Você está na página 1de 8

Colloids and Surfaces A: Physicochem. Eng.

Aspects 235 (2004) 121–128

A model on the temperature dependence of


critical micelle concentration
Hong-Un Kim, Kyung-Hee Lim∗
Department of Chemical Engineering, Chung-Ang University, Seoul 156-756, South Korea

Received 18 April 2003; accepted 23 December 2003

Abstract

An equation, which describes the dependence of critical micelle concentration (XCMC ) with temperature, has been derived on the basis of
G0 = −RT ln K, linear behavior of the enthalpy of micellization with temperature, and compensation phenomena in which the enthalpy
and the entropy of micellization change linearly with each other. The new equation has yielded excellent fitting results of XCMC (T) for
various surfactant systems. More interestingly, it yields d = 2, irrespective of surfactant system, in the power-law description of XCMC (T),
∗ ∗ ∗
|XCMC − XCMC | = const|T − T ∗ |d with the minimum CMC, XCMC , and the temperature, T∗ at XCMC . The value of d = 2 is confirmed from
the fits to reported literature data.
© 2004 Elsevier B.V. All rights reserved.

Keywords: Critical micelle concentration (CMC); Temperature dependence; Mass action law model; Compensation phenomena

1. Introduction When CMC is measured at various temperatures, the


thermodynamic potentials of micellization can be deter-
Micelles are one of the colloidal association structures, mined [3–10]. The standard Gibbs free energy of micelliza-
in which vesicles, microemulsions, and liquid crystals are tion per mole of surfactant or amphiphile, G0mczn , may
included. Micelles are formed above a narrow concentration be obtained from the relation G0mczn = −RT ln K/n with
range of surfactant and this range is called the critical micelle K and n being the equilibrium constant and association
concentration (CMC). number, respectively, for the micellization. K is expressed
Micellization is affected by various factors including rigorously in terms of the activities and can be approxi-
temperature, pressure, ionic strength, pH, surfactant species mated in terms of concentrations of the constituents, one of
such as hydrophobic volume, chain length, head group area, which is the concentration XCMC at the CMC. Then, the en-
etc. For ionics and amphoterics, micellization is affected thalpy of micellization, Hmczn0 , is obtained from G0mczn
by temperature as the hydrophobic and head group interac- and the Gibbs–Helmholtz relation, ∂(G0mczn /T)/∂T =
tions change with temperature. Accordingly, CMC versus −Hmczn0 /T 2 and the entropy of the micellization, Smczn
0 ,
temperature studies have been performed to obtain infor- from Smczn = (Hmczn − Gmczn )/T .
0 0 0
mation on these interactions [1]. For nonionic surfactants, In determining Hmczn 0 from G0mczn and the Gibbs–
the CMC decreases with an increasing temperature due to Helmholtz relation, the temperature dependence of G0mczn
an increase in hydrophobicity caused by the destruction of should be known so that its partial derivatives with respect
hydrogen bonds between water molecules and hydrophilic to T are calculated. This signifies that XCMC should be de-
groups [2] and the log CMC versus 1/T plot is nearly linear scribed as a function of temperature and this has been done
[2]. Meanwhile, for ionic surfactants the CMC decreases to by expressing XCMC as polynomials of temperature. How-
a minimum value and then increases, displaying a U-shaped ever, such description of XCMC (T) is not theory-based and
behavior [3,4]. The minimum is usually characterized by the therefore a theory-based and rigorous equation is desired.
minimum CMC, XCMC ∗ , and the temperature, T∗ at XCMC ∗ . In this article a new equation of XCMC (T) is derived on
the basis of G0mczn = −RT ln K/n, linear behavior of
∗Corresponding author. Tel.: +82-2-820-5275; fax: +82-2-826-3574. Hmczn
0 (T) [11–14], and compensation phenomena [15–21]
E-mail address: khlim@cau.ac.kr (K.-H. Lim). in which Hmczn 0 is linearly proportional to Smczn
0 . The

0927-7757/$ – see front matter © 2004 Elsevier B.V. All rights reserved.
doi:10.1016/j.colsurfa.2003.12.019
122 H.-U. Kim, K.-H. Lim / Colloids and Surfaces A: Physicochem. Eng. Aspects 235 (2004) 121–128

new equation, which has three fitting parameters, follows nS + mG = Sn Gm z = M z (1)


surfactant counterion micelle
measured CMC data very closely and yield better fitting re-
sults than the second-order polynomial which has also three where z is the charge or the valence of the micelles. The
fitting parameters. equilibrium constant K for the reaction is
A rigorous equation of XCMC (T) is required essentially  
in the power-law expression of CMC, that is, |XCMC − aM XM γM
K = n m = n m Kγ , Kγ ≡ n m (2)

