Next Article in Journal
Extractive Spectrophotometric Determination and Theoretical Investigations of Two New Vanadium(V) Complexes
Previous Article in Journal
Effects of 2,2′-Azobis(2-methylpropionamidine) Dihydrochloride Stress on the Gel Properties of Duck Myofibrillar Protein Isolate
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Ternary Mixed Micelle Hexadecyltrimethylammonium Bromide—Dodecyltrimethylammonium Bromide—Sodium Deoxycholate: Gibbs Free Energy of Mixing and Excess Gibbs Energy of Mixing

Department of Pharmacy, Faculty of Medicine, University of Novi Sad, Hajduk Veljka 3, 21000 Novi Sad, Serbia
*
Authors to whom correspondence should be addressed.
Molecules 2023, 28(18), 6722; https://doi.org/10.3390/molecules28186722
Submission received: 27 August 2023 / Revised: 11 September 2023 / Accepted: 18 September 2023 / Published: 20 September 2023
(This article belongs to the Section Physical Chemistry)

Abstract

:
Pharmaceutical, food, and cosmetic formulations often contain binary or ternary surfactant mixtures with synergistic interactions amongst micellar building blocks. Here, a ternary mixture of the surfactants hexadecyltrimethylammonium bromide, dodecyltrimethylammonium bromide, and sodium deoxycholate is examined to see if the molar fractions of the surfactants in the ternary mixed micellar pseudophase are determined by the interaction coefficients between various pairs of the surfactants or by their propensity to self-associate. Critical micelle concentrations (CMC) of the analyzed ternary mixtures are determined experimentally (spectrofluorimetrically using pyrene as the probe molecule). Thermodynamic parameters of ternary mixtures are calculated from CMC values using the Regular Solution protocol. The tendency for monocomponent surfactants to self-associate (lower value of CMC) determines the molar fractions of surfactant in the mixed micelle if there is no issue with the packing of the micelle building units of the ternary mixed micelle. If a more hydrophobic surfactant is incorporated into the mixed micelle, the system (an aqueous solution of surfactants) is then the most thermodynamically stabilized.

