You are on page 1of 9

J. Chem.

Thermodynamics 44 (2012) 107–115

Contents lists available at SciVerse ScienceDirect

J. Chem. Thermodynamics
journal homepage: www.elsevier.com/locate/jct

Aggregation behavior and intermicellar interactions of cationic Gemini


surfactants: Effects of alkyl chain, spacer lengths and temperature
Marjan Hajy Alimohammadi a, Soheila Javadian a,⇑, Hussein Gharibi a, Ali reza Tehrani-Bagha b,
Mohammad Rashidi Alavijeh a, Karim Kakaei c
a
Department of Physical Chemistry, Tarbiat Modares University, P.O. Box 14115-117, Tehran, Iran
b
Institute for Colour Science and Technology, Tehran, Iran
c
Department of Chemistry, Maragheh University, Maragheh, Iran

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

Article history: The aggregation behavior of the cationic Gemini surfactants CmH2m+1N(CH3)2(CH2)S (CH3)2 N CmH2m+1,2Br
Received 29 May 2011 with m = 12, 14 and s = 2, 4 were studied by performing surface tension, electrical conductivity, pulsed field
Received in revised form 2 August 2011 gradient nuclear magnetic resonance (PFG-NMR), and cyclic voltammetry (CV) measurements over the
Accepted 6 August 2011
temperature range 298 K to 323 K. The critical micelle concentration (CMC), surface excess (Umax), mean
Available online 16 August 2011
molecular surface area (Amin), degree of counter ion dissociation (a), and the thermodynamic parameters
of micellization were determined from the surface tension and conductance data. An enthalpy–entropy
Keywords:
compensation effect was observed and all the plots of enthalpy–entropy compensation exhibit excellent
Surface tension
Enthalpy–entropy compensation
linearity. The micellar self-diffusion coefficients (Dm) and intermicellar interaction parameters (kd) were
Thermodynamic obtained from the PFG-NMR and CV measurements. These results are discussed in terms of the intermicel-
Micellization lar interactions, the effects of the chain and spacer lengths on the micellar surface charge density, and the
Gemini surfactants phase transition between spherical and rod geometries. The intermicellar interaction parameters were
Aggregation found to decrease slightly with increasing temperature for 14–4–14, which suggests that the micellar sur-
Intermicellar interaction face charge density decreases with increasing temperature. The mean values of the hydrodynamic radius,
Rh, and the aggregation number, Nagg, of the Gemini surfactants’ m–4–m micelles were calculated from the
micellar self-diffusion coefficient. Moreover, the Nagg values were calculated theoretically. The experimen-
tal values of Nagg increase with increases in the chain length and are in good agreement with both previous
results and our theoretical results.
Ó 2011 Elsevier Ltd. All rights reserved.

1. Introduction factors underlying the variation of their thermodynamic properties


with the lengths of their chains and spacers [18–28].
Gemini surfactants are composed of two hydrophobic chains There is no doubt that understanding the micellization requires
and two hydrophilic head groups connected by a spacer that can its complete thermodynamic characterization. The thermodynamic
be hydrophilic or hydrophobic and rigid or flexible [1,2]. These sur- changes associated with adsorption and micelle formation have
factants have attracted considerable interest within both academic been investigated from the theoretical and experimental view-
and industrial circles over the last decade because Gemini surfac- points by many research groups.
tants are superior to the corresponding conventional monomeric In the present study, we carried out an investigation of the aggre-
surfactants in a number of aspects. Gemini surfactants have a high- gation behaviors of several Gemini surfactants, including their ther-
er surface activity, a lower CMC [3], better solubilizing power [4], modynamics of micellization, adsorption in the air–liquid interface,
much lower Krafft points [3–5], and better wetting [6], viscoelas- the structures of their aggregates, and the variation of their intermi-
ticity, gelification, and shear thickening [4] than the corresponding cellar interactions with temperature. We used techniques including
conventional monomeric surfactants. Owing to their remarkable surface tension, conductometry, PFG-NMR, and CV. First, the ther-
properties, considerable effort has been invested in the design modynamic properties of micellization were determined by using
and synthesis of new Gemini surfactants [7–12] to study the rela- CMC and a values. Second, the PFG-NMR and CV measurements en-
tionship between their molecular structures and their aggregation abled the investigation of the structures of the surfactants’ aggre-
morphologies in aqueous solution [13–17] and to examine the gates and their intermicellar interactions. Finally, by comparing
the results for these surfactants, we determined not only the effects
⇑ Corresponding author. Fax: +98 21 82883755. of varying the number of carbon atoms of the alkyl chain on the
E-mail addresses: javadian_s@modares.ac.ir, javadians@yahoo.com (S. Javadian). solution properties of the Gemini surfactants but also the effects

0021-9614/$ - see front matter Ó 2011 Elsevier Ltd. All rights reserved.
doi:10.1016/j.jct.2011.08.007
108 M. Hajy Alimohammadi et al. / J. Chem. Thermodynamics 44 (2012) 107–115

of varying the hydrophobicity of the spacer chain on micelle forma- electroactive probe and KCl (0.1 mol  dm3) was used as the sup-
tion. In addition to the determination of the abovementioned micel- porting electrolyte. Note that this concentration of the electroac-
lar parameters, this research demonstrates, for the first time, the tive probe does not significantly affect the micellization of the
usefulness of CV in the study of the effects of increasing the lengths surfactants and their mixtures [33].
of the hydrophobic tail and spacer of Gemini surfactants and those
of varying the temperature on their micellar parameters, intermi-
2.1.5. Computational methods
cellar interactions, and sphere to rod geometry changes.
DFT (density functional theory) calculations were carried out
with the Gaussian 98, Revision A.7 (Gaussian, Inc., Pittsburgh PA).
2. Experimental First, geometry optimization was performed at the B3LYP/6-
311+G(d,p) level in the gas phase. PCM (polarizable continuum
2.1. Materials model) calculations were then performed at the B3LYP level with
the 6-311+G(d,p) basis set and the gas-phase optimized geome-
Gemini surfactants with quaternary ammonium bromide head tries. The micellar radius was determined as the distance between
groups and linear alkyl tails with the general formula [CmH2m+1– the quaternary nitrogen and the farthest H of the terminal methyl
N+–(CH3)2–(CH2)n–(CH3)2–N+–CmH2m+1]2 Br, where m = 12 and group. The value of Nagg was calculated from this radius and the
14 and n = 2 and 4, referred to here as m–n–m, 2 Br surfactants, volume of the hydrated monomer, as obtained with the PCM by
were synthesized and purified by the known methods reported assuming that the micellar aggregates are spherical.
previously [29].