XCMC | = const|T − T ∗ |d . In such expression XCMC ∗ , and T∗ a S aG X S XG γS γG
have been found first by fitting a polynomial XCMC (T) to in which ai is the activity of the species i. Use is made
measured CMC(T) and then the exponent d is found. How- in Eq. (2) of the relation ai = γi Xi with γ and X being
ever, recently Kang et al. [10] have found that the accurate the activity coefficient and mole fraction, respectively. If we
determination of XCMC ∗ and T∗ is a prerequisite to obtain an assume that the activities are replaced by the mole fractions,
accurate and precise exponent for n, in support of the pre- then Kγ = 1 and therefore Eq. (2) becomes
vious finding that a small variation of 1% or so in XCMC ∗ or

T has a pronounced effect on d [10]. In their study, three XM
K= m. (3)
polynomials of 4th-, 5th-, and 6th- order were used to fit the XSn XG
measured CMC(T) of the cationic octadecyl trimethyl am-
For the monomer, aS = 1 amounts to assuming that the
monium chloride (OTAC) and anionic ammonium dodecyl
departure from ideal behavior is only the aggregation pro-
sulfate (ADS) surfactants. The polynomials resulted in vir-
∗ cess. In principle, it can be removed by estimating activity
tually the same XCMC (their differences were 1.2% at most)
∗ corrections from solution theory. However, it is quite prob-
and T with only a small uncertainty (4.6% at most). How-
lematic to assume ideal behavior for the micelles (i.e., γM =
ever, quite different d’s were identified. With the 4th and
1) because of the large size difference between monomers
5th order polynomials, the same T∗ s were obtained and the
∗ and micelles [24]. The micelles will also interact strongly
XCMC s only differed by 1.8%, but the d’s differed by 8%.
∗ and for ionics the interaction will become very significant
With the 5th and 6th order polynomials XCMC s were almost
∗ as soon as the mean separation is less than about eight to ten
the same and the T s were different by 9%, whereas the
times double layer thickness that occurs at surfactant con-
d’s were different by 32%. These results imply that poly-
centrations not far above the CMC. It must also be noted
nomials of XCMC (T) cannot provide an accurate power-law
that, when ionic micelles are formed, there is a strong ten-
description of CMC. In addition, the exponent d appears to
dency for the counterions to be associated closely with the
depend on surfactant system. For example, La Mesa found
head groups, because of the high electrical potential in that
that d = 1.73 ± 0.03 for nine ionic surfactants [22] and
region. This is another source of nonideality. Notwithstand-
Stasiuk and Schramm [23] obtained d = 3.54 for three
ing the problems attributed to the assumption ai = 1, Eq. (3)
commercial ionic surfactants and d = 5.80 for two com-
still represents micellization quite well.
mercial amphoteric surfactants. In contrast, Kang et al. [10]
The standard Gibbs free energy change, G0 , for the mi-
found d = 1.12 ± 0.15 for OTAC and d = 1.05 ± 0.20
cellization of Eq. (1) can be obtained from the well-known
for ADS.
thermodynamic result G0 = −RT ln K. From these equa-
The new equation of XCMC (T) presented in this article has
∗ tions one obtains.
been also used in determining XCMC , T∗ , and d. It yields one
∗ ∗
pair of (XCMC , T ) unequivocally and d = 2, irrespective of G0 = −RT ln K = −RT(ln XM − n ln XS − m ln XG ).
surfactant system. When it is tested for surfactant systems
of various kinds, it yields excellent results for the power-law (4)
description as well as for XCMC (T). Here G0 is also expressed as
G0 = µ0M − nµ0S − mµ0G (5)
2. Gibbs free energy of micellization and critical
in which µ0i is the standard chemical potential of species i. If
micelle concentration (CMC)
we define G0mczn as the standard Gibbs free energy change
of micellization per mole of amphiphile, i.e., G0mczn =
Thermodynamics of micellization has been often de-
G0 /n, then
scribed by the models of mass action law and phase sepa-
ration. Although the former is more complicated than the 1 0 m
G0mczn = µ − µ0S − µ0G (6)
latter, it sheds more light on quantitative understanding of n M n
micellization. Among the mass action law models we take and
the closed association model, which provides the essence
of micellization without loss of generality. In this model RT
G0mczn = − ln K
it is assumed that monodisperse micelles comprised of n n 
surfactant molecules are found, which are formed via the 1 m
= −RT ln XM − ln XS − ln XG (7)
following reaction: n n
H.-U. Kim, K.-H. Lim / Colloids and Surfaces A: Physicochem. Eng. Aspects 235 (2004) 121–128 123

