You are on page 1of 7

Colloids and Surfaces A: Physicochem. Eng.

Aspects 259 (2005) 95–101

Effect of salt on the micelles of cetyl pyridinium chloride


D. Varade a,∗ , T. Joshi a , V.K. Aswal b , P.S. Goyal c ,
P.A. Hassan d , P. Bahadur a
aDepartment of Chemistry, South Gujarat University, Surat 395007, India
b
Solid State Physics Division, BARC, Trombay, Mumbai 400085, India
c IUC-DAEF, Mumbai Centre, BARC, Trombay, Mumbai 400085, India
d Novel Materials and Structural Chemistry Division, BARC, Trombay, Mumbai 400085, India

Received 12 June 2004; accepted 14 February 2005


Available online 16 March 2005

Abstract

Surface tension, conductance, viscosity, dynamic light scattering (DLS) and small angle neutron scattering (SANS) measurements were
carried out in order to investigate the effect of added salt NaCl and NaBr on micellization and structure of cetylpyridinium chloride (CPyCl)
micelles. Critical micelle concentration (CMC) of CPyCl decreases with increase in salts concentration, while CMC and counter ion dissociation
(β) increases with increase in temperature. At a given CPyCl concentration, the micelle size increases with increasing salt concentration above
a threshold value. Micellar growth in presence of salts is clearly reflected by an increase in viscosity, hydrodynamic diameter (d) from DLS
and aggregation number from SANS measurements. The effect of NaBr on CPyCl micelles was more pronounced in comparison to NaCl.
The difference in solution properties of Br− and Cl− is that Cl− counter ions being more hydrated are less effective to neutralize the charge
on the micellar surface.
© 2005 Elsevier B.V. All rights reserved.

Keywords: Micellization; SANS; DLS; Micellar growth

1. Introduction Cationic surfactants offer some additional advantages over


other class of surfactants [3–6]. These substances besides
Micelles of ionic surfactants formed in aqueous media their surface activity do show antibacterial properties and
are aggregates of surfactant molecules with different sizes are used as cationic softeners, lubricants, retarding agents,
and shapes depending on given solution conditions. As for antistatic agents and in some cases consumer use etc. Ionic
the structure a micelle of ionic surfactants is composed of a micelles of various cationic surfactants having chloride or
compressive core surrounded by a less compressive surface bromide as counter ion undergo the salt induced sphere-
structure [1,2]. In addition, counter ions are bound to mi- to-rod transition, when NaCl or NaBr is added across a
celle surface. Looking at counter ion behavior, the binding certain threshold salt concentration. Experimentally, many
states are different depending on charge, size and polarity in techniques have been used in order to investigate the sur-
counter ions species themselves. The solution properties of factant concentration, added salt concentration and temper-
surfactant are all reflected from surfactant ions comprising ature dependences of micellar solution properties [7–10].
various combinations of hydrophobic tail with hydrophilic Cetyltrimethylammonium bromide (CTABr) forms rodlike
head and from counter ion species. micelles at concentration above 0.27 M in water at 28 ◦ C, on
the other hand, cetyltrimethylammonium chloride (CTACl)
∗ Corresponding author. Tel.: +91 261 2258384; do not show such a behavior, the aggregates remaining glob-
fax: +91 261 2256012/2227312. ular up to 0.7 M [11]. This point out the specific role played
E-mail address: dharmeshvarade2004@yahoo.com (D. Varade). by the counter ion and existence of spherical and nonspher-

