25,26]. Taking this fact into account Eq. (5) can be expressed in
where a is the constant for particular ionic head at a given tem-
the form:
perature and b is the constant independent of the length of the
GTot = GTWT
LW AB
+ GTWT LW
+ GHWH AB
+ GHWH EL
+ GHWH (6) straight-chain hydrophobic part of surfactant, joined with it.
For soaps, alkane sulfonates, alkyl sulfates, alkylammonium
LW + GAB = −2 AB AB
Because GTWT TWT TW and GHWH + GHWH = chlorides, alkyltrimethylammonium bromides the experimental b
−2HW we obtain: constant is equal to log2 [1,22]. From Eqs. (4), (9) and (10a)–(10g)
it is possible to obtain the equation:
GTot = −2(TW + HW ) + GHWH
EL
(7)
 EL

The contribution to the Gibbs energy of the electrostatic forces 5.34 + [(−2WH + GHWH )] × SC (H)
log CMC =
between two planes, including the surfactant ions, separated by a 4.606kT
distance d and with similar double layers can be expressed as [73]:
5.34
ε × 10−21 − × 10−21 n (12)
EL
GHWH = 2
+ 1 − coth d) 4.606kT
4
0 (cosech d (8)

where ε is the dielectric constant of water, 0 is the surface poten- The comparison of Eq. (11) with (12) indicates that the constants
tial for the surfactant ions and is the reciprocal of Debye length. a and b fulfill the conditions:
To compare the Gibbs energy of surfactant ions or molecules  EL

5.34 + [(−2WH + GHWH )] × SC (H)
interactions through the water phase calculated from Eq. (7) with a= × 10−21 (13)
that from Eq. (4) the contactable area of the surfactant tail, SC (T) 4.606kT
and head, SC (H) must be known. Then it follows from Eq. (7) that
5.34
[26]: b= × 10−21 (14)
4.606kT
EL
G = −2TW × SC (T ) − (2HW − GHWH ) × SC (H) (9) As follows from Eq. (13) the constant a depends on the temperature
It is commonly assumed that one or two CH2 groups are and the kind of the head as well as assumed different values for each
included in the surfactant head [1]. Therefore the contactable area homologous series of surfactants, while the constant b depends
of the tails of straight chain surfactants is the sum of that of CH3 only on the temperature (Eq. (14)). It appears that the constant b
and n-3 of CH2 groups, where n is the number of the carbon at 293 K is equal to 0.2867 and is close to that proposed by Klevens
atoms in the surfactant molecule. [1,22] for the homologous series of ionic surfactants type 1:1 elec-
From the study of strong adsorption of long-chain n-paraffins trolyte and only for some surfactants to those given by Huibers
on graphitized carbons, Groszek [74] proposed that the area per [75].
CH2 group should be that of one hexagon equal to 5.24 Å2 . At The calculations of the interactions between the surfactants
the first approximation, it can be assumed that the area of CH3 is molecules through the water phase indicate that on the basis of
equal to two hexagons. the surface free energy of the “tail” and “head” of surfactants and
Taking into account the geometrical size of heads of surfactants, electrostatic interactions between the charged heads it is possible
water-tail and water-head interfacial tension (numerically equal to predict the tendency of surfactants to aggregate in the aqueous
to the interfacial free energy) as well as Gibbs energy resulting solutions.
from the electrostatic interactions, it is possible to calculate G. In
such case for the ionic surfactants there are fulfilled the following 3.5. Thermodynamic parameters of micellization
equations:
Two general approaches have been employed to tackle the prob-
G = [(n − 1) × −5.34 + 19.38] × 10−21 J/mol (SDDS) (10a)
lem of micelle formation [21]. The first called the phase separation
G = [(n − 1) × −5.34 + 22.18] × 10−21 J/mol (SHS) (10b) model is the simplest one and treats micelles as a simple phase. In
this model, micelle formation is considered as a phase separation
−21
G = [(n − 1) × −5.34 + 23.39] × 10 J/mol (CTAB, TTAB) (10c) phenomenon. In the second called the mass action model, micelles
−21 and single surfactant molecules or ions are considered to be in the
G = [(n − 1) × −5.34 + 21.46] × 10 J/mol (DDEAB) (10d)
association-dissociation equilibrium.