Introducing the ratio σ = XM /XCMC as Tanford did [25] vS XSt = XS + nXM = XS + nKXn+mS
and recognizing that, at the CMC, XS = XCMC − XM and n+m−2
that XG ≈ XS (= XCMC ) when the ionic surfactant SvS GvG = XS + XS
(n + m)(2n + 2m − 1)
is like a symmetric electrolyte, i.e., νS = νG , we obtain
(n + m − 1)(n + m + 1)
  = XS (15)
G0mczn m 1 1 (n + m)(2n + 2m − 1)
= 1+ − ln XCMC − ln σ + ln(1 − σ).
RT n n n At the CMC, XSt = XCMC and therefore
(8)
1 (n + m)(2n + 2m − 1)
XS = XCMC (16)
The choice of σ has only a small effect on the free energy vS (n + m − 1)(n + m + 1)
of micellization because it ranges usually from 0.01 to 0.10.
and its substitution to Eq. (14) yields
Moreover, for large n (for example, n ≥ 50) the 1/n terms
and the last term in the right-hand side are negligible. Hence, G0mczn
in this case Eq. (8) is reduced to
RT   
G0mczn m 1
= (1 + β) ln XCMC (9) = 1+ − ln XCMC
RT n n
 
where β≡m/n is the degree of counterion binding. This is 1 n(n + m)(2n + 2m − 1)
+ ln
the simpler relationship between G0mczn and XCMC from n n+m−2
  
which G0mczn is estimated with XCMC measured. m 1 1 (n + m)(2n + 2m − 1)
+ 1+ − ln
The last two terms of Eq. (8) may be elaborated [26] n n vS (n + m − 1)(n + m + 1)
using the fact that physical properties of surfactant solutions (17)
change abruptly at the CMC. For a micellar system, the
number of degrees of freedom is three. Accordingly, the total For n > 50, the third term in the first bracket and the quanti-
surfactant concentration can determine the concentrations of ties in the second and third brackets become negligible and
every chemical species at constant temperature and pressure. Eq. (9) is recovered [29].
On the basis of one mole of the total concentration the mass Eq. (17) is also employed for the micellization of nonionic
balances for counterions and surfactant ions yield. amphiphiles. In this case m = 0 or β = 0 and therefore the
equation is reduced to
vS XSt = XS + nXM (10)
    
G0mczn 1 1 n2 (2n − 1)
vG XSt = XG + mXM (11) = 1− ln XCMC + ln
RT n n n−2
  
where vs and vG are the valences of the surfactant ion and 1 1 n(2n − 1)
counterion, respectively, of the surfactant SvS GvG and XSt is + 1− ln (18)
n vS n 2 − 1
the fraction of total surfactant. The CMC can be identified
by the Philip’s criterion [28], d3 B/dXSt3 = 0. Here B is a Dropping 1/n terms in Eqs. (17) and (18) makes them re-
property of the micellar solution, which may be represented duced further to Eqs. (19a) and (19b).
by 
G0mczn m 1
B = bS XS + bG XG + bM XM (12) = 1+ ln XCMC + ln 2n(n + m)
RT n n
 m 2
where bi is the contribution factor of each species and is + 1+ ln (for ionics) (19a)
related to the partial molar quantity of respective species. n vS
After substantial mathematical manipulations [24,26,27]
for d3 B/dXSt3 = 0, one obtains
G0mczn 1 2
= ln XCMC + ln 2n2 + ln (for nonionics)
n(n + m)(2m + 2n − 1) n+m−1 RT n vs
−1
K = XS (13) (19b)
n+m−2
and from this equation along with G0mczn /RT = −ln K/n For ionics Eqs. (9) and (19a) are good approximates to
Eq. (17). Likewise, for nonionics Eq. (9) with β = 0 and
 
G0mczn m 1 Eq. (19b) are good approximates to Eq. (18). How good
= 1+ − ln XS the approximates are can be seen in Table 1 which contain
RT n n
1 n(n + m)(2n + 2m − 1) sample calculations for the nonionics.
+ ln . (14) When n is as small as 10, the error for Eq. (9) is large in
n n+m−2
absolute value (−22 %) but is small (+0.4%) for Eq. (19b).
For the surfactant with vS = vG , XG = XS , and along with Hence, Eq. (19b) can be as good as Eq. (17) within 0.5%
Eqs. (3) and (13) Eq. (10) leads to error. For Eq. (9), the larger the n, the smaller the error; the
124 H.-U. Kim, K.-H. Lim / Colloids and Surfaces A: Physicochem. Eng. Aspects 235 (2004) 121–128

Table 1
Errors in the values of G0mczn /RT calculated from approximate equations
for nonionic surfactants
n 10 20 60 100 500