3. Results and discussion


2.1.1. Surface tension measurements
The surface tension measurements were performed with a
3.1. The CMC and the thermodynamics of micellization
Krüss K12 tensiometer by using the ring method [30,31]. The plat-
inum ring was thoroughly cleaned and flame dried before each
The CMC values for m–s–m, (m = 12, 14 and s = 2, 4) have been
measurement. The uncertainty of the measurements was
obtained with conductivity and surface tension measurements;
±0.1 mN  m1. In all cases, more than three successive measure-
representative examples for both techniques are shown in figures
ments were carried out, and the standard deviation did not exceed
SM-1 and SM-2 respectively. The agreement between the CMCs de-
more than 0.08 mN  m1. The temperature was maintained at a
rived from these two methods is satisfactory; note that the CMC
constant value by circulating the thermostatted water through
generally depends on the determination method (table 1). As indi-
the jacketed vessel containing the solution. The temperature con-
cated in table 1, the CMC values for 14–2–14 and 14–4–14 are not
trol accuracy was within ±0.1 K.
strongly dependent on temperature within the range 298 K to
318 K, whereas the CMC values of 12–2–12 and 12–4–12 initially
2.1.2. Electrical conductivity measurements decrease and then increase as the system temperature increases.
The measurements of the conductivity of the surfactant solu- The initial decrease in the CMC is a direct consequence of the de-
tions were performed with a conductometer, Jenway 4510. After crease in the hydrophilicity of the surfactant molecules
investigating the conductivity of the solvent, three successive mea- [18,19,34]. The increase in temperature also results in an increase
surements of the conductivity of the surfactant solutions were car- in the breakdown of the structure of the water molecules sur-
ried out at a controlled constant temperature. The uncertainty of rounding the hydrophobic alkyl group, which is unfavorable to
the measurements was ±0.01 lS  cm1. The employed conductiv- the formation of micelles. As a result, the onset of micellization
ity cell was shown in SM. tends to occur at higher concentrations as the temperature in-
creases [34]. The value of the CMC decreases with increases in
2.1.3. NMR measurements the hydrophobic chain length of the Gemini surfactants (table 1).
NMR self-diffusion measurements were performed on a Bruker Increasing the spacer length of the Gemini surfactants also results
DRX 500 Avance NMR spectrometer at room temperature. A longi- in increases in the CMC (table 1). These results are in good agree-
tudinal eddy-current delay with a bipolar pulse pair (LEDBPP) ment with those of previous studies [17,21].
pulse sequence [32] was used to determine the self-diffusion coef- According to the Frahm’s Method, the a was calculated from the
ficients (D). In this technique, micelle self-diffusion is monitored by ratio of the slopes above and below the CMC in plots of conductiv-
analyzing the signals from trace amounts of tetramethylsilane ity against surfactant concentration [35]. The a values are tabu-
(Me4Si) added to the solution, which are assumed to be completely lated in table 1 [35]. Frahm’s method is an approximate method
solubilized in the micellar phase. All NMR measurements were since the contribution of conductivity of micelles above the CMC
performed at 25 °C. The basic sequence was used with pulse field is neglected. Then, the a values were calculated for 12–4–12 and
duration, d, of 5 ms and a time interval, D, of 200 ms between gra- 14–4–14 in T = 298 K using Evans’ method (table 1) [36]. As shown
dient pulses. in table 1, the difference of a values calculated by Evans to the
method of Frahm is more than 100%. Consequently, ignoring the
2.1.4. CV measurements micellar contribution leads to large differences in the a values cal-
CV was performed by using an electrochemical analyzer, SAMA culated by two methods. However, use of the Evans method re-
500. A three-electrode system consisting of a working Pt electrode quires knowledge of Nagg, and these are not always available.
(the interface at which the redox reaction of interest occurs), a sat- As can be seen in table 1, the a increases with increases in the
urated Ag/AgCl reference electrode with a salt bridge containing chain length of the m–n–m Gemini surfactant homologues. This re-
3 M aqueous KCl solution (the electrode with a stable potential sult indicates that the binding of the bromide counter ions to the
against which the potential at the working electrode is measured) surfactant ammonium heads weakens with increasing hydropho-
and the counter electrode (a Pt wire through which all the current bicity. The increases in a (table 1) within the series of surfactants
required to maintain the redox reaction at the working electrode indicate that increasing the length of the hydrophobic spacer
flows). The working electrode was polished with slurry of alumina means that the distance between the charges increases, and as a
powder and then washed carefully with distilled water before each result the necessity of partial neutralization by counter ions be-
measurement. Ferrocene (0.001 mol  dm3) was used as the comes less. Table 1 shows that the a value of each species tends
M. Hajy Alimohammadi et al. / J. Chem. Thermodynamics 44 (2012) 107–115 109

TABLE 1
CMC, a, interfacial and thermodynamic parameters of micellization of Gemini surfactants at different temperatures