G = [(n − 1) × −5.34 + 19] × 10−21 J/mol (CPyB) (10e) In the equilibrium state of solution at the concentration of the
surfactant equal to the CMC the chemical potential of the surfactant
G = [(n − 1) × −5.34 + 18.7] × 10−21 J/mol (SDSa) (10f)
in the monomeric form fulfills the equation [2]:
−21
G = [(n − 1) × −5.34 + 22.84] × 10 J/mol (BDDAB) (10g)
1 = 1 + RT ln a1 (15)
It should be stressed that the Gibbs energy of interactions does
The chemical potential of the surfactant in the micelle can be
not correspond to that of the head and tail of one molecule and is
expressed in the form:
only calculated on the contactable area of this part of surfactant.
Assuming that the values of G calculated from Eqs. (10a)–(10g) m = 0m + RT ln am (16)
(Table S3) are close to those obtained from Eq. (4); it is possible to
determine the CMC of the studied surfactants. It appears that the where 1 and m are the chemical potentials of the surfactant in
CMC values calculated in this way are close to those obtained exper- the monomeric and aggregated forms, respectively, and 1 and 0m
imentally from the surface tension [27], density (Fig. 1), viscosity are their standard chemical potentials at a given temperature, a1
(Fig. 2) and conductivity (Fig. 3) measurements at 293 K (Table 1). and am refer to the activity of the surfactant in the bulk phase and
For homologous straight-chain ionic surfactants in aqueous micelle, however, defined in different ways, namely asymmetri-
solutions Klevens [1,22] proposed a simple equation to express the cally and symmetrically.
132 A. Zdziennicka et al. / Fluid Phase Equilibria 322–323 (2012) 126–134

It means that a1 = xCMC f1∗ = xCMC when xCMC → 0 and am = The obtained values of Gmic 0 of ionic surfactants are lower
xm fm = xm = 1, where xCMC and xm are the surfactant molar frac- than those for nonionic Triton’s (Table S2), which can probably
tions at CMC and in the micelle, and f1∗ and fm are the activity result from higher hydrophobic interactions between the tails of
coefficients of the surfactant in the bulk phase and the micelle, the studied ionic surfactants which causes strong water molecules
respectively. 0
reorientation. This process causes probably larger increase of Smic
According to the asymmetric and symmetric definition of activ- than in the case of nonionic surfactants, because their contactable
ity, f1∗ → 1 if xCMC → 0 and fm → 1 if xm → 1. area with water molecules is lower than that of the studied ionic
Because in the equilibrium state 1 = m and at the first surfactants. For these surfactants the process of the counterions
approximation am = 1 the standard Gibbs energy of micellization binding with the surface active ions in the micellar state contributes
per mole of monomer (Gmic 0 ) can be expressed by the equation
to the standard Gibbs energy of aggregation. It is confirmed by
[21]: Hmic0 values which are positive for the anionic surfactants and
0 negative for the cationic ones. However, the positive values are
Gmic = 0m − 1 = RT ln a1 ≈ RT ln xCMC (17)
close to those for Triton’s (Table S2). Negative enthalpy for the
It seems that the assumption that a1 = xCMC f1∗ = xCMC is reason- cationic surfactants suggests that their heads may be hydrated
able because for most surfactants the CMC corresponds to a very to a smaller degree than the anionic ones and energy needed for
small value of xCMC [1,21]. For this reason it can be assumed that dehydration is lower. Of course, the thermodynamic parameters
xCMC = CMC/ω where ω is the number of water moles in 1 dm3 at obtained on the basis of CMC determined by different methods are
a given temperature and the CMC is expressed in mol/dm3 . Taking not the same.