CMC (mM) 2.2 3.7 5.4 5.6 6.5


Eq. (18) −4.9775 −5.0030 −4.9938 −5.0388 −5.0006
Eq. (9) with β = 0 −6.1193 −10.127 −5.2276 −5.1807 −5.0353
Eq. (19b) −4.9775 −4.9852 −4.9863 −5.0341 −4.9996

errors by Eq. (9) are 11.6, 4.6, 3.0, and 0.72% for n = 20,
60, 100, and 500, respectively. Hence, for n > 50 which is
quite usual, Eq. (9) is as good as the rigorous one, Eq. (18),
within 5% error. This may be understood easily by exam-
ining the quantities in the second and third brackets. They
decrease with increasing n. Besides, they contribute much
less to G0mczn /RT than the quantity in the first bracket.
Table 2 shows the relative magnitude of the quantities in
the brackets. The quantity in the second bracket contributes
within 10% to G0mczn /RT at small n and its contribution
decreases with increasing n. The quantity in the third bracket
contributes less than 1% even at small n.
Fig. 1. Calorimetric enthalpy change with temperature of micelle formation
Eqs. (19a) and (19b) with νs = 1, presented in [26], have in various aqueous surfactant solutions. (a) Dimethyldecylphosphine oxide
been used successfully for nonionics and for n
1, which [11] (䉬); dimethyldodecylphosphine oxide (䉫); sodium dodecyl sulfate
may be accounted for by the above arguments. (䊉); sodium dodecoyl sarcosinate (䊊); sodium octyl sulfate (䉱); sodium
decyl sulfate ( ); dodecylpyridinium bromide (䊏); dodecylpyridinium
chloride (䊐); and (b) [12] for perflorinated ionic surfactants, Na-PFPE
(䉫); H-PFPE (䊊); Na-PFO (䊐); H-PFO ( ).
3. Heat of micellization and its temperature
dependence
Since ∂Hmczn
0 /∂T = Cp,mczn
0 , the slope AH should be
Heat of micellization or enthalpy of micellization, equivalent to Cp,mczn , the heat capacity change of micel-
0
Hmczn
0 , has been determined in two ways; one is direct lization. The linearity between Hmczn
0 and T indicates that
measurement with a calorimeter and the other is calculation Cp,mczn is constant at the observed temperature intervals
0
from Hmczn
0 and the Gibbs–Helmholtz relation.
and that, as shown in Fig. 1a and b, Cp,mczn
0 is positive
for hydrocarbon surfactants and negative for perfluorinated
3.1. Direct method
ionic surfactants. If Hmczn
0 = 0 at the (reference) temper-
ature TH=0 , the linear behavior of Hmczn
0 with T can be
It has been observed that experimentally measured
stated as
Hmczn
0 in aqueous surfactant solutions varies linearly with
temperature [11–13], namely Hmczn 0 = AH T + BH , where Hmczn
0
= Cp,mczn
0
(T − TH=0 ) (20)
AH and BH are the slope and intercept, respectively, in
the Hmczn
0 versus T plot. Fig. 1 shows such observations; The heat capacity change, which is characteristic of the re-
Fig. 1a for eight hydrocarbon surfactants [11] and Fig. 1b for actions in surfactant systems, is a consequence of solvation
perfluoropolyether carboxylic acid and salt (H-, Na-PFPE) and particularly of the structuring of water molecules in the
and perfluoroocatanoic acid and salt (H-, Na-PFO) vicinity of surfactant chains related to hydrophobicity, and
[12]. is normally negative [30].

3.2. Indirect method


Table 2
Contributions of the quantities in the brackets of Eq. (18) to G0mczn /RT
Hmczn
0 may be determined from G0mczn and the
n 10 20 60 100 500 Gibbs–Helmholtz equation,
CMC (mM) 2.2 3.7 5.4 5.6 6.5  
−5.5074 −5.3195 −5.1343 −5.1331 −5.0259
∂G0mczn /T Hmczn
0
1st bracket =− . (21)
2nd bracket 0.547 0.3382 0.1485 0.0992 0.0263 ∂T T2
3rd bracket −0.0371 −0.0217 −0.0080 −0.0049 −0.0010
Since droping 1/n terms would cause the error within
Total −4.9975 −5.0030 −4.9938 −5.0388 −5.0006
0.5%, as demonstrated in Table 1, Eq. (19a) can be used as
H.-U. Kim, K.-H. Lim / Colloids and Surfaces A: Physicochem. Eng. Aspects 235 (2004) 121–128 125