Compound T/k CMCa  104/ CMCb  104/ ab Cmax  106/ Amin/ cCMC/ DHfe/ Tc/K
(mol  dm3) (mol  dm3) (mol  m2) (nm2  molecule1) (mN  m1) (kJ  mol1)
12–2–12 298 8.09 8.40, 8.05c 0.20, 0.13c 1.69 0.98 26.1 34.78 ± 0.01 293.3 ± 0.1
303 7.75 9.09, 9.1c 0.21, 0.13c 1.71 0.97 26.9
308 7.75 9.35, 9.1c 0.23, 0.13c 1.70 0.98 26.9
313 8.42 10.1.10c 0.23, 0.11c 1.63 1.02 26.6
318 9.09 10.9 0.24 1.50 1.11 26.2
12–4–12 298 9.45 10.7 0.24(0.14) 1.58 1.05 36.2 32.09 ± 0.05 290.0 ± 0.3
303 8.70 11.0, 15d 0.27, 0.32d 1.53 1.09 35.4
308 8.93 11.6, 16d 0.29, 0.35d 1.45 1.14 34.9
313 9.45 12.5, 19d 0.31, 0.38d 1.27 1.30 32.1
318 10.7 12.6, 19d 0.34, 0.41d 1.18 1.40 31.6
14–2–14 308 1.47 1.49, 1.9c 0.24, 0.15c 2.23 0.74 27.8 40.40 ± 0.03 304.3 ± 0.1
313 1.47 1.49, 1.9c 0.25, 0.16c 2.19 0.76 27.2
318 1.55 1.77 0.29 1.85 0.90 27.1
323 1.55 1.78 0.31 1.42 1.17 27.2
333 1.62 1.78 0.36 1.07 1.55 25.7
14–4–14 298 1.61 1.46 0.34(0.160) 1.68 0.99 34.1 36.09 ± 0.01 295.9 ± 0.1
303 1.59 1.51 0.35(0.165) 1.64 1.01 32.7
308 1.59 1.72 0.37(0.168) 1.61 1.03 32.3
313 1.59 1.76 0.38(0.171) 1.40 1.18 32.5
318 1.61 1.75 0.40(0.180) 1.24 1.33 32.4

The a values in parentheses were calculated using Evans method.


a
The CMC values obtained from surface tension measurements.
b
The CMC and a values obtained from conductometry measurements.
c
Data taken from reference [18].
d
Data taken from reference [23].
e
The thermodynamic parameters were calculated by using the CMC values obtained from surface tension measurements and a values calculated according to the Frahm’s
Method.
f
These temperatures are below the Kraft temperature of 14–2–14.

to increase with temperature. Note that even though a shallow orientation of the surfactant molecule at the interface is almost
minimum is observed in the CMC–temperature curves, the overall perpendicular to the interface. As expected, the trend in Amin is in-
trend in the respective curves is of increases in the CMC with tem- verse with temperature (table 1).
perature. If the CMC is regarded as reflecting the solubility of the The variations of CMC and a with temperature enable us to
micelles, a increases with temperature in parallel with the in- analyze the thermodynamic energetics of micelle formation in
creased solubility of the micelles. The a is determined by various water. The collected data for the variations of CMC and a with tem-
factors such as the temperature, the electrical potential around perature are given in table 1 and enable us to analyze the thermo-
the micelle, the micellar radius, and the dielectric constant of the dynamic energetics of micelle formation in water. First, the
medium, etc. [37]. The gross features of counter ion dissociation standard Gibbs free energy change upon micelle formation was
or distribution between the kinetic micelle and the bulk solution examined. By applying mass action law, the Gibbs free energy of
can be described with simplified electrostatic models [37], in micellization has been derived for general type of ionic surfactant
which the stabilization of micelles by counter ion binding results containing i polar groups of valency zs bonded to j alkyl chains,
from a balance between the kinetic motion of the ions and the elec- with counterions of valency zc as follow [21,40]:
tric potential on the micellar surface; this balance eventually re-  
1 i
sults in increases in a with increases in temperature. In other DGmic =ðkJ  mol Þ ¼ RT 1=j þ ð1  aÞ jzs =zc j ln X CMC
words, the binding of counter ions (their adsorption onto the j
micellar surface) is an exothermic process whereas the electrolytic þ RTði=jjzs =zc jð1  aÞ ln i=jjzs =zc j  ln j=jÞ; ð1Þ
dissociation of micelles (the desorption of counter ions) is an endo-
where XCMC and a are the CMC value on the mole fraction scale and
thermic process.
the degree of counterion dissociation respectively. For the dimeric
The surface excess {Cmax/(mol  cm2)} is an effective measure
surfactants (i = j = 2 and jzs j ¼ jzc j ¼ 1), equation (1) becomes:
of adsorption at the air/water interface. The concentration of the
surfactant is always greater at the surface than in the bulk. The sur- 1
DGmic =ðkJ  mol Þ ¼ RTð1:5  aÞ ln X CMC  ðRT=2Þ ln 2: ð2Þ
face excess (Cmax) and the minimum area per molecule (Amin) were
calculated by using the Gibbs adsorption equation [38,39]. In table The second term in the right hand side of equation (2) is small with
1, it can be seen that the surface excess concentrations, Cmax, of the respect to the first one and negligible. The values of DGmic obtained
Gemini surfactants decrease with increasing temperature. This by using equation (2) are plotted in figure 1. The DGmic values of
behavior arises because thermal motions increase as the tempera- these surfactants do not change significantly with increases in tem-
ture increases. perature. The Gibbs free energy of micellization becomes more neg-
The surface excess concentration, Cmax, decreases with in- ative with increases in the hydrophobicity of a surfactant, which
creases in the spacer length because the surface area per surfactant means that for the studied surfactants micelle formation is most
molecule occupied by head groups is larger for larger spacers. facile for 14–2–14.
Due to hydrophobic interactions between the chains, the mag- By applying the Gibbs–Helmholtz equation to equation (2), the
nitude of Cmax is lower for longer hydrophobic chains, suggesting enthalpy of micellization can be obtained [18]. This method is an
that the air/water interface is close-packed and therefore that the easy one, as it only involves CMC and a determinations, but several
110 M. Hajy Alimohammadi et al. / J. Chem. Thermodynamics 44 (2012) 107–115

80

Thermodynamic parameters/kJmol -1
60 a
40

20

-20

-40

-60
295 300 305 310 315 320
T/K
80
Thermodynamic parameters/kJmol-1

60 b
40

20

-20

-40
295 300 305 310 315 320
T/K

60

c
Thermodynamic parameters/kJmol-1

40

20

-20

-40

-60
305 310 315 320 325 330 335
T/K
50
Thermodynamic parameters/kJmol-1

40
30
d
20
10
0
-10
-20 295 300 305 310 315 320
-30
-40
-50
295 300 305 310 315 320
T/K