this into account Eq. (17) assumes the form: It seems that each method of CMC determination is sensitive
to different sizes of aggregates. Comparing the thermodynamic
0 CMC
Gmic = RT ln (18) parameters determined by us with those in the literature
ω
[1,5,9,18–20,29,38,39,42,61,65,77–79] three cases should be taken
It should be stated that Gmic 0 ought to be calculated using 0
into account. First, it is possible to find out that Gmic was cal-
CMC/ω but not only CMC as can be found in many cases in the culated by using CMC instead of xCMC . At this point it should be
literature [1,15,16,21]. stressed that in the literature the Gibbs energy of micellization
Eq. (18) seems to be reasonable for nonionic surfactants but for is called free energy of micellization [1,21,61]. Of course, these
the ionic ones the equation derived by using the mass action model two thermodynamic functions are equal because the micelliza-
is more adequate. tion process is considered as an isothermic–izobaric–izochoric one
0
For the ionic surfactants a convenient solution for relating Gmic [1,21]. In other cases, these functions have different values. The
to CMC was given by Philips [76] who arrived at the following second case deals with the application of the same equations for
expression: the ionic and nonionic surfactants and in the third the degree
 p
 CMC of dissociation at the micelle–water interface is not taken into
0
Gmic = 2− RT ln (19) account in the calculations of Gibbs energy of micellization and
n∗ ω
the degree of surfactant dissociation is exchangeably used with the
where n* is the number of surfactant ions forming a micelle and p degree of bonding of counterions to the surface active ions. There-
is the number of counterions bonded to the micelle. fore, it is very difficult to compare the thermodynamic parameters
It means that 1 − (p/n∗ ) = ˛ (˛ is the degree of dissociation). It of surfactant studies obtained by us to those from the literature
is clear if (p/n∗ ) → 1 (the counterions are bounded to micelles), Eq. [1,5,9,18–22,29,38,39,61,65,77–79]. However, in some cases the
(19) assumes the form of Eq. (18), but if p → 0 i.e., when counterions values of Gmic0 , H 0 and S 0 calculated by us are close to those
mic mic
are not bounded to the micelle, it assumes the following form: in the literature [29,38,39,65,80].
CMC As the CMC obtained on the basis of Gibbs energy of molecules
0
Gmic = 2RT ln (20) or ions of surfactant through the water phase interactions are close
ω
to those determined experimentally, it means that from the macro-
Since scopic properties of surfactants such as surface free energy of the
0
Gmic 0
= Hmic 0
− TSmic (21) hydrophobic and hydrophilic groups and electrostatic interactions
it is possible to predict the micellization process.
and
0 )
d(Gmic 0
= −Smic (22)
dT
4. Conclusion
0
if Hmic (enthalpy of micellization) is constant over the tempera-
ture range investigated. The CMC values of CTAB, CPyB, SDDS, SHS, TTAB, DDEAB,
Alternatively, SDSa, BDDAB, TX-114, TX-100 and TX-165 determined from the
0 /T ) isotherms of surface tension, density, viscosity and conductivity
d(Gmic
T2 0
= −Hmic (23) are somewhat different for a given surfactant. It confirms the sug-
dT
gestion that CMC is not one pinpoint value but it is a range of
0 (entropy of micellization) is constant over the temperature
if Smic concentrations at which aggregates are formed. However, not for
range investigated. all studied surfactants the minimal values of their CMC were con-
As mentioned before there are some differences in CMC for firmed by light scattering measurements. In the case of the ionic
the same surfactant determined from the surface tension, den- surfactants the ranges of their concentration at which the aggre-
sity, viscosity and conductivity measurements (Table 2), therefore gates can be formed include the values predicted by the Klevens
to calculate Gmic0 of surfactants the CMC values obtained by all equation. It appears that the empirical constants in this equation
methods and the degree of dissociation obtained from the conduc- can be deduced on the basis of surface free energy of tail and head
tivity measurements (Table S2) were taken into account. However, of surfactants, and electrostatic interactions between heads of sur-
in order to determine Hmic0 and S 0 only the data obtained from factants in aqueous media if the contactable area of tail and head
mic
the surface tension measurements were considered. is possible to determine.