a good approximate for G0mczn . Then one obtains


 
Hmczn
0 ∂ G0mczn ∂
− = = (1 + β) ln XCMC
RT 2 ∂T RT ∂T
∂β ∂ ∂n
+ln XCMC + f(n; β)
∂T ∂n ∂T
∂ ∂f(n; β) ∂n
= (1 + β) ln XCMC + (22)
∂T ∂n ∂T
where f(n;β)is defined as
1  m 2
f(n; β) = ln 2n(n + m) + 1 + ln
n n vS Fig. 2. Changes in thermodynamic potentials of tetradecyltrimethylam-
1 2 monium bromide (TTAB) as a function of temperature.
= ln 2n2 (1 + β) + (1 + β) ln (23)
n vS
The quantity f(n;β) is, in principle, a function of n and
β(≡ m/n). However, β, the degree of counterion bind-
ing, is weakly varying, although oscillatory in some cases
[10,31], or almost constant with temperature. For exam-
ple, for α-sulfonatomyristic acid methyl ester in water the
measured β was 0.712 ± 0.002 and the variation 0.002
amounts to 0.3% in the range from 5 to 50 ◦ C [7c]. Hence,
f(n;β) may be considered as a function of n only and this is
reflected in Eq. (22).
Determination of Hmczn 0 by the indirect method has
been done usually by using Eq. (22) without second
term, i.e., neglecting the contribution by the second term
[∂f(n; β)/∂n][∂n/∂T ]. Some researchers [32,33] have as-
Fig. 3. Linear behavior of Hmczn
0 (T) and Smczn
0 (T).
serted that such calculation of Hmczn
0 is not proper because
of the large changes in aggregation number that can oc-
where TS=0 is the temperature at which Smczn
0 = 0. Com-
cur with temperature. However, Krescheck and Hargraves
bining Eqs. (20) and (24), and eliminating T results in
[11,34] and Kiraly and Dekany [13] have defended this
procedure, chiefly on the ground that the Hmczn 0 thus Cp,mczn
0
calculated agree with the calorimetric estimates. Hmczn
0
= Smczn
0
+ Cp,mczn
0
(TS=0 − TH=0 )
AS
= Tcom Smczn
0
+ Icom (25)
4. Compensation phenomena Tcom (≡ Cp,mczn
0 /AS ) and Icom are the slope and the inter-
cept in the plots of Hmczn
0 versus Smczn
0 . Fig. 3 shows
For micellization in aqueous surfactant solutions, the en-
the linear behavior of Hmczn with Smczn , which various
0 0
thalpy and entropy of micellization experience large change
surfactant systems display. Tcom is called the compensation
with temperature. However, such large changes in enthalpy
temperature and characterizes the compensation phenom-
and entropy compensate each other, leaving the free energy
ena. In Fig. 3 Tcom or the slope appears to be identical and
change almost invariant and small in magnitude. This has
indeed its value falls in 307 ± 7 K for most of surfactant
been known as compensation phenomena and is illustrated
systems [20,21].
in Fig. 2 [35]. Hmczn0 is positive and large at low tem-
peratures and decreases almost linearly with temperature,
as stated before. At high temperatures micellization occurs
5. Dependence of critical micelle concentration on
just as readily as at low temperatures, since G0mczn is al-
temperature
most the same value as those at low temperatures. Smczn 0

is now large and negative, reflecting the ordering associated


Although it is widely known that the CMC is affected by
with the aggregation of monomers. Micellization is therefore
temperature, the correct relationship between the two has not
driven entirely by enthalpy changes at high temperatures.
been identified. Thus, a polynomial-function description of
Fig. 2 displays another important feature. Smczn0 also
CMC with respect to T has been used for most of surfactant
changes almost linearly with temperature, that is,
systems until now. However, polynomials of various orders
Smczn
0
= AS (T − TS=0 ) (24) can be used for a given surfactant system with acceptable
126 H.-U. Kim, K.-H. Lim / Colloids and Surfaces A: Physicochem. Eng. Aspects 235 (2004) 121–128

errors of fits to measured CMC. Hence, a rigorous relation


between CMC and T is desired so that the thermal behav-
ior of CMC can be described adequately for the surfactant
system of interest.
A new description of the critical micelle concentration
as a function of temperature is derived from G0mczn and
the linear behavior of Hmczn
0 (T) and Smczn
0 (T), i.e., from
Eqs. (9), (20) and (24). From the last two equations one
obtains

G0mczn = Hmczn
0
− TSmczn
0

Cp,mczn
0
= Cp,mczn
0
(T − TH=0 )−T (T − TS=0 )
Tcom
  Fig. 4. Comparison of fits by 2nd-order polynomial (dashed line) and Eq.
−Cp,mczn
0
TS=0
= T 2 + Cp,mczn
0
1+ T (29) (full line) for dodecyl-4-methoxypyridinium chloride (open circle,
Tcom Tcom measured [4]). .
−Cp,mczn
0
TH=0 . (26)
excellent agreement with the data. Figs. 5–8 show the re-
Combining Eqs. (19a) and (26) now yields
sults for nonionic n-dodecyl polyethylene glycol monoether
C (C12 E4 , C12 E6 , C12 E8 ) [8,37], α-sulfonatomyristic acid
ln XCMC = A + BT +
T methyl ester [7c], ethyl ester [20], cationic alkyltrimethy-
1 2 lammonium bromide(Ci (TAB), i = 6 [38], 8 [9,37], 10
− ln[2n2 (1 + β)] − ln , (27)
(1 + β)n vS
where the constants A, B, and C are defined as
 