FIGURE 1. Plot of thermodynamic parameters of micellization (- - - the Gibbs free energy of micellization ðDGmic Þ, — standard enthalpy ðDHmic Þ and    standard entropy
ðT DSmic Þ) of Gemini surfactants: (h) 12–2–12; (⁄) 12–4–12; (D) 14–2–14; (s) 14–4–14 versus temperatures. The thermodynamic parameters were calculated by using the
CMC values obtained from surface tension measurements and a values calculated according to the method of Frahm.

factors that affect DHmic , are not explicitly considered in data treat- accuracy. For example, the use of a small number of data points
ment by Van’t Hoff analysis [21,41]. In the van’t Hoff approach, usually leads to a linear dependence of DGmic =T on 1/T, hence to
DGmic =T versus 1/T is plotted for several temperatures. The quality a single value of DHmic for the temperature range studied. In this
of these plots depends on the number of data points and their work, this problem can be circumvented by fitting the curve with
M. Hajy Alimohammadi et al. / J. Chem. Thermodynamics 44 (2012) 107–115 111

60

40

20

ΔΗ˚mic/kJmol-1
0

-20

-40

-60

-80
-0.03 0.02 0.07 0.12 0.17 0.22 0.27
ΔS˚mic/kJmol-1K-1

FIGURE 2. Plot of standard enthalpy ðDHmic Þ versus standard entropy ðDSmic Þ, for Gemini surfactants: (h) 12–2–12; (⁄) 12–4–12; (D) 14–2–14; (s) 14–4–14.

TABLE 2
Parameters obtained in the presence of supporting electrolyte for Gemini Surfactants at their respective CMC values

Compound T/K Dma  1011/(m2  S1) Dmb  1011/(m2  S1) Dm  1011 =ðm2 S1 Þ Rh =Å Nagg kd/(dm3  mmol1) (Nagg)th Nagg

12–2–12 298 2.25 1.54 8.86 – – 0.042 – 48c


12–4–12 298 7.61 5.85 11.38 2.153 35 0.025 27 36d,55e
14–2–14 313 – – 9.08 – – 0.028
14–4–14 298 5.38 4.83 9.4 2.606 51 0.026 38
303 6.31 10.7 2.599 50 0.021
308 7.06 11.7 2.678 54 0.019
313 9.32 13.4 2.620 51 0.015
318 11.30 14.2 2.747 59 0.011
a
The diffusion coefficient values obtained from the PFG-NMR technique in the absence of supporting electrolyte (Ctot = 0.019 mol  dm3).
b
The diffusion coefficient values obtained from CV technique in the presence of supporting electrolyte (Ctot = 0.019 mol  dm3).
c
To calculate P, we calculated a0 values using surface tension plots and volume and length values were determined by Tanford equations (the values in parentheses) and
theoretical calculations. The volume and length values obtained by two methods are given in SM.
d
Data taken from reference [24].
e
Data taken from reference [13].

polynomial of higher order to get temperature-dependent DHmic. sation phenomenon has been found for a variety of processes of
More importantly, however, is that there is no provision in the small solutes [46], including those of surfactants [47–51].
van’t Hoff equation for factors that are important for micelle for- In general, such compensation phenomena can be described
mation of ionic surfactants, in particular, the dependence of micel- with a linear relationship of the form [34,52]:
lar geometry, aggregation number, surface-charge density, and
1
hydration of the head group on T [41]. On the other hand, the ef- DHmic =ðkJ  mol Þ ¼ DHmic þ T C DSmic : ð3Þ
fects of T on the above-mentioned micellar parameters are in-
cluded in the direct (i.e., calorimetric) determination of DHmic . An acceptable linear relationship was obtained in all compensation
The enthalpy of micellization, DHmic , is dependent on tempera- plots, as shown in figure 2; these lines have approximately the same
ture (figure 1). The temperature dependence of DHmic might be due slope but different intercepts at DSmic ¼ 0. Note that the intercept
to changes in aggregation number (size) and shape [42,43]. Posi- DHmic gives the enthalpy effect under the condition DSmic ¼ 0. An in-
tive values of DHmic , such as those observed at low temperatures, crease in the value of the intercept DHmic thus corresponds to a de-
are usually attributed to the release of structural water from the crease in the stability of the structure of the micelles. It was found
hydration layers around the hydrophobic parts of the surfactant that the intercept, DHmic , that characterizes the solute–solute inter-
molecule [44]. It has been suggested [45] that negative DHmic val- action is dependent on not only the hydrophobic chain length but
ues are evidence that London-dispersion interactions are a major also on the length of the spacer group. Our results indicate that
force in micellization. 14–2–14 has a larger negative value (DHmic ) than the other surfac-
The entropy contribution to the micellization process can be tants studied here. The micellar stabilities decrease in the order 14–
estimated from the calculated enthalpy and Gibbs free energy val- 2–14 > 14–4–14 > 12–2–12 > 12–4–12. The compensation tempera-
ues [21]. From Gibbs free energy relationship, any uncertainty tures range approximately 295 K to 304 K for all these surfactants.
introduced in the calculation of DHmic will be carried over to
DSmic [41].
It can be seen in figure 1that the absolute value of T DSmic for all 3.2. Intermicellar interactions
these surfactants decreases with increasing temperature, which
indicates that raising the temperature leads to a reduction in the From a structural point of view, the most important parameters
entropy or randomness of the system. The comparison of the of a micellar system are the mean micellar aggregation number
enthalpy and entropy terms (see figure 1) suggests the presence and the hydrodynamic radius, which were determined by using
of a mutual relationship in the form of the enthalpy–entropy self-diffusion coefficient measurements with PFG-NMR spectros-
compensation phenomenon [43]. The enthalpy–entropy compen- copy and CV techniques.
112 M. Hajy Alimohammadi et al. / J. Chem. Thermodynamics 44 (2012) 107–115

16
14
12

10-11Dm/m2.s -1
10
8
6
4
2
0
295 300 305 310 315 320
T/K

FIGURE 3. Self-diffusion coefficient, Dm as a function of temperature in the presence of various concentrations of 14–4–14, 2Br: (⁄) 0.010; (h) 0.015; (D) 0.020 mol  dm3.

0.03

0.025
-1
kd/dm 3mmol

0.02

0.015

0.01

0.005

0
295 300 305 310 315 320
T/K

FIGURE 4. Plot of intermicellar interaction parameter (kd) against temperature in micellar solutions (14–4–14).