0927-7757/$ – see front matter © 2005 Elsevier B.V. All rights reserved.
doi:10.1016/j.colsurfa.2005.02.018
96 D. Varade et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 259 (2005) 95–101

ical aggregates can be understood in terms of more or less measurement. The flow time for constant volume of solution
efficient repulsions between the polar groups at the micellar through the capillary was measured with a calibrated stop-
surface. It is well known that salts such as KBr and sodium watch.
salicylate (NaSal) induce pronounced growth of CTABr mi-
celles due to charge neutralization at the micellar surface by 2.4. Dynamic light scattering
these salts [12,13]. The counter ion distribution or conden-
sation and size of the ionic micelles play an important role DLS measurements were performed using a Malvern 4800
to decide the effect of salt on the micellar systems [14]. We- Autosizer employing 7132 digital correlator. The light source
ican et al. [15] by means of laser light scattering, 1 H NMR was argon ion laser operated at 514.5 nm with a maximum
measurements and fluorescence probe reported that CTABr output power of 2 W. The average decay rate (Γ ) of the elec-
forms rod-like micelles at 0.1 mol L−1 KBr and the worm tric field autocorrelation function, g1 (τ), was estimated us-
like micelles are formed at above 0.2 mol L−1 KBr. ing the method of cumulants. The apparent diffusion coef-
In this paper we have reported systematic studies on influ- ficients (D) of the micelles were obtained from the relation
ence of NaCl and NaBr salt on the micellization and struc- Γ = Dq2 (q is the magnitude of the scattering vector given
ture of the micelles in cetylpyridinium chloride (CPyCl) so- by q = [4πn sin(θ/2)]/λ, n being the refractive index of the
lution as studied by surface tension, conductance, viscos- solvent, λ, the wavelength of laser light and θ is the scatter-
ity, dynamic light scattering (DLS) and small angle neutron ing angle) and the corresponding hydrodynamic diameters
scattering (SANS) measurements. The results are interpreted (d) were calculated using the Stokes–Einstein relationship.
in terms of counter ion condensation on the ionic micelles, For all the solutions, Γ varies linearly with ‘q2 ’ indicating
which depends on the hydrated size of the counter ions. translational diffusion of the scatterers.

2.5. Small angle neutron scattering


2. Materials and methods
For SANS measurements, the sample solutions were pre-
The cationic surfactant cetylpyridinium chloride (CPyCl), pared in D2 O. The use of D2 O as solvent instead of H2 O
salt NaCl and NaBr were highly pure samples purchased from provides better contrast in neutron scattering experiments.
Fluka, Switzerland and were used as received. Triply distilled In all the measurements, CPyCl concentration (0.25 M) and
water from an all Pyrex glass apparatus was always used for temperature (30 ± 0.1) ◦ C were kept constant. The solutions
the preparation of solutions for surface tension, conductance were held in a quartz cell of 0.5 cm thickness with tight-fitting
and viscosity measurements. For DLS measurement Milli-Q Teflon stoppers. The SANS experiments were performed us-
water having specific resistance 18.2 M cm was used. The ing a SANS diffractometer at the Dhruva reactor, BARC,
samples for SANS experiments were prepared in D2 O. Trombay [17]. The diffractometer uses a polycrystalline BeO
filter as a monochromator. The mean wavelength of the in-
2.1. Conductance cident neutron beam is 5.2 Å with a wavelength resolution
( λ/λ) of approximately 15%. The angular distribution of
Conductometric measurements were made with a digital the scattered neutron was recovered by a linear 1 m long He3
conductivity meter (Phillips, India) using a dip-type cell. All position sensitive detector (PSD). The data were recorded in
measurements were done in a jacketed vessel, which was the Q range of 0.02–0.24 Å−1 . All the measured SANS distri-
maintained at the appropriate temperature (±0.1 ◦ C). The er- butions were corrected for the background and solvent con-
rors in the conductance measurements were within ±0.5%. tributions and the data were normalized to the cross-sectional
The conductance was measured after thorough mixing and unit using standard procedures [17].
temperature equilibrium at each dilution.