A. Zdziennicka et al. / Fluid Phase Equilibria 322–323 (2012) 126–134 133

Contrary to some investigators, for determination of standard EL


G1W the Gibbs energy of the interactions between surface
1
Gibbs energy of micelle formation, the molar fraction of surfac- active ions through the water resulting from the electro-
tant at CMC instead of the CMC expressed in mol/dm3 should be static forces (J), Eq. (5)
taken into account. Also for ionic surfactants the degree of dissoci- EL
GHWH the Gibbs energy of the interactions between surfactant
ation of surface active ions at the micelle-solution interface, which heads through the water resulting from the electrostatic
depends on the type of surface active ions and counterions, should forces (J), Eq. (6)
be considered. Moreover, the temperature cannot be neglected. LW
G1W the Gibbs energy of the interactions between surface
1
As the values of standard enthalpy and entropy of micellization active ions through the water resulting from the Lifshitz-
depend on the way of standard Gibbs energy determination, there- van der Waals forces (J), Eq. (5)
fore in many cases, our thermodynamic parameters are somewhat LW
GHWH the Gibbs energy of the interactions between the sur-
different from those reported in the literature. factant heads through the water resulting from the
The standard Gibbs energy of ionic surfactants under investiga- Lifshitz-van der Waals forces (J), Eq. (6)
tions can be predicted from Gibbs energy of interactions between LW
GTWT the Gibbs energy of the interactions between the sur-
the surfactants ions through the water phase including those of the factant tails through the water resulting from the
Lifshitz-van der Waals, acid–base and electrostatic forces knowing Lifshitz-van der Waals forces (J) Eq. (6)
the contactable area of tail and head of surfactant. 0
Hmic the standard enthalpy of micellization (kJ/mol), Eq. (21)
0
Smic the standard entropy of micellization (J/mol K), Eq. (21)
List of symbols ˛ dissociation degree
a the constant for particular ionic head at a given tempera- HW the surfactant head-water interfacial free energy (mJ/m2 ),
ture, Eq. (11) Eq. (7)
a1 the activity of the surfactant in the bulk phase, Eq. (15) TW the surfactant tail-water interfacial free energy (mJ/m2 ),
am the activity of the surfactant in the micelle, Eq. (16) Eq. (7)
A the constant, Eq. (3) ε the dielectric constant of water, Eq. (8)
b the constant independent of the length of the straight-  the viscosity (mPa s)
chain hydrophobic part of surfactant, Eq. (14) the reciprocal of Debye length, Eq. (8)
B the constant, Eq. (3)  the conductivity (␮S/cm)
C the molar concentration (mol/dm3 ), Eq. (1) 1 the chemical potential of the surfactant in the monomeric
Cp the mass percentage concentration, Eq. (2) form, Eq. (15)
D the constant, Eq. (3) m the chemical potential of the surfactant in the aggregated
d the distance between the surfactant ions in water (nm), form, Eq. (16)
Eq. (8) 1 the standard chemical potential of surfactant in the
f1∗ the activity coefficient of the surfactant in the bulk phase monomeric form at a given temperature, Eq. (15)
fm the activity coefficient of the surfactant in the micelle 0m the standard chemical potential of surfactant in the aggre-
k the Boltzman constant (J/mol K), Eq. (4) gated form at a given temperature, Eq. (16)
l the molecule length (nm)  the density of the surfactant solution (g/cm3 ), Eq. (1)
MS the molecular weight of the surface active agent (g), Eq. o the density of “pure” solvent (g/cm3 ), Eq. (1)
(1) v the apparent molar volume, Eq. (1)
n the number of carbon atoms in the surfactant molecules, 0 the surface potential for the surfactant ions, Eq. (8)
Eq. (10a) ω the number of water moles in 1 dm3 , Eq. (18)
n* the number of surfactant ions in the micelle, Eq. (19)
Nagg the aggregation number
Acknowledgment
p the number of surfactant ion in micelle bonding with
counterion, Eq. (19)
The financial support from Polish Ministry of Science and Higher
R the gas constant (J/mol K), Eq. (15)
Education, Project No. N N204 352040 is gratefully acknowledged
SC (H) the contactable area of the surfactant head (nm2 ), Eq. (9)
SC (T ) the contactable area of the surfactant tail (nm2 ), Eq. (9)
T absolute temperature (K), Eq. (4) Appendix A. Supplementary data
V̄M the partial molar volume, Eq. (2)
xCMC the surfactant molar fraction at CMC Supplementary data associated with this article can be found, in
xm the surfactant molar fraction in the micelle the online version, at doi:10.1016/j.fluid.2012.03.018.