1.8
1.7
1.6
1.5
1.4
U(r

1.3
1.2
1.1
1
0.0 1 0.01 2 0.01 4 0.01 6 0.01 8 0.0 2 0.02 2 0.024
kd/ dm3 mmol-1

FIGURE 5. Plot of the micellar interaction energy (U(r)) against intermicellar interaction parameter (kd) in micellar solutions.

The self-diffusion coefficient values obtained with PFG-NMR are not perturb the micelle and that its rates of entrance/exit into the
listed in table 2. The increases in the self-diffusion coefficient are aggregates with fast and reversible electron transfer are at least
mainly due to decreases in the size of the micelles. The low self-dif- comparable to those of the surfactant monomers. The electrochem-
fusion coefficient values of 12–2–12 and 14–2–14 indicate that the istry of ferrocene in non-aqueous, aqueous, and micellar environ-
presence of these Gemini surfactants with short spacers results in ments is described elsewhere [54–58].
the formation of aggregates that are less curved. It was found [13] The cyclic voltammograms for the one- electron oxidation of
in an electron microscopy experiment that a Gemini surfactant with ferrocene (equation (4)) in 12–4–12 at various scan rates for the
s = 2 forms rod-like micelles. Electron micrographs of 12–2–12 solu- potential range 0.0–0.5 V are given in figure SM 3.
tions have found long cylindrical micelles [15,21]. An investigation
Fc () Fcþ þ e : ð4Þ
with cryogenic transmission electron microscopy found that 12–s–
12 surfactants with 4 6 s 6 12 form spheroidal micelles [13]. In CV, the peak current ipa for a redox-active reversible system is
The CV measurements were used to determine the diffusion given by the Randles–Sevcik equation [59]:
coefficients of the micelles and the intermicellar interaction  1
parameters [53]. In the present study, ferrocene was chosen as nF mD 2
ipa ¼ 0:4463FACn ; ð5Þ
the electroactive probe; ferrocene can be used provided that it does RT
M. Hajy Alimohammadi et al. / J. Chem. Thermodynamics 44 (2012) 107–115 113

where n is the number of electrons involved in oxidation or reduc- supporting material). Note no significant variation in Rh with
tion, A is the area of the electrode (A was determined as the method increasing temperature, which might be due to an increase in the
described in SM), F is Faraday’s constant, R is the gas constant, T is aggregation number and a decrease in micellar solvation. Previ-
the absolute temperature, D is the diffusion coefficient of the elec- ously, Borse and co-workers found that the hydration of micelles
troactive (EC) probe, C is the concentration of the EC probe in the (hE) decreases in cationic Gemini surfactants with increasing tem-
solution, and v is the scan rate. For micellar systems involving an perature [66]. By assuming that there is no substantial change in
EC probe completely solubilized in the micelles, the D in equation micellar solvation, the aggregation numbers can be calculated at
(5) corresponds to the micelle diffusion coefficient Dm since the various temperatures. The values in table 2 demonstrate that the
probe diffuses with the micelles. Hence, from the slope of the ipa Nagg values of 14–4–14 remain predominantly constant with the
versus v1/2 plot, the Dm values were obtained for various micelle increase in temperature. A small variation in Nagg values of 14–
concentrations; the concentration of ferrocene was fixed at 4–14 are in line with the similar variation in the CMC values of this
1  103 mol  dm3. The Dm was found to decrease with decreasing surfactant. Consistent with the present work, the previous studies
spacer length for surfactants 12–s–12 and 14–s–14 (table 2), which have noted a small variation of Nagg with the increase in the tem-
is in good agreement with the PFG-NMR data. In CV measurements, perature for cationic Gemini surfactants [18,21].
the addition of supporting electrolyte causes a decrease in Dm or The values of the intermicellar interaction coefficient (kd) were
micellar growth due to reduction electrostatic repulsion between calculated from the slopes of plots of Dm versus the surfactant con-
head groups (table 2). These results are in agreement with previous centration (see figure SM4), as described by linear interaction the-
findings [55]. ory [60]. In table 2, it can be seen that the kd values are influenced
At concentrations above, but not far from, the CMC micelles by increases in the lengths of the tail and the spacer. The intermi-
mutually interact; consequently, the calculated values of diffusion cellar interaction parameter decreases with increases in the chain
coefficients are lower than that at the CMC. As attenuation of the lengths of the Gemini surfactants. This behavior is mainly due to
mass-transport parameters is a linear function of surfactant con- the growth of micelles and decreases in the micellar surface charge
centration therefore linear interaction theory (see SM) was used density [18,66]. On the other hand, an unexpected increase in kd
for calculation of intermicellar interaction parameters [60,61]. was found for decreasing spacer length. This increase is due to
Consequently, extrapolation to infinite dilution (to the CMC) yields the change from sphere to rod micellar morphology. Small angle
mass-transport parameters that are independent of interparticle neutron scattering and cryo-TEM studies of the m–s–m Gemini sur-
interactions factants revealed the formation of less curved aggregates when the
spacer is short enough [21]. Cryo-TEM showed that the 12–2–12
Dm ¼ D0m ½1  kd ðC s  CMCÞ; ð6Þ micelles are threadlike, whereas those of 12–4–12 are spheroidal
[13,21].
where kd is the intermicellar interaction parameter, D0m is the self- In order to examine the effects of varying the temperature on
diffusion coefficient in the absence of micellar interaction, CMC is intermicellar interactions, voltammetric experiments were carried
the critical micelle concentration, and Cs is the surfactant concen- out for 14–4–14 over the temperature range 298 K to 318 K. Figure
tration. The representative profiles of Dm versus (Cs-CMC) for 12– 3 shows the variation of Dm with temperature at three concentra-
2–12, 12–4–12, 14–4–14, and 14–2–14 (in the presence of support- tions. As expected, the Dm values increase with increasing temper-
ing electrolyte) are shown in figure SM 4. The values of D0m or DCMC ature. The kd values were obtained with linear interaction theory
were calculated from the dependence of Dm on surfactant concen- and are plotted as a function of temperature in figure 4. It is obvi-
tration. The results are shown in table 2. These data were used to ous that kd decreases linearly with increasing temperature. Two
calculate the micellar hydrodynamic radius by using the Stokes– factors control the value of kd in colloidal systems: inter-particle
Einstein equation [62–65]. attractive or repulsive interactions. According to our results,
For s = 2, spherical micelles cannot form and the Stokes–Ein- decreasing kd with temperature suggests a decrease in repulsive
stein equation is not valid. For a Gemini surfactant with a longer interactions due to a reduction in surface charge density and an in-
spacer, the surface area per surfactant molecule occupied by head crease attractive interactions. This temperature effect is consistent
groups is larger than for a Gemini surfactant with a shorter spacer, with the results of a previous study by Borse et al. [66]. They re-
i.e., there is a smaller packing parameter, which facilitates the for- ported that the equilibrium distances (d) between the head groups
mation of spherical rather than worm-like micelles. According to of surfactant 12–4–12 are 0.742 and 0.771 nm at T = 303 K and
the experimental data [21], 12–4–12 and 14–4–14 are likely to 323 K, respectively. An increase in the attractive interaction or de-
form spherical micelles. crease in repulsive interaction leads to micellar growth. The small
The hydrodynamic radius values are shown in table 2. The Rh magnitude of the increase in the micellar size or Nagg (table 2) sug-
values can be used to compute the aggregation number of the mi- gests that the increase in attractive interaction or decrease in
celles. By assuming spherical micelles and neglecting their intrinsic repulsive interaction has only a small effect on micelle size, but
instability, the aggregation number can be computed. The Nagg val- in our work the structure of Gemini surfactant is the dominant fac-
ues are collected in table 2, along with the theoretically calculated tor for determination of the micelle size. While, the results of
aggregation numbers, Nagg,theor, and those obtained with other Charlton et al. showed a decrease in interaction parameter with
techniques. The CV-based Nagg values are in excellent agreement temperature due to increase in attractive interaction, then the in-
with the values obtained using other techniques [13,29]. The Nagg crease in the attractive interaction leads to micellar growth for Tri-
values are dependent on the chain length of the surfactant hydro- tonX-100 [56]. Further, the intermicellar interaction parameters
phobic tail in the presence of electrolyte. The increases in Nagg with for the monomeric form of these surfactants such as cetyltrimeth-
increases in the length of the R group are due to increases in the ylammonium bromide (CTAB) [67], cetyltrimethylammonium
packing factor, because such increases affect the volume of the chloride (CTACl) [68] and sodium dodecyl benzene sulfonate
hydrophobic group more than its length (for the same head group). (SDBS) [69] increase with increasing temperature.
In addition, the actual diffusion coefficient, D0m , hydrodynamic ra- To examine further the observed behavior, the intermicellar
dius, Rh, and aggregation numbers, Nagg of the 14–4–14 micelles interaction energy (Coulombic potential) for two identical spheri-
are shown for various temperatures in table 2. The results indicate cal macro-ions of diameter r was calculated [70]. A plot of U(r) ver-
an increase in D0m with increasing temperature. At least part of this sus kd is shown in figure 5. It is evident that the intermicellar
change in D0m is due to the decrease in solvent viscosity (see the interaction parameter is a function of the Coulombic interaction
114 M. Hajy Alimohammadi et al. / J. Chem. Thermodynamics 44 (2012) 107–115