Cp,mczn
0
TS=0
A≡ 1+ <0 (28a)
(1 + β)R Tcom
−Cp,mczn
0
B≡ >0 (28b)
(1 + β)RTcom
−Cp,mczn
0 TH=0
C= >0 (28c)
(1 + β)R
The constant A is negative, and B and C are positive, since
Cp,mczn
0 is negative. Tcom is around 307 K and TS=0 ≈
383 K at which the entropy of dissolution of liquid hydro- Fig. 5. Fits by Eq. (29) to the data of nonionic surfactant solutions: (䊐)
carbons is zero [36]. C12 E4 (correlation coefficient, r = 0.9934); (䉫) C12 E6 (r = 0.9984);
The fourth and the fifth terms of Eq. (27) is negligibly ( ) C12 E8 (r = 0.9899); (䊊) α-sulfonatomyristic acid methyl ester
small compared to the first three terms and therefore they (r = 0.9980); (+) α-sulfonatomyristic acid ethyl ester (r = 0.9994)).
can be dropped to obtain
C
ln XCMC = A + BT + . (29)
T

6. Results and discussion

Because Eq. (29) has three fitting parameters like a


2nd-order polynomial, ln XCMC = DT2 + ET + F , fits by
both equations are compared. Fig. 4 shows the results of the
fits by these equations for dodecyl-4-methoxypyridinium
chloride [4]. Both equations fit the data well and Eq. (29)
fits better than the 2nd-order polynomial (correlation coef-
Fig. 6. Fits by Eq. (29) to the data of cationic surfactant Ci (TAB)
ficients, 0.9958 versus 0.9883). solutions: (䊐) i = 6 (correlation coefficient, r = 0.9989); (䉫) i = 8
Eq. (29) has been employed to fit CMC data of various (r = 0.9812); ( ) i = 10 (r = 0.9924); (䊏) i = 12 (r = 0.9882); (䉬)
surfactant systems and it yields the results which are in i = 14 (r = 0.9777); (䉱) i = 16 (r = 0.9996).
H.-U. Kim, K.-H. Lim / Colloids and Surfaces A: Physicochem. Eng. Aspects 235 (2004) 121–128 127

Fig. 9. Hmczn
0 by Eq. (30) and the portion at T > 273 K for which
Fig. 7. Fits by Eq. (29) to the data of anionic surfactant sodium alkyl: Hmczn
0 changes linearly with temperature.
(䊐) octyl (correlation coefficient, r = 0.9866); (䉫) decyl (r = 0.9992);
( ) dodecyl (r = 0.9984); (䊊) tetradecyl (r = 0.9997) sulfate solutions.

Cp,mczn
0
[37,39], 12 [22,37], 14 [37,40], 16 [37,41]), anionic sodium Hmczn
0
= (Θ + 273)2 − Cp,mczn
0
TH=0
alkyl sulfate (SOS, SDeS, SDS, STS [3,42,43]) surfac- Tcom
 
tants, and amphiphilic drugs (cloxacillin, dicloxacillin [44], Cp,mczn
0
Θ 2
= 2732 1 + − Cp,mczn
0
TH=0
imipramine, clomipramine [45]), respectively. The results Tcom 273
in these figures imply that Eq. (29) describes correctly the  
(546)Cp,mczn
0
74529
temperature dependence of CMC for most of surfactant ≈ Θ+Cp,mczn
0
−TH=0 .
systems. Tcom Tcom
From G0mczn of Eq. (26) and the Gibbs–Helmholtz re- (31)
lation, ∂(G0mczn /T)/∂T = −Hmczn 0 /T 2 , Hmczn
0 can be
Here, use is made of the fact that Θ/273 < 1 and that
calculated. It appears that the calculated Hmczn differs
0
(1 + Θ/273)2 ≈ 1 + 2Θ/273. Fig. 9 shows the overall
from the observed one described by Eq. (20). Eq. (22) and
behavior of Eq. (30) and the portion for T > 273 K. At T >
the Gibbs–Helmholtz relation gives
273 K the quadratic behavior is approximated quite well by
Cp,mczn
0 linear one (correlation coefficient = 0.9990), as described
Hmczn
0
= T 2 − Cp,mczn
0
TH=0 . (30) by Eq. (31). This result implies that Eq. (30) is consistent
Tcom
with the observed linear behavior of Hmczn
0 at T > 273 K.
Hmczn
0 should vary quadratically with respect to T and this Eq. (29) is also used for the description of power law
behavior of CMC(T), |XCMC − XCMC ∗ | = const|T − T ∗ |d
appears to be inconsistent with the observed linear behavior
of Hmczn
0 . Eq. (20) represents the observations at T > Eq. (29) yields