1. Introduction

Surfactants are molecules or ions that contain both hydrophilic and hydrophobic regions in their structure. In an aqueous solution, surfactants are embedded in the aqueous solution–air interface and the hydrophobic molecular segment of the surfactant is oriented toward the air; while the hydrophilic polar head is orientated toward the aqueous phase, the Gibbs free energy of the water–surfactant system decreases. On a particular total surfactant concentration, known as the critical micelle concentration (CMC), the air and water phase’s boundary surfaces are completely saturated with surfactant particles. In order for the Gibbs free energy to further decrease in the water–surfactant system, surfactants form micelles, molecular aggregates in which the hydrophobic part of the surfactant forms the hydrophobic core of the micelle, while the polar groups of the surfactants form the polar outer shell of the micelle [1,2,3,4,5]. It has been demonstrated that below the CMC, surfactants form aggregates of a few (2–4) surfactants called premicellar aggregations, which can exhibit catalytic capabilities in certain chemical processes [6,7,8]. Both monocomponent surfactants and a precisely calculated mixture of surfactants can generate micelles, which are known as mixed micelles. Surfactant mixes that are either binary or ternary are typically utilized, and critical micellar concentrations for these mixtures have also been identified (i.e., determined experimentally) [9,10,11,12,13]. According to the phase separation model (applicable for micelles with large aggregation numbers, i.e., the number of surfactant particles in one micelle), mixed micelles (as micellar pseudophases) are formally created by mixing monocomponent micellar pseudophases [14,15,16,17,18]. The molar Gibbs free energy of the formation of a ternary mixed micellar pseudophase from a ternary mixture of surfactants dissolved in the aqueous phase (solution) is ( f g m M 0 ):
f g m M 0 = x 1 f g M 1 0 + x 2 f g M 2 0 + x 3 f g M 3 0 + g m i x i d + g m i x E
In Equation (1), f g M i 0 = R T ln C M C i   ( i = 1,2 , 3 ) represents the molar Gibbs free energy of the formation of monocomponent micellar pseudophases from which a ternary mixed micellar pseudophase is formally obtained (1 mol); if the critical micelle concentration of the monocomponent (pure) surfactant is C M C i , x i   ( i = 1,2 , 3 ) represents the molar fractions of surfactants in the ternary mixed micelle, i.e., micellar pseudophase; g m i x i d corresponds to the change in molar Gibbs free energy when obtaining a mixed micellar pseudophase as an ideal mixture of monocomponent micellar pseudophases under the conditions of constant pressure and temperature (i.e., ideal Gibbs free energy of mixing) [14,15,16]. The last term in Equation (1) presents the excess molar Gibbs free energy and contains energetic (entalpic) and entropic effects (interactions) not present in monocomponent micellar pseudophases [19]:
f g m M 0 x 1 f g M 1 0 + x 2 f g M 2 0 + x 3 f g M 3 0 = g m i x i d + g m i x E = g m i x r e l g m i x i d
The sum g m i x i d + g m i x E presents the real molar Gibbs free energy of mixing ( g m i x r e l ). Generally for ideal mixtures, g m i x E = 0 . In accordance, g m i x E describes the thermodynamic stabilization of real mixed micellar pseudophases, related to the ideal mixed micellar pseudophases [16]. According the to Regular Solution Theory (RST), the excess molar Gibbs free energy originates from energetic interactions between the first neighbors of structurally different surfactants, while surfactant conformations in mixed micellar pseudophases are identical to conformations of the same surfactants in monocomponent pseudophases [20,21]. Different types of particles are randomly distributed across the crystal lattice. This means that there is no excess molar entropy in RST (compared to an ideal mixture) [21,22]. The excess molar Gibbs free energy for the binary micellar pseudophase based on RST is expressed by the following symmetric Margules function of the first order:
g m i x E = h m i x E = R T β 12 x 1 x 2 ;   s m i x E = 0 R S T