2.2. Surface tension 3. Results and discussion

The surface tension of CPyCl solutions in absence and Conductivity measurements were performed in water for
presence of salts was measured by drop weight method using CPyCl at various temperatures (30–60 ◦ C at an interval of
a modified stalagmometer [16]. 10 ◦ C) in order to evaluate the CMC and the degree of counter
ion dissociation, β. The plots of conductance data are pre-
2.3. Viscosity sented as specific conductance versus concentration in Fig. 1.
The values of CMC and β at various temperatures for CPyCl
The viscosity measurements were carried out using an solution are recorded in Table 1. The values of CMC and
Ubbelohde suspended level capillary viscometer. The vis- β increase with temperature in the range investigated. The
cometer was always suspended vertically in a thermostat with effect of temperature on the CMC of surfactants in aqueous
a temperature stability of ±0.1 ◦ C in the investigated region. solution is usually analyzed in terms of two opposing factors.
The viscometer was cleaned and dried every time before each First, as the temperature increases the degree of hydration of
D. Varade et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 259 (2005) 95–101 97

Table 1
The CMCs, degree of counter ion dissociation (β), standard molar Gibbs energies ( G◦ ), enthalpies ( H◦ ) and entropies ( S◦ ) of micellization for CPyCl in
water at different temperatures
Temperature (K) CMC (mM) β G◦ (kJ mol−1 ) H◦ (kJ mol−1 ) S◦ (kJ mol−1 )
303.2 0.98 0.41 −27.59 −10.17 0.057
313.2 1.11 0.43 −28.17 −10.85 0.055
323.2 1.30 0.45 −28.65 −11.55 0.053
333.2 1.45 0.46 −29.23 −12.28 0.050

the hydrophilic group decreases, which favors micellization; Fig. 2. The slope in the plot was taken as [(∂ln XCMC )/∂T].
however, an increase in temperature also causes the disrup- The standard molar entropy of micelle formation ( Sm◦ ) was

tion of the water structure surrounding the hydrophobic group calculated from
and this is unfavorable to micellization. It seems from the data
 
in Table 1 that this second effect is predominant in the tem- ◦ Hm◦ − G◦m
perature range studied. On the other hand, β, also increases Sm = (3)
T
regularly with temperature increase. This observed increase
in β is probably due to a decrease in the charge density at the
The entropy change is positive in all cases. However, it
micellar surface caused by the decrease in the aggregation
decreases with increasing temperature. The thermodynamic
number of the micelle.
parameters of micellization obtained by following the above
The CMC values, as determined at various temperatures,
procedure are summarized in Table 1.
were further used for calculation of the thermodynamic pa-
The surface tension (γ) of CPyCl solutions in absence and
rameters of micellization. In the charged pseudo-phase model
presence of NaCl and NaBr was measured for a range of con-
of micelle formation, the standard free energy of micelle for-
centrations above and below the critical micelle concentration
mation per mole of surfactant is given by
(CMC). As shown in the Fig. 3 (for NaCl), a linear decrease
G◦m = (2 − β)RT ln XCMC (1) in surface tension was observed with increase in surfactant
concentrations up to the CMC, beyond which no considerable
where R is the gas constant, T is the temperature in Kelvin change was noticed. The electrical atmosphere in the aqueous
scale and XCMC stands for the CMC in the mole fraction unit. surfactant solution altered in presence of NaCl which neutral-
The standard enthalpy of micelle formation ( Hm◦ ) can be izes the effective head group charge resulting in reducing the
derived by the van’t Hoff equation electrostatic repulsion between the polar head groups and the
 
∂ ln XCMC micelles are formed at much lower concentration as compared
Hm◦ = −(2 − β)RT 2 (2) to that in pure water. The CMC of the surfactant decreased
∂T
in the presence of NaCl and NaBr, the decrease being de-
The enthalpy of micellization can be obtained if the de- pendent upon the concentration of added salt. The effect was
pendence of the CMC on temperature is known. One can see more pronounced in case of NaBr. The area occupied by the
that the standard enthalpy of micelle formation is more neg- surfactant molecules in Å2 , at the air–water interface at satu-
ative or exothermic on higher temperature side. A linear plot ration monolayer was estimated using the surface tension (γ)
was observed of ln XCMC against T for CPyCl as shown in