G the Gibbs energy of interactions between the ions of sur-
factant of 1:1 type electrolytes (J), Eq. (4) References
GTot the total Gibbs energy of interactions between the sur-
factant molecules through the water phase (kJ/mol), Eq. [1] J.M. Rosen, Surfactants and Interfacial Phenomena, 3rd ed., Wiley Interscience,
New York, 2004.
(5) [2] A.W. Adamson, A.P. Gast, Physical Chemistry of Surfaces, 6th ed., Wiley Inter-
0
Gmic the standard Gibbs energy of micellization (kJ/mol), Eq. science, New York, 1997.
(17) [3] J. Eastoe, J.S. Dalton, Adv. Colloid Interface Sci. 85 (2000) 103–144.
AB [4] J. Fendler, E. Fendler, Catalysis in Micellar and Macromolecular Systems, Aca-
G1W 1
the Gibbs energy of the interactions between the surface demic, New York, 1975.
active ions through the water resulting from the Lewis [5] M. Czerniawski, Roczniki Chem. 40 (1966) 1265–1271 (in polish).
acid–base forces (J), Eq. (5) [6] P. Carpena, J. Agular, P. Bernaola-Galván, C. Carnero Ruiz, Langmuir 18 (2002)
AB 6054–6058.
GHWH the Gibbs energy of the interactions between heads of
[7] A. González-Pérez, J.L. Del Castillo, J. Czapkiewicz, J.R. Rodrigez, Colloid Surf. A
surfactant through the water resulting from the Lewis 232 (2004) 183–189.
acid–base forces (J), Eq. (6) [8] R. Barchlui, R. Pottel, J. Phys. Chem. 98 (1994) 7899–7905.
AB
GTWT the Gibbs energy of the interactions between tails of [9] C. Carnero Ruiz, Colloid Surf. A 147 (1999) 349–357.
[10] K. Szymczyk, A. Zdziennicka, B. Jańczuk, W. Wójcik, Colloid Surf. A 264 (2005)
surfactant through the water resulting from the Lewis 147–156.
acid–base forces (J), Eq. (6) [11] K. Szymczyk, B. Jańczuk, Colloid Surf. A 293 (2007) 39–50.
134 A. Zdziennicka et al. / Fluid Phase Equilibria 322–323 (2012) 126–134

[12] F. Ysambertt, F. Vejart, J. Paredes, J.-L. Salager, Colloid Surf. A 137 (1998) [45] Y. Rhabi, M.A. Winnik, Adv. Colloid Interface Sci. 89–90 (2001) 25–46.
189–196. [46] S.W. Musselman, S. Chander, J. Colloid Interface Sci. 256 (2002) 91–99.
[13] J.H. Ayala, A.M. Afonso, V. González-Diaz, Microchem. J. 60 (1998) 101–109. [47] M.N. Jones, Int. J. Pharm. 177 (1999) 137–159.
[14] P.H. Elworthy, K.J. Mysels, J. Colloid Sci. 21 (1966) 331–347. [48] E. Ghzaoui, E. Fabregue, G. Cassanas, J.M. Fulconis, J. Delagrange, Colloid Polym.
[15] F. Andriamainty, J. Čižmárik, M. Holíková, Sci. Pharm. 72 (2004) 221–225. Sci. 278 (2000) 321–328.