where R is the universe gass constant, T presents the thermodynamic temperature of the system, while β 12 is the interaction coefficient (interaction parameter) between surfactants 1 and 2 in the binary mixed micellar pseudophase [14,15,16,20,21,22]. The coefficient of the interaction depends on the geometry of binary mixed micellar pseudophase which is represented as a quasi-crystalline structure whose parameter is the coordination number [20,21,22]. If β 12 < 0 , then there are synergistic interactions between surfactants 1 and 2 ( g m i x r e l < g m i x i d ), i.e., the real binary mixed micellar pseudophase is thermodynamically more stable than the ideal binary mixed micellar pseudophase. On the contrary, if β 12 > 0 , there are antagonistic interactions between surfactants 1 and 2, and real mixed micelle is thermodynamically less stable than the ideal mixed micelle ( g m i x r e l > g m i x i d ) [16,23,24,25]. According to Porter-u and van Laar, the Margules function of the first order (3) when s m i x E 0 can also describe the excess Gibbs free energy of mixing [26,27,28]. This is the case if the phenomenon of enthalpy–entropy compensation applies to the thermodynamic process of mixed micelle formation. Then, the excess molar Gibbs free energy is a symmetric function of the mole fraction of surfactants from the binary mixed micellar pseudophase [21,22,29,30,31]. In the case of the ternary micellar pseudophase, the excess molar Gibbs energy of mixing is the function of interaction parameters between structurally different surfactant particles: g m i x E = f β 12 , β 13 , β 23 β i j = β j i , Therefore, it is necessary to know the interaction parameters in binary mixed micellar pseudophases [11,32,33].
Monocomponent, binary, and ternary mixtures of surfactants are usually applied in pharmaceutical, food, and cosmetic formulations; in petro chemistry in the micellar catalysis of chemical reactions/synergistic interactions between surfactants, in addition to the thermodynamic parameters of micellization, they can also occur in different properties of surfactants, such as the solubilization capacity of hydrophobic molecules. If there are synergistic interactions between surfactants, and if β i j is negative enough, then the critical micelle concentration of the binary mixture of surfactants can have a lower value than the critical micelle concentration of the more hydrophobic surfactant in the mixture.
This means that to achieve the same (surface) effect from a binary mixture of surfactants, a smaller amount of surfactants is used than from a more hydrophobic surfactant—the ecological footprint of the binary mixture is reduced compared to monocomponent surfactants [4,5,6,7,8,9,10,11,12,16,34,35,36,37].
The aim is to determine whether the interaction coefficients between binary pairs of surfactants ( β i j ) or the tendency towards the self-association of monomeric surfactants determines the molar fractions of surfactants in the ternary mixed micellar pseudophase of hexadecyltrimethylammonium bromide (1)—dodecyltrimethylammonium bromide (2)—sodium deoxycholate (3) (Figure 1, Appendix A). Building units of the examined ternary micelle are cationic and anionic surfactants, so it is expected that there are synergistic interactions between cationic and anionic surfactants due to Columb’s electrostatic attractive interactions. The different geometry of the hydrophobic segment of cationic (1) (2) on one side and anionic (3) surfactants on the other side results in the excess of molar (conformational) entropy [38]. The different lengths of the hydrocarbon chains between the examined cationic surfactants (1) and (2) means that surfactants (1) and (2) have different tendencies towards self-association, i.e., the cationic surfactant with the longer hydrocarbon chain has the lower value of critical micelle concentration (i.e., greater tendency towards self-association). In the ternary mixture of surfactants, the molar fraction of sodium-deoxycholate is constant ( α 3 = 0.6 ), whereas the cationic surfactant ratio changes: α 1 + α 2 + 0.6 = 1 (the molar fraction of surfactants in the starting ternary mixture of surfactants—which dissolves in the aqueous phase—differs from the molar fraction of the same surfactants in the formed ternary mixed micelles x i α i ). The mole fraction of sodium deoxycholate (3) was chosen based on our earlier investigations into the solubilization capacity of binary mixtures, i.e., for α 3 = 0.6 in a binary mixed micelle, it approximately corresponds to x 3 0.5 , which is desirable for the solubilization of flavonoids, isoflavonoids, and polyphenols (bile acid anions form hydrogen bonds with flavonoids and, thus, are incorporated into the hydrophobic core of the mixed micelle) [39].