Fig. 1. Plots of specific conductivity vs. concentration of CPyCl in water at


different temperatures. Fig. 2. Plots of ln XCMC vs. temperature for CPyCl in water.
98 D. Varade et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 259 (2005) 95–101

Fig. 3. Plots of surface tension vs. log CPyCl concentration in absence and Fig. 4. Relative viscosity vs. concentration of CPyCl in absence and presence
presence of varying amount of NaCl at 30 ◦ C. of NaCl concentration at 30 ◦ C.

data at temperature T and the Gibbs adsorption isotherm:


 
1 dγ
Γs − (4)
nRT d ln C
where Γ s is surface excess, R is the universal gas constant,
C is the concentration of the surface active compound and
n is taken as 2 because of the ionic nature of the surfactant.
Interfacial properties of CPyCl in water at 30 ◦ C are recorded
in Table 2.
The viscosity measurement for micellar solution of CPyCl
were made in absence and presence of NaCl and NaBr at
30 ◦ C in order to study the transition of micelle shape induced
by the change in ionic strength. The results obtained in NaCl
were quite interesting (Fig. 4). CPyCl shows a gradual in-
crease in viscosity with increase in concentration in absence Fig. 5. Variation of hydrodynamic diameter of 0.25 M CPyCl in varying salt
of NaCl. But the electrical atmosphere in the aqueous sur- concentration at 30 ◦ C.
factant solution is profoundly altered by the added NaCl and
this is turn influences the viscous flow of CPyCl. Initially, crease in viscosity at low salt concentration followed by a
the solution of CPyCl becomes less viscous with increase in dramatic increase at high salt concentration. CPyCl shows
the concentration in presence of lower NaCl concentration, an increase in viscosity (figure not shown) at much lower
but a dramatic increase in viscosity was observed in presence concentration of NaBr (0.2 M) as compared to NaCl (1.0 M).
of 1.0 M NaCl. This is evidently an indication of decrease The changes in the aggregation behavior of CPyCl mi-
in the electrostatic interaction between micelles which will celles in presence of NaCl and NaBr can be inferred from
be reflected as a decrease in the “effective volume fraction” DLS measurements. The apparent hydrodynamic diameter
in the presence of low NaCl concentration [19]. From the (d) of the micelles in presence of different amounts of NaCl
above results we can conclude that only above a certain salt and NaBr are depicted in Fig. 5. Considering the ionic nature
concentration there is a transition from globular to rodlike of the surfactant, in the absence of any salt, the measured
micelles. Almgren and coworkers [20] has also studied the diffusion coefficients will be modified due to the presence of
effect of varied NaCl concentration on the relative viscosity repulsive intermicellar interactions. Such repulsive interac-
of hexadecyltrimethylammonium bromide (CTABr, 25 mM) tions will lead to an increase in the diffusion coefficient and
solution, in which they have reported the similar initial de- hence a decrease in the apparent diameter of the micelles.

Table 2
Interfacial properties of CPyCl in water at 30 ◦ C
Surfactants CMC (mM) C␲ = 20 (mM) γ CMC (mN m−1 ) Area/molecule (Å2 )
Experimental Literature [reference]
CPyCl 0.95 1.00 [18] 0.5 42.0 42.0
D. Varade et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 259 (2005) 95–101 99