1
potential. The interaction parameter decreases gradually with DHmic =ðkJ  mol Þ ¼ RT 2 ð1:5  aÞð@ ln X CMC =@TÞp þ
increases in temperature as the Coulombic interaction (U(r)/kT)
RT 2 LnX CMC ð@ a=@TÞ: ð11Þ
varies between 1.74 and 1.13.
To evaluate the enthalpy of micelle formation, the CMC and a values
4. Conclusion were first correlated by using a polynomial equation [18]:
3
In our surface tension and conductometry study of ‘nX CMC ðTÞ ¼ a þ bT þ CT 2 þ dT ; ð12Þ
CmH2m+1N(CH3)2(CH2)S (CH3)2 N CmH2m+1,2Br with m = 12, 14
and s = 2, 4 over the temperature range 298 K to 323 K, the varia- a ¼ a0 þ b0 T þ c0 T 2 ; ð13Þ
tions of CMC and a with temperature have been obtained. The
where constants a, b, c, d, a’, b’, and c’ were determined by using
CMC values pass through a minimum as the temperature is in-
least-squares regression analysis.
creased from 298 K to 323 K. The a values increase slightly with
The enthalpy of micelle formation was then calculated numer-
temperature. This study also enabled us to obtain the thermody-
ically by substituting equations (13) and (14) into equation
namic parameters of micellization. The Gibbs free energy of micel-
(12).The entropy contribution to the micellization process can be
lization has a weak dependence on temperature for these Gemini
estimated from the calculated enthalpy and Gibbs free energy val-
surfactants. A clear linear relation (enthalpy–entropy compensa-
ues as
tion relation) was found between DH0m and DS0m for these Gemini
surfactants. T DSmic ¼ DHmic  DGmic : ð14Þ
By performing CV measurements, several important micellar
The micellar hydrodynamic radius was calculated by using the
parameters were estimated and the intermicellar interactions were
Stokes–Einstein equation [62–65]:
analyzed. The effects of varying the temperature can be interpreted
in terms of decreasing electrostatic repulsions of head groups and kT
Rh ¼ ; ð15Þ
increasing attractive interactions of spacers. This study has shown 6pgD0m
that CV is a simple and sensitive technique for probing the struc-
tural changes of micelles and in particular their dependence on where k is the Boltzmann constant, T is the absolute temperature,
the spacer and chain lengths of the hydrophobic tails of Gemini and g is the viscosity of the solution, which can be approximated
surfactants, and for obtaining information about intermicellar as that of water.
interactions in aqueous solutions. The aggregation number can be computed with the following
equation:

5. Calculations ð4=3ÞpR3h
Nagg ¼ ; ð16Þ
½V A þ nh;A V H2 O þ bðV con þ nh;con V H2 O Þ
As can be seen from figure SM2, plots of solution conductivity
where VA, Vcon, and V H2 O are the molecular volumes of the Gemini
versus surfactant concentration are composed of two straight lines
surfactants, the counter ion, and H2O, and nh,A and nh,con are the
intersecting at the CMC. We employed the method of Frahm and
hydration numbers of the head groups of the surfactants and coun-
that of Evans to calculate the a value.
ter ions [71]. Finally, b is defined as the degree of attachment.The
The method of Frahm is given by [35]
intermicellar interaction energy (Coulombic potential) for two iden-
a ¼ s2 =s1 ; ð7Þ tical spherical macro-ions of diameter r was calculated [70]

where s1 and s2 are the slopes k plots below and above the CMC, UðrÞ ¼ pee0 r2 w20 exp½jðr  rÞ=r; ð17Þ
respectively. where r is the micellar center-to-center distance, e is the dielectric
The method of Evans for calculating a is given by [36]: constant of the medium (H2O), e0 is the permittivity of free space, k
is the Debye–Huckle inverse screening length as determined from
1000s2 ¼ ða2 =N2=3
agg Þð1000s1  KBr Þ þ aKBr ;
  ð8Þ the ionic strength of the solution, and r ¼ 2R0h , where R0h is the
micellar hydrodynamic radius. R0h can be obtained from D0m , which
where s1, s2 are those defined above; the KBr is the equivalent con-
can be determined with the Stokes–Einstein relation (equation
ductivity of Br at infinite dilution.
(16)). The value for r can be calculated from the micellar volume
The surface excess (Cmax) and the minimum area per molecule
fraction (u) (obtained from R0h ) as follows:
(Amin) were calculated by using the Gibbs adsorption equation
[38,39], r ¼ 2R0h þ l; ð18Þ

Cmax =mol  m2 ¼ ð1=2:303RTÞ½dc=d log CT;P ð9Þ 3


l ¼ ½8p=21=3 uðR0h Þ: ð19Þ
and

1 1018 Appendix A. Supplementary material


Amin =nm2  molecule ¼ ; ð10Þ
NA Cmax
Supplementary data associated with this article can be found, in
the online version, at doi:10.1016/j.jct.2011.08.007.
where dc/d log C is the maximum slope in each case. The R, T, C,
and N are the gas constant, absolute temperature, concentration,
References
and Avogadro’s constant, respectively. The slope of the tangent at
a given concentration of the c versus log C plot was used to calcu- [1] F.M. Menger, C.A. Littau, J. Am. Chem. Soc. 115 (1993) 10083–10090.
late C by using curve fitting to a polynomial equation of the form [2] F.M. Menger, C.A. Littau, J. Am. Chem. Soc. 113 (1991) 1451–1452.
y = ax2 + bx + c.By applying the Gibbs–Helmholtz equation to [3] M.J. Rosen, Chemtech 23 (1993) 30–33.
[4] F. Devinsky, I. Lacko, T. Imam, J. Colloid Interf. Sci. 143 (1991) 336–342.
equation (2), the enthalpy of micellization can be obtained from [5] G. Bai, J. Wang, H. Yan, Z. Li, R.K. Thomas, J. Phys. Chem. B105 (2001) 3105–
[18]: 3108.
M. Hajy Alimohammadi et al. / J. Chem. Thermodynamics 44 (2012) 107–115 115