273 K and Eq. (30) at T > 273 K also displays a linear XCMC = exp(−A + 2 BC) (32a)
behavior with T. Let Θ = T − 273 (Θ is the temperature 
in centigrade) and rewrite Eq. (30) in terms of Θ. Then we ∗ C
T = (32b)
obtain B
and one obtains, after substantial mathematical manipula-
tions [29],

|XCMC − XCMC | = const|T − T ∗ |2 (33)

The most noticeable finding with Eq. (33) is that the ex-
ponent d should be 2, irrespective of surfactant systems,
which is in disagreement with the previous values of d =
1.73 ± 0.03 by La Mesa [22], d = 3.54 for three com-
mercial ionic surfactants and d = 5.80 for two commercial
amphoteric surfactants by Stasiuk and Schramm [23], and
d = 1.05 ± 0.20 for ADS and d = 1.12 ± 0.15 for OTAC by
Kang et al. [10] The power-law equation, Eq. (33), is tested
for 28 surfactant systems including those examined by La
Fig. 8. Fits by Eq. (29) to the data of amphiphilic drug: (䊐) cloxacillin Mesa, Stasiuk and Schramm, and Kang et al. and it follows
(correlation coefficient, r = 0.9725); (䉫) dicloxacillin (r = 0.9765); ( ) closely the data, as shown in Fig. 10; the best fit is obtained
imipramine (r = 0.9913); (䊊) clomipramine (r = 0.9733) solutions. with d = 1.955 ± 0.015 (correlation coefficient = 0.9925).
128 H.-U. Kim, K.-H. Lim / Colloids and Surfaces A: Physicochem. Eng. Aspects 235 (2004) 121–128

[7] (a) M. Okawauchi, M. Hagio, Y. Ikawa, G. Sugihara, Y. Murata, M.


Tanaka, Bull. Chem. Soc. Jpn. 60 (1987) 2718;
(b) P. Mukerjee, K. Korematsu, M. Okawauchi, G. Sugihara, J. Phys.
Chem. 89 (1985) 5308;
(c) M. Fugiwara, T. Okano, T.-H. Nakashima, A.A. Nakamura, G.
Sugihara, Colloid Polym. Sci. 275 (1977) 474;
(d) T. Okano, T. Tamura, T.-Y. Nakano, S.-I. Ueda, S. Lee, G.
Sugihara, Langmuir 16 (2000) 3777.
[8] L.-J. Chen, S.-Y. Lin, C.-C. Huang, E.-M. Chen, Colloid Surf. A
135 (1998) 175.
[9] R. Zielinski, J. Colloid Interface Sci. 235 (2001) 201.
[10] K.-H. Kang, H.-U. Kim, K.-H. Lim, Colloid Surf. A 189 (2001) 113.
[11] G.C. Kresheck, W.A. Hargraves, J. Colloid Interface Sci. 48 (1974)
481.
[12] V. Tomasic, A. Chittofrati, N. Kallay, Colloid Surf. A 104 (1995) 95.
Fig. 10. Fit by Eq. (33) to CMC(T) data for various surfactant systems. [13] A. Kiraly, I. Dekany, J. Colloid Interface Sci. 242 (2002) 214.
[14] S. Paula, W. Sus, J. Tuchtenhagen, A. Blume, J. Phys. Chem. 99
This result may imply that Eq. (33) holds without regard to (1995) 11742.
surfactant system, at least for those examined. [15] P. Gilli, V. Ferretti, G. Gilli, P.A. Borea, J. Phys. Chem. 98 (1993)
1515.
[16] B. Madan, B. Lee, Biophys. Chem. 51 (1994) 279.
7. Conclusions [17] R. Lumry, S. Rajender, Biopolymers 9 (1970) 1125.
[18] C. Jolicoeur, R.P. Philip, Can. J. Chem. 52 (1974) 1834.
[19] V.C. Krishnan, L.H. Friedman, J. Solution Chem. 2 (1974) 37.
A new equation, ln XCMC = A + BT + (C/T)(Eq. (29)),
◦ [20] G. Sugihara, M. Hisatomi, J. Colloid Interface Sci. 219 (1999) 31.
has been derived on the basis of G = −RT ln K, lin- [21] H.U. Kim, K.-H. Lim, manuscript in preparation.
ear behavior of the enthalpy of micellization with tempera- [22] C. La Mesa, J. Phys. Chem. 94 (1990) 323.
ture, and compensation phenomena. This equation describes [23] E.N.B. Stasiuk, L.L. Schramm, J. Colloid Interface Sci. 324 (1996)
the temperature dependence of critical micelle concentra- 178.
[24] R.J. Hunter, Foundations of Colloid Science, vol. 1, Clarendon Press,
tion (CMC), XCMC , with three parameters, A, B, and C. The
Oxford, 1987, p. 575 (Chapter 10).
equation is tested for surfactant systems of various kinds [25] C. Tanford, Hydrophobic Effects, Wiley, New York, 1980.
such as nonionic, ionic, and zwitterionic surfactants, and [26] Y. Moroi, Micelles: Theoretical and Applied Aspects, Plenum Press,
amphiphilic drugs and it is found that the equation fits excel- New York, 1992.
lently the measured CMC of these surfactant systems. This [27] H.U. Kim, Ph.D. Dissertation, Chung-Ang University, Seoul, Korea,
2003.
result may imply that the new equation holds without regard
[28] J.N. Phillips, Trans. Faraday Soc. 151 (1955) 561.
to surfactant systems. [29] S.D. Hamman, Aust. J. Chem. 31 (1978) 919.
Eq. (29) has been also used for a power-law description for [30] N. Muller, Langmuir 9 (1993) 96.
the CMC(T), |XCMC − XCMC ∗ ∗
| = const|T − T ∗ |d with XCMC [31] G. Sugihara, Y. Arakawa, K. Tanaka, S. Lee, Y. Moroi, J. Colloid
and T∗ being the minimum CMC and the temperature at Interface Sci. 170 (1995) 399.