Thus, the observed diameter of the micelles in the absence of 1/2


x = Q[a2 µ2 + b2 (1 − µ2 )] (9)
salts will be slightly underestimated as no account of inter-
particle interaction has been taken into consideration. This where a and b are, respectively, the semimajor and semiminor
is evident from the smaller diameter of micelles (2 nm) as axes of the ellipsoidal micelle and µ is the cosine of the angle
compared to that expected from the length of hydrocarbon between the directions of a and the wave vector transfer Q.
chain of the surfactant (3.2 nm). With addition of salts, an in- In general, micellar solutions of ionic surfactants show
crease in the apparent diameter of the micelles from ∼2 nm a correlation peak in the SANS distribution [24]. The peak
to ∼8 nm is observed suggesting an increase in size of the arises because of the corresponding peak in the interparticle
micelles. This observed change in apparent size arises from structure factor S(Q) and indicates the presence of electro-
both decrease of intermicellar interaction and growth of the static interactions between the micelles. S(Q) specifies the
micelles. Considering large changes in the relative viscosity correlation between the centers of different micelles and it
of the solution with addition of salt, a prolate ellipsoidal or is Fourier transform of the radial distribution function g(r)
rod-like growth of the micelles can be envisioned. The ob- for the mass centers of the micelle. Unlike the calculation of
served micellar growth is significantly different for NaCl and P(Q), it is quite complicated to calculate S(Q) for any other
NaBr. In presence of NaCl, the apparent diameter reaches a shape than spherical. This is because of S(Q) depends on the
value of 8 nm at a salt concentration of 1.2 M while the same shape and as well as on the orientation of the particles. To
effect could be observed for NaBr even at 0.2 M. This sug- simplify this, prolate ellipsoidal micelles are assumed to be
gests that the nature of the condensation of counter ions on equivalent spherical. We have calculated S(Q) as derived by
the micelle surface is different for Cl− and Br− . Hayter and Penfold [25] from the Ornstein–Zernike equa-
In SANS experiments one measures the differential scat- tion and using the mean spherical approximation. The mi-
tering cross-section per unit volume (dΣ/dΩ) as a function of celle is assumed to be a rigid equivalent sphere of diameter
scattering vector Q. For a system of monodisperse interacting σ = 2(ab2 )1/3 interacting through a screened Coulomb poten-
micelles (dΣ/dΩ) is given by [21] tial, which is given by
dΣ exp[−κ(r − σ)]
(Q) = n(ρm − ρs )2 V 2 [F (Q)2  u(r) = u0 σ , r > σ, (10)
dΩ r
+F (Q)2 (S(Q) − 1)] + B (5) where κ is the Debye–Huckel inverse screening length and is
calculated by
where n denotes the number density of the micelles, ρm and ρs
are, respectively, the scattering length densities of the micelle  1\2
8πNA e2 I
and the solvent and V is the volume of the micelle. F(Q) is the κ= (11)
103 εkB T
single particle form factor and S(Q) is the interparticle struc-
ture factor. B is a constant term that represents the incoherent defined by the ionic strength I of the solution
scattering background, which is mainly due to hydrogen in
I = CMC + 21 αC + Cs, (12)
the sample.
The micelles formed at the CMC are spherical. If the so- where I is determined by the CMC, dissociated counter ions
lution conditions (e.g., concentration, ionic strength, etc.) of from the micelles and the salt concentration. The fractional
the micellar solutions are changed that favors the growth of charge α (=Z/N, where Z is the micellar charge) is the charge
the micelles, they grow along one of the axial directions of per surfactant molecule in the micelle and is a measure of
the micelles. The growths of the micelles along other two the dissociation of the counter ions of the surfactant in the
axial directions are restricted by the length of the surfactant micelle. C and Cs present the concentrations of the surfactant
molecule to avoid the energetically unfavorable any empty and salt in the solution, respectively. The contact potential u0
space or water penetration inside the micelle [22,23]. The is given by
prolate ellipsoidal shape (a = b = c) of the micelles is widely
used in the analysis of SANS data because it also represents Z 2 e2
u0 = (13)
the other different possible shapes of the micelles such as πεε0 σ(2 + κσ)2
spherical (a = b) and rod-like (a b). For such an ellipsoidal
where ε is the dielectric constant of the solvent medium, ε0
micelle
 1 is the permittivity of free space and e is the electronic charge.
The dimensions of the micelle, aggregation number and
F 2 (Q) = [F (Q, µ)2 dµ] (6)
0 the fractional charge have been determined from the anal-
 2 ysis. The semimajor axis (a), semiminor axis (b = c) and
1
the fractional charge (α) are the parameters in analyzing the
F (Q)2 = F (Q, µ) dµ (7)
0 SANS data. The aggregation number is calculated by the
relation N = 4πab2 /3v, where v is the volume of the sur-
3(sin x − x cos x) factant monomer. Throughout the data analysis corrections
F (Q, µ) = (8)
x3 were made for instrumental smearing [26]. The parameters
100 D. Varade et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 259 (2005) 95–101