[6] R. Zana, Dimeric (Gemini) Surfactants, in: K. Esumi, M. Ueno (Eds.), Structure– [38] M.J. Rosen, Surfactants and Interfacial Phenomena, third ed., John Wiley and
Performance Relationships in Surfactants, Dekker, New York, 1997, p. 255. Sons, New Jersey, 2004.
[7] M.J. Rosen, J.H. Mathias, L. Davenport, Langmuir 15 (1999) 7340–7346. [39] M.J. Rosen, Langmuir 7 (1991) 885–888.
[8] M. Dreja, W. Pyckhout-Hintzen, H. Mays, B. Tieke, Langmuir 15 (1999) 391– [40] R. Zana, Langmuir 12 (1996) 1208–1211.
399. [41] P. Galgano, O. El Seoud, J. Colloid Interf. Sci. 345 (2010) 1–11.
[9] T.J. Gao, M.J. Rosen, J. Am. Oil Chem. Soc. 71 (1994) 771–776. [42] M. Hisatomi, M. Abe, N. Yoshino, S. Lee, S. Nagadome, G. Sugihara, Langmuir 6
[10] F.L. Duivenvoorde, M.C. Feiters, S.J. Van der Gaast, J.B.F.N. Engberts, Langmuir (1999) 1515–1521.
13 (1997) 3737–3743. [43] G.C. Kresheck, W.A. Hargraves, J. Colloid Interf. Sci. 48 (1974) 481.
[11] M. Blanzat, E. Perez, I. Rico-Lattes, D. Prome, J.C. Prome, A. Lattes, Langmuir 15 [44] G.C. Kresheck, in: F. Franks (Ed.), Water: A Comprehensive Treatise, Plenum,
(1999) 6163–6169. New York, 1975.
[12] Y.-P. Zhu, A. Masuyama, Y. Kirito, M. Okahara, M.J. Rosen, J. Am. Oil chem. Soc. [45] D. Rubingh, in: K.L. Mittal (Ed.), Solution Chemistry of Surfactants, Plenum,
69 (1992) 626–632. New York, 1979.
[13] D. Danino, Y. Talmon, R. Zana, Langmuir 11 (1995) 1448–1456. [46] C.V. Krishnan, H.L. Friedman, J. Solution Chem. 2 (1973) 37–51.
[14] R. Oda, I. Huc, D. Danino, Y. Talmon, Langmuir 16 (2000) 9759–9769. [47] T. Okano, T. Tamura, T.-Y. Nakano, S.-I. Ueda, S. Lee, G. Sugihara, Langmuir 16
[15] A. Bernheim-Groswasser, R. Zana, Y. Talmon, J. Phys. Chem. B 104 (2000) (2000) 3777–3783.
4005–4009. [48] J.A. Molina-Bolivar, J. Aguiar, J.M. Peula-Garcia, C.C. Ruiz, J. Phys. Chem. B 108
[16] X. Huang, Y.C. Han, Y.X. Wang, Y. Wang, J. Phys. Chem. B 111 (2007) 12439– (2004) 12813–12820.
12446. [49] J.L. López-Fontán, A. González-Pérez, J. Costa, J. Ruso, G. Prieto, P. Schulz, F.
[17] R. Zana, J. Colloid Interf. Sci. 248 (2002) 203–220. Sarmiento, J. Colloid Interf. Sci. 297 (2006) 10–21.
[18] M.S. Bakshi, A. Kaura, R.K. Mahajan, Colloids and Surfaces A: Physicochemical [50] D. Das, K. Ismail, J. Colloid Interf. Sci. 327 (2008) 198–203.
and Engineering Aspects 262 (2005) 168–174. [51] J. Das, K. Ismail, J. Colloid Interf. Sci. 337 (2009) 227–233.
[19] A. Bendjeriou, G. Derrien, P. Hartmann, C. Charnay, S. Partyka, Thermochim. [52] R. Lumry, S. Rajender, Biopolymers 9 (1970) 1125.
Acta 434 (2005) 165–170. [53] B. Geetha, A. Mandal, Langmuir 13 (1997) 2410–2413.
[20] G. Bai, H. Yan, R.K. Thomas, Langmuir 17 (2001) 4501–4504. [54] I.D. Charlton, A.P. Doherty, J. Phys. Chem. B 104 (2000) 8327–8332.
[21] R. Zana, J. Xia, Gemini Surfactants: Synthesis, Interfacial and Solution Phase [55] T.L. Ferreira, B.M. Sato, O.A. El Seoud, M. Bertotti, J. Phys. Chem. B 114 (2009)
Behaviour, and Application, Marcel Dekker, New York, 1998. 857–862.
[22] L. Grosmaire, M. Chorro, C. Chorro, S. Partyka, S. Lagerge, Thermochim. Acta [56] I.D. Charlton, A.P. Doherty, Colloids and Surfaces A: Physicochemical and
379 (2001) 255–260. Engineering Aspects 182 (2001) 305–310.
[23] M.S. Borse, S. Devi, Adv. Colloid Interf. Sci. 123 (2006) 387–399. [57] I.D. Charlton, A.P. Doherty, Langmuir 15 (1999) 5251–5256.
[24] L. Grosmaire, M. Chorro, C. Chorro, S. Partyka, R. Zana, J. Colloid Interf. Sci. 246 [58] Z.X. Chen, S.P. Deng, X.K. Li, J. Colloid Interf. Sci. 318 (2008) 389–396.
(2002) 175–181. [59] K. Chokshi, S. Qutubuddin, A. Hussam, J. Colloid Interf. Sci. 129 (1989) 315–
[25] R. Oda, S.J. Candau, I. Huc, Chem. Commun. (1997) 2105–2106. 326.
[26] S. Lagerge, A. Kamyshny, S. Magdassi, S. Partyka, J. Therm. Anal. Calorim. 71 [60] E. Dickinson, Chapter 2. Annual Reports on the Progress of Chemistry Section
(2003) 291–310. C: Physical Chemistry 80 (1983) 3–37.
[27] L. Chen, Y. Li, H. Xie, J. Dispers. Sci. Tech. 29 (2008) 1098–1102. [61] A.B. Mandal, Langmuir 9 (1993) 1932–1933.
[28] Y. Jiang, H. Chen, X.-H. Cui, S.-Z. Mao, M.-L. Liu, P.-Y. Luo, Y.-R. Du, Langmuir 24 [62] Z. Yang, J. Zhao, Y. Xie, Y. Bai, Z. Du, Z. Phys. Chem. 217 (2003) 1109–1118.
(2008) 3118–3121. [63] H. Gharibi, B. Sohrabi, S. Javadian, M. Hashemianzadeh, Colloids Surf. A:
[29] A.R. Tehrani Bagha, H. Bahrami, B. Movassagh, M. Arami, F.M. Menger, Dyes Physicochem. Eng. Aspects 244 (2004) 187–196.
Pigments 72 (2007) 331–338. [64] A. Einstein, R. Fürth, Investigations on the Theory of the Brownian Movement,
[30] M.A. Rodriguez, M. Munos, M.M. Graciani, M.S.F. Pachon, M.L. Moya, Colloids Dover Publications, New York, 1956.
and Surfaces A: Physicochemical and Engineering Aspects 298 (2007) 177– [65] J.N. Israelachvili, D.J. Mitchell, B.W. Ninham, J. Chem. Soc., Faraday Trans. 2
185. (72) (1976) 1525–1568.
[31] C.C. Ruiz, J.A. Molina-Bolivar, J. Aguiar, G. MacIsaac, S. Moroze, R. Palepu, [66] M. Borse, V. Sharma, V.K. Aswal, P.S. Goyal, S. Devi, J. Colloid Interf. Sci. 284
Langmuir 17 (2001) 6831–6840. (2005) 282–288.
[32] D.H. Wu, A.D. Chen, C.S. Johnson, J. Magn. Reson. Ser. A 115 (1995) 260–264. [67] R.B. Dorshow, J. Briggs, C.A. Bunton, D.F. Nicoli, J. Phys. Chem. 86 (1982) 2388–
[33] Z. Yang, Y. Lu, J. Zhao, Q. Gong, X. Yin, J. Phys. Chem. B 108 (2004) 7523–7527. 2395.
[34] L.-J. Chen, S.-Y. Lin, C.-C. Huang, J. Phys. Chem. B 102 (1998) 4350–4356. [68] R.B. Dorshow, C.A. Bunton, D.F. Nicoli, J. Phys. Chem. 87 (1983) 1409–1416.
[35] J. Frahm, S. Diekmann, A. Haase, Ber. Bunsenge. Phys. Chem. 84 (1980) 566– [69] D.C.H. Cheng, E. Gulari, J. Colloid Interf. Sci. 90 (2004) 410.
571. [70] J.B. Hayter, J. Penfold, Mol. Phys. 42 (1981) 109–118.
[36] H. Evans, J. Chem. Soc 78 (1956) 579–586. [71] J. Zhao, S. Deng, J. Liu, C. Lin, O. Zheng, J. Colloid Interf. Sci. 311 (2007) 237–242.
[37] Y. Moroi, in: Micelles: Theoretical and Applied Aspects, Plenum, New York,
1992, pp. 41–90.
JCT-11-213

You might also like