2. Results and Discussion

At some values of α 1 + α 2 + 0.6 = 1 , a reaction of ternary mixed micelle formation is observed (as a summary process of successive and parallel association processes [40], since the tested micelle does not bind counterions, the charge of the micelle was not taken into consideration):
n i 1 ( a q ) + n 2 2 a q + n 3 3 a q m M x 1 = n 1 n 1 + n 2 + n 3 , x 2 = n 2 n 1 + n 2 + n 3 ,   x 1 + x 2 + x 3 = 1   ( a q )
where the micelle’s composition and the micelle’s aggregation number correspond to the mean values over the reaction system. The chemical potentials of surfactants, i.e., ternary mixed micelles as separate particles, can be expressed by Henry’s law, where the standard state refers to an infinitely (ideally) diluted solution. Thus, in the equilibrium state for the process (4), the following holds:
G ξ p , T = 0 = μ m M 0 n 1 μ 1 0 n 2 μ 2 0 n 3 μ 3 0 + R T ln x m M b n 1 ln x 1 b n 2 ln x 2 b n 3 ln x 3 b
In the above equation, μ m M 0 is the standard chemical potential of the ternary mixed micelles in an aqueous solution. At the same equation, μ i 0   ( i = 1,2 , 3 ) represents the standard chemical potential of a surfactant (monomer) in an aqueous solution; x m M b and x i b   ( i = 1,2 , 3 ) correspond to the molar fractions of the mixed micelles and surfactants in an aqueous solution. Multiplying Equation (5) by the reciprocal of the aggregation number of the ternary mixed micelle ( 1 / n 1 + n 2 + n 3 and n = n 1 + n 2 + n 3 ), we obtain the equation:
μ m M 0 n 1 μ 1 0 n 2 μ 2 0 n 3 μ 3 0 n = + R T ln x m M b 1 n n 1 n ln x 1 b n 2 n ln x 2 b n 3 n ln x 3 b
If the aggregation number of the ternary micelle tends to infinity, then the following limit value applies:
lim n ln x m M b 1 n = 0
Therefore, Equation (6) is:
μ m M 0 n 1 μ 1 0 n 2 μ 2 0 n 3 μ 3 0 n = R T x 1 ln x 1 b + x 2 ln x 2 b + x 3 ln x 3 b
f g m M 0 = R T x 1 ln x 1 b + x 2 ln x 2 b + x 3 ln x 3 b
In Equation (9), x i b   ( i = 1,2 , 3 ) can be expressed as products: critical micellar concentrations of the ternary mixture of the surfactant ( C M C 123 ) and the molar fraction of the corresponding surfactant from the binary mixture of surfactants ( α i ):
f g m M 0 = R T x 1 ln α 1 C M C 123 + x 2 ln α 2 C M C 123 + x 3 ln α 3 C M C 123
If the limiting value (7) is valid, formation of the ternary mixed micelle can equally be described using the law of mass action (i.e., association reaction (4)) and the phase separation method. In the state of equilibrium, the chemical potential of the surfactant from the ternary mixed micellar pseudophase is equal to the chemical potential of the same surfactant from the aqueous phase [14,15,16]:
μ ( a q ) 0 + R T l n C M C i μ m m 0 + R T l n x i f i m i x e d   m i c e l l a r   p s e u d o p h a s e = μ ( a q ) 0 + R T l n α i C M C 123 w a t e r   p h a s e ,   ( i = 1,2 , 3 )
From the equality of chemical potentials follows the expression (11):
C M C i x i f i = α i C M C 123 ,   ( i = 1,2 , 3 )  
where f i represents the coefficient of the activity of surfactants in ternary mixed micelle pseudophases (if the activity coefficient of a particle in a real mixture has a lower value than 1, then the observed particle is thermodynamically more stable than in a hypothetical ideal mixture under the same conditions [27]). By introducing expression (12) into Equation (10):
f g m M 0 R T = x 1 ln C M C 1 + x 2 ln C M C 2 + x 3 ln C M C 3 x 1 f g M 1 0 + x 2 f g M 2 0 + x 3 f g M 3 0 + x 1 ln x 1 + x 2 ln x 2 + x 3 ln x 3 g m i x i d + x 1 ln f 1 + x 2 ln f 2 + x 3 ln f 3 g m i x E
we get Equation (13) that is equivalent to expression (1); of course, if the limit value (7) holds. Rearranging Equation (10) gives:
f g m M 0 = R T x 1 ln C M C 123 + x 2 ln C M C 123 + x 3 ln C M C 123 + x 1 ln α 1 + x 2 ln α 2 + x 3 ln α 3
and if the limiting value exists:
lim n ln α i 1 n = 0 ,   ( i = 1,2 , 3 )
then Equation (14) is:
f g m M 0 = R T x 1 + x 2 + x 3 1 ln C M C 123 + n 1 ln α 1 1 n + n 2 ln α 2 1 n + n 3 ln α 3 1 n 0
f g m M 0 = R T ln C M C 123
Thus, if the aggregation number of the ternary mixed micelle is relatively high, then the same expression (17) is obtained for f g m M 0 , starting from the association reaction (4) or from expression (1) in the phase separation theory (Figure 2). If classic surfactants with hydrocarbon chains form mixed micelles with bile acid anions, then molecular dynamic simulations show that the applicability of limit values (7) and (15) is justified [41]. In Equation (17), C M C 123 is expressed in mol·dm−3 or in the surfactant mole fraction. For surfactants with low C M C 123 values (<10 mM), the values of f g m M 0 would only differ by the constant term ≈ ln 55.5 = 4.02 when using one or the other unit [42].