Table 3
Micellar parameters from 0.25 M CPyCl in presence of varying NaCl and NaBr concentration at 30 ◦ C
Salt [Salt] (M) Aggregation number, N Fractional charge, α Semiminor axis, b = c (Å) Semimajor axis, a (Å) Axial ratio a/b
NaCl 0.0 104 0.2 18.1 44.3 2.4
0.1 113 0.2 18.7 45.3 2.4
0.5 131 0.2 19.5 48.4 2.5
1.0 229 0.2 18.4 94.4 5.1
NaBr 0.1 139 0.2 19.2 52.7 2.7
0.2 368 0.07 19.5 136.0 7.0

in the analysis were optimized by means of nonlinear least- relation peak is an indication of strong repulsive interaction
square fitting program and the errors of the parameters were between the positively charged CPyCl micelles. The peak
calculated by the standard methods used [27]. usually occurs at Qm ∼ 2π/d, where d is the average distance
Figs. 6 and 7 show the SANS data from 0.25 M CPyCl between the micelles and Qm is the value of Q at the peak
micellar solution with and without the addition of salt NaCl position. As shown in the figure cross-section increases and
and NaBr at temperature 30 ◦ C, respectively. The SANS dis- the peak position shift to lower Q values with the increase in
tribution from a pure 0.25 M CPyCl shows a well defined the salt concentration. This reflects the increase in the size of
peak at the wave vector transfer Q ∼ 0.076 Å−1 . This cor- the micelles in the presence of salt [7,14]. The broadening of
the correlation peak is due to screening of charge by the salt
between the micelles. The micellar parameters obtained for
these systems are given in Table 3.
It is seen that in case of NaCl the fractional charge on
the micelle remains nearly same but the aggregation num-
ber increases when the NaCl concentration in the micellar
solution is increased. This indicates that the counter ion has
equal affinity to stay in water as well as condense on the
micelles. For the same concentration of the salt NaCl and
NaBr, the charge neutralization on the ionic CPyCl micelles
is different for these salts, and this leads to the formation of
different micellar structures when NaCl and NaBr are added.
For example the aggregation number of CPyCl micellar so-
lution increases from 104 to 229 upon addition of 1 M NaCl;
while aggregation number increases upto 368 by the addi-
tion of only 0.2 M NaBr. These results are well supported by
viscosity and DLS measurements discussed earlier.
The charge neutralization at the surface of the micelle by
Fig. 6. SANS data from 0.25 M CPyCl micellar solution with varying con-
centration of NaCl. the increase in the counter ion condensation decreases the
effective head group area for the surfactant monomer to oc-
cupy in the micelle and hence the increase in the aggregation
number of the micelle. The difference in solution properties
of Br− and Cl− is that Cl− counter ions are more hydrated.
This suggests that the hydrated size of the counter ion similar
to the size of the micelle plays an important role to control
the properties of micellar solution. The comparison of the
micellar parameters in Table 3 suggests that counter ion con-
densation is more effective to neutralize the charge on the
micellar surface when the hydrated size is small. The solid
lines in Figs. 6 and 7 are fitted curves to the experimental
data using Eq. (5).