[16] N. Dubey, J. Chem. Eng. Data 56 (2011) 3291–3300. [49] M.E. McCarroll, A.G. Joly, Z. Wang, D.M. Friedrich, R. von Wandruszka, J. Colloid
[17] R.A. El-Ghazawy, R.A. Raheem, A.M. Al-Sabagh, Polym. Adv. Technol. 15 (2004) Interface Sci. 218 (1999) 260–264.
244–250. [50] M.K. Purkait, S. DasGupta, S. De, J. Hazard. Mater. 137 (2006) 827–835.
[18] J. Aguiar, J.A. Molina-Bolívar, J.M. Peula-García, C. Carnero Ruiz, J. Colloid Inter- [51] N. Kimura, J. Umemura, S. Hayashi, J. Colloid Interface Sci. 182 (1996) 356–364.
face Sci. 255 (2002) 382–390. [52] A. Zdziennicka, J. Colloid Interface Sci. 335 (2009) 175–182.
[19] J.M. del Rio, G. Prieto, F. Sarrniento, V. Mosquera, Langmuir 11 (1995) [53] A. Zdziennicka, Langmuir 26 (3) (2010) 1860–1869.
1511–1514. [54] B. Jańczuk, J.M. Bruque, M.L. González-Martín, C. Dorado-Calasanz, Langmuir
[20] T.M. Perger, M. Bešter-Rogač, J. Colloid Interface Sci. 313 (2007) 288–295. 11 (1995) 4515–4518.
[21] Th.F. Tadros, Surfactants in Agrochemicals, Marcel Dekker Inc., New York, 1994. [55] B. Jańczuk, W. Wójcik, A. Zdziennicka, J. Colloid Interface Sci. 157 (1993)
[22] H.B. Klevens, J. Phys. Colloid Chem. 52 (1948) 130–148. 384–393.
[23] C.J. van Oss, P.M. Constanzo, J. Adhesion Sci. Technol. 4 (1992) 477–487. [56] S. Javadian, H. Gharibi, B. Sohrabi, H. Bijanzadeh, M.A. Safarpour, R.
[24] B. Waligóra, D. Góralczyk, Bull. Acad. Polon Sci. XXII (1974) 901–905. Behjatmanesh-Ardakani, J. Mol. Liquids 137 (2008) 74–79.
[25] C.J. van Oss, R.E. Giesse, P.M. Constanzo, Clay Clay Miner. 38 (1990) 151–159. [57] R. Sowada, Tenside Surfact. Deterg. 31 (1994) 195–199.
[26] B. Jańczuk, J.A. Méndez Sierra, M.L. González-Martín, J.M. Bruque, W. Wójcik, J. [58] S.Y. Kwon, S.H. Lee, Curr. Appl. Phys. 11 (2011) 1042–1047.
Colloid Interface Sci. 184 (1996) 607–613. [59] E. Junquera, E. Aicart, Langmuir 18 (2002) 9250–9258.
[27] A. Zdziennicka, K. Szymczyk, J. Krawczyk, B. Jańczuk, Fluid Phase Equilibr. [60] P. Lianos, R. Zana, J. Colloid Interface Sci. 88 (1982) 594–598.
doi:10.1016/j.fluid.2012.01.014. [61] V.I. Martin, J. Colloid Interface Sci. 363 (2011) 284–294.
[28] K. Tsubone, M.J. Rosen, J. Colloid Interface Sci. 244 (2001) 394–398. [62] D.N. Rubingh, P.M. Holland, Cationic Surfactants: Physical Chemistry. Surfac-
[29] Ch. Das, B. Das, J. Chem. Eng. Data 54 (2009) 559–565. tants Science Series, vol. 37, Marcel Dekker, New York, 1991.
[30] R. Nagarajan, Ch. Wang, Langmuir 16 (2000) 5242–5251. [63] E. Rodenas, C. Dolcet, M. Valiente, E.C. Valeron, Langmuir 10 (1994) 2088–2094.