In deriving Equation (17) from the associated reaction (4), the micelle charge and the binding of counterions to the micelle were not taken into account, i.e., partial neutralization of the micellar charge. The function of the dependence of the specific conductivity of the aqueous solution of the ternary mixture of the investigated surfactants on the concentration of the ternary mixture (Figure 3) ( C M C 123 ) was determined spectrofluorimetrically, with pyrene as the probe molecule (Table 1). This means that the examined ternary mixed micelle does not bind counterions to its outer shell, which is a consequence of the presence of bile acid anions (3), which, with their steroid skeleton on the surface of the micelle, disturb the continuous electrostatic potential of spheroidal symmetry and which exists on the surface of cationic monocomponent micelles. This then makes it difficult to form a Helmholtz layer in the Stern double layer—micellar building elements behave like individual surfactant particles in terms of charge, i.e., they are in a completely dissociated state [21,41].
In the examined ternary mixed micellar pseudophase, each surfactant on each temperature has the negative value of the logarithm of the coefficient of activity; also, the excess molar Gibbs energy is negative, which means that the real ternary mixed micellar pseudophase is thermodynamically more stable than the ideal micellar pseudophase (Table 1). By applying Nagarajan’s molecular theory of the micellar state, the following expression can be written for the excess molar Gibbs free energy [19]:
g m i x E = x 1 h o r d R T ln Ω m M Ω M 1 + x 2 h o r d R T ln Ω m M Ω M 2 + x 3 h o r d R T ln Ω m M Ω M 3 0 + γ a x 1 a 1 x 2 a 2 x 3 a 3 + x 1 g d i p o l + x 2 g d i p o l + g C o l u m b
The first three terms present the change in enthalpy ( h o r d ) and entropy (expressed over the number of microstates in the mixed micelle ( Ω m M ) and in monocomponent micelle ( Ω M )) for each surfactant due to the change in the conformational states of the hydrocarbon chain in the mixed micelle compared to the conformations in monocomponent micelles [38]. As the anion of deoxycholic acid (3) contains a conformationally rigid steroid skeleton in its structure, the conformation of the anion of deoxycholic acid is independent of the micellar packing, i.e., the structure of the neighboring surfactant and the coordination number of the micellar pseudophase. The fourth term of the expression (18) originates from the fact that the effective surface area of the hydrophobic core of the ternary mixed micelle (a) is not a linear combination of the product of the mole fractions of the surfactants and the effective surface area of the hydrophobic cores of monocomponent micelles ( x i a i ,   ( i = 1,2 , 3 ) ) [19]. The fifth term ( x 1 g d i p o l + x 2 g d i p o l ) relates to the cation– dipole interaction between cationic surfactants and the C3 pseudoaxial and C12 axial OH groups of the deoxycholic acid anion (3) [22]. The last term takes into account the reduction in repulsive electrostatic interactions between identical charges in monocomponent micelles during the formation of ternary mixed micelles with attractive electrostatic interactions between anionic and cationic surfactants.
In the system of three structurally different surfactants in the aqueous solution, surfactant (1) is the most hydrophobic (has the lowest critical micellar concentration, Figure 1 and Appendix B); thus, surfactant (1) probably first forms a monocomponent micelle in which there are particles (1) with elongated hydrocarbon chains and more or less globular conformations [43] (in order to evenly fill the micellar core). The deoxycholic acid anion incorporates in the micelle by substituting surfactant (1) in globular conformation—the steroid skeleton of surfactant (3) is localized in the micellar groove, i.e., between elongated conformations of surfactant (1) as the convex surface (β side of the steroid skeleton with angular methyl groups) particles (3) are oriented towards the interior of the micelle, while the axial OH groups and the C17 side chain with carboxylate group towards the aqueous solution orientation of the deoxycholic acid anion is confirmed by cross peaks in a 2D ROESY [44,45] experiment (Figure 4 and Figure 5). Surfactant (2), the least hydrophobic surfactant in the examined mixture, probably incorporates with the elongated conformations in the micelle of surfactants (1) and (3).
Therefore, in ternary micelles of cationic surfactants, (1) and (2) decrease the number of conformational microstates related to the monocomponent micelles; so, in the expression (18), the first and the second term are:
x 1 h o r d R T ln Ω m M Ω M 1 + x 2 h o r d R T ln Ω m M Ω M 2 > 0
i.e., the effect of reducing the number of conformational microstates in ternary mixed micelles thermodynamically destabilizes the real mixed micelle in relation to the state of the ideal mixed micelles. In monocomponent micelles of surfactants (1) and (2), when the hydrocarbon chain of cationic surfactant is in globular conformation, the hydrophobic molecular surface is exposed to hydration [43]. The substitution of globular conformations of cationic surfactants with deoxycholic acid anions results in a decrease in the degree of hydrophobic hydration since the effective hydrophobic surface of the ternary mixed micelle decreases compared to monocomponent micelles (Figure 5). Thus, the fourth term in the expression (18) is:
γ a x 1 a 1 x 2 a 2 x 3 a 3 < 0
and thermodynamically stabilizes real mixed micelles towards the ideal mixed micelles. Since the intermolecular interactions between the dipole OH surfactant (3) and cationic surfactants (1) and (2) do not exist in monocomponent micelles (1) and (2), the last terms of expression (18) have a stabilizing contribution to the ternary mixed micelle compared to the ideal ternary mixed micelle (Figure 6):
x 1 g d i p o l + x 2 g d i p o l + g C o l u m b < 0
The ternary mixed micelle’s excess molar Gibbs free energy is g m i x E < 0 (Table 1), hence, it must be:
x 1 g o r d 1 + x 2 g o r d 2 < γ a x 1 a 1 x 2 a 2 x 3 a 3 + x 1 g d i p o l + x 2 g d i p o l + g C o l u m b
where g o r d = h o r d R T l n ( Ω m M / Ω M ) , ord = ordering.
In the case of a ternary mixture of surfactants with the following composition: α1 = 0.05; α2 = 0.35; α3 = 0.6 (Table 1), in the formed ternary mixed micellar pseudophase, the molar fractions of cationic surfactants are approximately equal to each other; although, in the initial mixture, the amount of cationic surfactants with a shorter hydrocarbon chain (2) is seven times greater than the amount of cationic surfactants with a longer hydrocarbon chain (1).
In binary micellar pseudophases, both cationic surfactants have synergistic interactions of similar strengths with the deoxycholic acid anion (Appendix B). This means that in the formation of the ternary mixed micellar pseudophase system (i.e., surfactant aqueous solution), the more hydrophobic surfactant (1) that is incorporated into the mixed micelle, the micelle is more stabilized. In this way, a larger amount of water molecules from the hydration layer above the hydrophobic molecular surface of the monomeric surfactant (water molecules that have a lower entropy compared to water molecules from the interior of the aqueous solution [46]) move into the interior of the aqueous solution; thus, the entropy of the system increases (especially at lower temperatures [47,48,49,50,51,52]). The relation g m i x E / f g m M 0 grows with the temperature rise (Table 1), since with increasing temperatures the difference in entropy between water molecules in the hydration layer above the hydrophobic molecular surface and water molecules inside the aqueous solution decreases [46], i.e., in Equation (1), the absolute value of first three terms decreases ( x 1 f g M 1 0 + x 2 f g M 2 0 + x 3 f g M 3 0 ), in which, according to Nagarajan and Tanford, the entropy change due to the transfer of the hydrophobic molecular segment from the aqueous phase to the hydrophobic environment is incorporated (along with the dehydration of the hydrophobic surface) [19]. However, since g m i x E contains electrostatic attractive interactions between cationic and anionic surfactants that do not change with the temperature, the relation g m i x E / f g m M 0 slightly decreases ( g m i x E according to Equation (1) is incorporated in f g m M 0 ; both Gibbs energies slighly increase with temperature (Table 1) due to the increase in hydrophobic interactions (i.e., van der Waals interactions [46]) that do not exist in monocomponent micelles as they are not linear functions of hydrophobic interactions from monocomponent micelles). Similarly, g m i x E / f g m M 0 changes in the case of the ternary mixtures of the aqueous solutions of surfactants of the composition α1 = 0.1; α2 = 0.3; α3 = 0.6 (Table 1). However, in this case with a temperature rise, it decreases the molar fraction of surfactant (1), while the molar fraction of surfactant (2) grows in the micellar pseudophase. At higher temperatures, surfactant (1) probably takes some globular conformations as well, so the surface below globular surfactant (1) easily (without steric repulsive interactions) fills with surfactants with shorter hydrocarbon chains. For other examined mixtures, the relation g m i x E / f g m M 0 does not change with the temperature, but remains more or less constant (with some fluctuation). Namely, with the increase in α1 in the ternary mixed micellar pseudophase, the probability increases, especially at higher temperatures, so that a certain fraction of surfactant (1) receives a more or less globular conformation, whereby in the hydrocarbon chains of surfactant (1), partial synclinal and synperiplanar conformations are formed (i.e., conformations with repulsive interactions). [38]. Thus, the part of the electrostatic attractive interaction (energy) is used for achieving certain conformation states (especially at higher temperatures) [21], which results in a more or less constant value of the relation g m i x E / f g m M 0 in the temperature dependence.