4. Conclusions

Fig. 7. SANS data from 0.25 M CPyCl micellar solution with varying con- We have studied the effect of added salt NaCl and NaBr
centration of NaBr. on micellization and structure of cetylpyridinium chloride
D. Varade et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 259 (2005) 95–101 101

(CPyCl) micelles. It was found that the micelle size increases [8] V.K. Aswal, P.S. Goyal, Phys. Rev. 67 (2003) 051401.
with increasing salt concentration above a threshold value. [9] B. Simoncic, J. Span, Acta Chim. Slov. 45 (1998) 143.
[10] K. Fujio, T. Mitsui, H. Kurumizawa, Y. Tanaka, Y. Uzu, Colloid
The observed micellar growth is significantly different for
Polym. Sci. 282 (2004) 223.
NaCl and NaBr. The effect of NaBr on CPyCl micelles was [11] J. Ulmius, B. Lindman, G. Lindblom, T. Drakenberg, J. Colloid
more pronounced in comparison to NaCl. The results are Interface Sci. 65 (1978) 88.
interpreted based on the dependence of the counter ion con- [12] M. Cates, S.J. Candau, J. Phys. Condens. Matter 2 (1990) 6889.
densation. [13] H. Rehage, H. Hoffman, Mol. Phys. 74 (1991) 933.
[14] V.K. Aswal, P.S. Goyal, Phys. Rev. 61 (2000) 2947.
[15] Z. Weican, L. Ganzuo, M. Jianhai, S. Qiang, Z. Liqiang, L. Haojun,
W. Chi, Chin. Sci. Bull. 45 (2000) 1854.
Acknowledgements [16] J. Mata, D. Varade, G. Ghosh, P. Bahadur, Colloids Surf. A 245
(2004) 69.
Financial assistance from IUC-DAEF Project No. [17] V.K. Aswal, P.S. Goyal, Curr. Sci. 79 (2000) 947.
IUC/CRS-M-103/2001 and help from J. Mata for DLS mea- [18] M.S. Bakshi, S. Mahajan, J. Jpn. Oil Chem. Soc. 49 (2000) 17.
surements is gratefully acknowledged. [19] K.R. Choudhury, D.K. Majumdar, Electrochim. Acta 28 (1983)
597.
[20] M.S. Vethamuthu, M. Almgren, G. Karlsson, P. Bahadur, Langmuir
12 (1996) 2173.
References [21] S.H. Chen, T.L. Lin, in: D.L. Price, K. Skold (Eds.), Methods of
Experimental Physics, vol. 23B, Academic Press, New York, 1987,
[1] Y. Moroi, Micelles Theoretical and Applied Aspects, Plenum Press, p. 489.
New York, 1992. [22] V. Degiorgio, M. Corti, Physics of Amphiphiles: Micelles, Vesicles
[2] D.M. Bloor, J. Gormally, E.W. Jones, J. Chem. Soc. Faraday Trans. and Microemulsion, Amsterdam, North-Holland, 1985.
80 (1984) 1915. [23] J.N. Israelachvili, Intermolecular and Surface Forces, Academic
[3] E. Jungerman, Cationic Surfactants, Marcel Dekker, New York, 1969. Press, New York, 1992.
[4] J. Cross, E.J. Singer, Cationic Surfactants: Analytical & Biological [24] S.H. Chen, E.Y. Sheu, J. Kalus, H. Hoffmann, J. Appl. Cryst. 21
Evaluation, Marcel Dekker, New York, 1994. (1988) 751.
[5] , P.M. Holland, D.N. Rubingh (Eds.), Cationic Surfactants: Physical [25] J.B. Hayter, J. Penfold, Colloid Polym. Sci. 261 (1983) 1022.
Chemistry, Marcel Dekker, New York, 1991. [26] J.S. Pedersen, D. Posselt, K. Mortensen, J. Appl. Cryst. 23 (1990)
[6] J.M. Richmond, Cationic Surfactants: Organic Chemistry, Marcel 321.
Dekker, New York, 1990. [27] P.R. Bevington, Data Reduction and Error Analysis for Physical Sci-
[7] V.K. Aswal, P.S. Goyal, Chem. Phys. Lett. 357 (2002) 491. ences, McGraw-Hill, New York, 1969.

You might also like