[31] S. Skerjanc, K. Kogej, J. Cerar, Langmuir 15 (1999) 5023–5028. [64] J. Haldar, V.K. Aswal, P.S. Goyal, S. Bhattacharya, J. Phys. 63 (2) (2004) 303–307.
[32] H. Hoffmann, A. Rauscher, M. Gradzielski, S.F. Shultz, Langmuir 8 (1992) [65] J.J. Galan, A. González-Pérez, J.R. Rodriguez, J. Therm. Anal. Calorimetr. 72 (2003)
2140–2146. 465–470.
[33] A. Callaghan, R. Doyle, E. Alexander, R. Palepu, Langmuir 9 (1993) 3422–3426. [66] K. Behera, M.D. Pandey, M. Porel, S. Pandey, J. Chem. Phys. 127 (2007), art. no.
[34] H.S. Courtney, W.A. Simpson, E.H. Beachey, Infect. Immun. 51 (1986) 184501.
414–418. [67] G.K. Hiller, N. Calkins, R. Wandruszka, Langmuir 12 (1996) 916–920.
[35] P. Murkejee, K.J. Mysels, Critical Micelle Concentrations of Aqueous Surfactant [68] R.J. Robson, E. Dennis, J. Phys. Chem. 81 (1977) 1074–1078.
Systems, NSRDS-NBS 36, National Bureau of Standards, United States Govern- [69] K.M. Kale, R. Zana, J. Colloid Interface Sci. 61 (1977) 312–322.
ment Printing Office, Washington, DC. [70] N. Boden, S.A. Corne, K.W. Jolley, J. Phys. Chem. 1 (1987) 4092–4105.
[36] M. Tanaka, S. Kaneshina, R. Matuura, K. Fukuoka, Univ. Sci. Rep. 4 (1974) [71] L. Benjamin, J. Phys. Chem. 70 (1966) 3790–3797.
131–153. [72] R. De Lisi, S. Milioto, R.E. Verrall, J. Solut. Chem. 19 (1990) 665–692.
[37] E.A.M. Gad, M.M.A. El-Sukkary, D.A. Ismail, JAOCS 74 (1997) 43–47. [73] A. Kitahara, A. Watanabe, Electrical Phenomena at Interfaces, Surfactant Science
[38] S.K. Mehta, K.K. Bhasim, R. Chauhan, S. Dham, Colloid Surf. A 255 (2005) Series, vol. 15, Dekker, New York, 1984.
153–157. [74] A. Groszek, J. Proc. R. Soc. 314 (1970) 473–478.
[39] E. Fisicaro, M. Biemmi, C. Compari, E. Duce, M. Peroni, J. Colloid Interface Sci. [75] P.D.T. Huibers, Langmuir 15 (1999) 7546–7550.
305 (2007) 301–307. [76] J.N. Philiphs, Trans. Faraday Soc. 51 (1955) 561–569.
[40] O. Kosaka, P. Sehgal, H. Doe, Food Hydrocolloid 22 (2008) 144–149. [77] S. Dai, K.C. Tam, Colloid Surf. A 229 (2003) 157–168.
[41] C. Gamboa, A.F. Olea, Colloid Surf. A 278 (2006) 241–245. [78] A. Taheri-Kafrani, A.-K. Bordbar, J. Therm. Anal. Calorim. 98 (2009) 567–575.
[42] H. Benalla, M.J. Meziani, J. Zajac, Colloid Surf. A 238 (2004) 99–108. [79] M.H. Ropers, G. Czichocki, G. Brezesinski, J. Phys. Chem. B 107 (2003)
[43] J. Zajac, Colloid Surf. A 167 (2000) 3–19. 5281–5288.
[44] M. Kumbhakar, S. Nath, T. Mukhrjee, H. Pal, J. Chem. Phys. 121 (2004) [80] S. Mukherjee, D. Mitra, S.C. Bhattacharya, A.K. Panda, S.P. Moulik, Colloid J. 71
6026–6033. (2009) 677–686.

Vous aimerez peut-être aussi