3. Materials and Methods

All chemicals were used from the original manufacturer’s packaging (Table 2).

3.1. Determinations of Critical Micellar Concentration

Stock solutions of surfactants (hexadecyltrimethylammonium bromide (1), dodecyltrimethylammonium bromide (2), sodium deoxycholate (3)) prepared in deionized water were mixed in the ternary mixtures with different molar ratios of surfactants (0.5:3.5:0.6, 1:3:6, 2:2:6, 3:1:6, 0.35:0.05:0.6). In the ternary mixture of the ratio 0.5:3.5:0.6, the ratio between surfactants 1:2 is 0.125:0.875; 1:3 is 0.077:0.923, and 2:3 is 0.368:0.632; these binary mixtures of surfactants were also measured in order to calculate the interaction factors of surfactants in the binary mixtures that were used for calculations in the ternary mixture. Accordingly, in Table 3, all the measured ternary and binary mixtures of surfactants are presented.
Critical micellar concentrations of monocomponent surfactants and their secondary and ternary mixtures were measured spectrofluorimetrically on Cary Eclipse fluorescence spectrophotometer (Agilent, Waldbronn, Germany) using pyrene as the probe molecule. The ratio of the pyrene fluorescence intensities of the first and the third vibronic peaks (a measure of the environmental polarity) was measured in the function of total concentration of surfactant on temperatures (278.1–313.1 K, in intervals of 5 K; the temperature variation is 0.1 K) and critical micelle concentrations were determined by curve fitting with the Boltzmann equation (OriginaLab 9 software) (Appendix C). Physico-chemical parameters of examined surfactant mixtures were calculated using Mathlab.
Thermodynamic parameters of ternary mixed micelle (Table 1) are calculated according to RST protocol [11,32,33] from data (coefficients of interaction) of binary systems (Appendix B).

3.2. Conductivity Measurements

The goal of the conductometry measurement was to determine the fraction of counter ion binding to the mixed micelle. Conductivity was measured by gradual dilution of surfactant mixture solutions with deionized water. The data were acquired using a Consort C 860 conductometer. Equipment was calibrated with KCl solution ranging from 0.01 to 1.0 mol·dm−3 of known κ (specific conductivity). The cell containing solutions was immersed in a water bath, controlling the temperature variation at 0.1 K. The temperature was kept constant at 293.1 K.

3.3. Conductivity Measurements

For the NMR experiments, samples were prepared as D2O solution and 0.7 mL of solution were used for the measurement. Spectra were recorded on a Bruker AVANCE III HD 400 MHz spectrometer, equipped with Prodigy cooled probe head. For ROESY experiments, standard Bruker pulse program with water suppression (roesyphpr.2) was used with the spin-lock pulse length of 350 ms.