XCMC , respectively. It yields d = 2 irrespective of surfactant [32] N. Muller, in: K.L. Mittal (Ed.), Micellization, Solubilization, and
Microemulsions, vol. 1, Plenum Press, New York, 1977, pp. 229–239.
systems. [33] A. Holtzer, M.F. Holtzer, J. Phys. Chem. 78 (1974) 1442.
[34] G.C. Kresheck, in: F. Franks (Ed.), Water—a Comprehensive Trea-
tise, Plenum press, New York, 1975, pp. 95–167 (Chapter 2).
Acknowledgements [35] D.F. Evans, Langmuir 4 (1988) 3.
[36] S.J. Gill, N.F. Nichols, I. Wadsö, J. Phys. Chem. 8 (1976) 445.
H.-U. Kim was supported by the Korea Research Foun- [37] P. Mukerjee, K.J. Mysel, Critical Micelle Concentrations of Aqueous
Surfactant Systems, Nat. Stand. Ref. Data Ser. Nat. Bur. Stand. (US)
dation through the grant No. EN0037. The authors would
36 (1971).
like to thank one of the reviewers for his comments and [38] V. Mosquera, J.M. del Rip, D. Attwood, M. Garcia, M.N. Jones,
suggestions. G. Prieto, M.J. Suarez, F. Sarmiento, J. Colloid Interface Sci. 206
(1998) 66.
[39] N. Yoshida, K. Matsuoka, Y. Moroi, J. Colloid Interface Sci. 187
References (1997) 388.
[40] A. Castedo, J.L. Del Castillo, M.J. Suarez-Filloy, J.R. Rodriguez, J.
[1] D.D. Miller, L.J. Magid, D.F. Evans, J. Phys. Chem. 94 (1990) 5921. Colloid Interface Sci. 196 (1997) 148.
[2] P. Becher, in: M.J. Schick (Ed.), Nonionic Surfactants, Marcel [41] D.J. Lee, Colloid Polym. Sci. 273 (1995) 539.
Dekker, New York, 1967 (Chapter 15). [42] Y. Moroi, N. Nishikino, H. Uehara, R. Matuura, J. Colloid Interface
[3] B.D. Flockhart, J. Colloid Interface Sci. 16 (1961) 484. Sci. 50 (1975) 254.
[4] J.A. Stead, H.J. Taylor, J. Colloid Interface Sci. 30 (1969) 482. [43] J.M. Ruso, P. Taboada, V. Mosquera, F. Sarmiento, J. Colloid Inter-
[5] M.J. Blandamer, P.M. Cullis, L.G. Soldi, J.B.F.N. Engberts, A. face Sci. 214 (1999) 292.
Kacperrska, N.M. Van Os, M.C.S. Subha, Adv. Colloid Interface Sci. [44] P. Taboada, D. Attwood, M. Garcia, M.N. Jones, J.M. Ruso, V.
58 (1995) 171. Mosquera, F. Sarmiento, J. Colloid Interface Sci. 221 (2000) 242.
[6] K. Shinoda, M. Kobayashi, N. Yamaguchi, J. Phys. Chem. 91 (1987) [45] D. Attwood, E. Boitard, J.-P. Dubes, H. Tachoire, J. Colloid Interface
5292. Sci. 227 (2000) 356.

Você também pode gostar