4. Conclusions

The examined ternary mixed micelle is thermodynamically more stable than a hypothetical ideal mixed micelle ( g m i x E < 0 ). Thermodynamic stabilization is most likely a result of electrostatic attractive interactions between cationic surfactants and anions of deoxycholic acid (3), as well as ion–dipole interactions involving the OH groups of the surfactant (3). If a ternary mixture has an α1 lower than 0.2, then the g m i x E / f g m M 0 grows with the temperature; if α1 is higher than 0.2, then the ratio of Gibbs free energies g m i x E / f g m M 0 is more or less constant.

Author Contributions

Conceptualization, visualization, writing—original draft preparation, supervision, M.P.; methodology, software, writing—original draft preparation, A.P.; formal analysis, data curation, resources, V.T.; investigation I.V. and G.P. All authors have read and agreed to the published version of the manuscript.

Funding

This study was supported by the Republic of Serbia, Autonomous Province of Vojvodina, Provincial Secretariat for High Education and Scientific Research, grant no. 142-451-3155/2022-02, and Project of Ministry of Science and Technological Development of Republic of Serbia, grant no. 451-03-47/2023-01/200114.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Not applicable.

Appendix A

Bile salts are biosurfactants, with significant physiological roles. Due to their capacity to alter different physiological barriers (membranes or cell contacts in the body), they are used as drug carriers in pharmaceutical formulations and have a positive influence on the transport process of some drugs as well as on their pharmacodynamics. This increases the transport of some drugs to specific tissues (modulation of blood–brain barrier) or into the interior of cells. Bile salt anions create relatively tiny micelles with up to 20 structural units. Additionally, when combined with other surfactants, they can create mixed micelles with a significant hydrophobic domain, increasing their ability for solubilization [53,54,55].
Quaternary ammonium salts hexadecyltrimethylammonium bromide (1) and dodecyltrimethylammonium bromide (2) are surfactants that have diverse, important applications: (1) is used as a tumoricidal irritant in colorectal cancer surgery and in hydatid cyst operations, has shown anticancer cytotoxicity with various cancer cell lines and in pharmaceutical industry, and it is used as an antimicrobial agent. (2) acts as a foaming agent, stabilizer, paint stripper dispersing agent, and bactericidal lotions, plays an important function in dying process of fabrics and in textile industry and waste water treatment and it is used in the extraction process of DNA by using meat products [56,57,58].

Appendix B

Critical micelle concentrations, composition and interaction parameter of binary mixtures hexadecyltrimethylammonium bromide (1)—dodecyltrimethylammonium bromide (2)—sodium deoxycholate (3): (A) 1 + 2, (B) 1 + 3 and (C) 1 + 2.
Table A1. (A).
Table A1. (A).
278.1 K283.1 K288.1 K
α1CMCx1βCMCx1βCMCx1β
011.541 11.336 11.968
0.1251.13470.54−5.831.22360.53−5.591.23730.53−5.82
0.251.2920.62−3.941.32090.61−3.931.39950.61−3.98
0.51.35050.8−1.981.39370.79−1.971.48570.78−2.03
0.751.11860.92−1.441.15320.91−1.531.2310.9−1.67
0.8750.65950.79−5.390.98290.91−2.530.93420.85−3.75
10.9218 0.9627 1.043
293.1 K298.1 K303.1 K
α1CMCx1βCMCx1βCMCx1β
013.431 14.195 14.154
0.1251.27690.54−5.981.290.53−6.361.33180.51−6.69
0.251.59740.62−3.691.77610.62−3.731.84510.59−4.12
0.51.4540.77−2.461.47850.73−3.041.55620.69−3.56
0.751.40660.980.111.5480.93−1.051.77020.87−1.81
0.8751.00480.87−3.531.20320.88−3.131.02550.77−5.47
11.0795 1.2513 1.5614
308.1 K313.1 K
α1CMCx1βCMCx1β
018.59 21.161
0.1251.42710.53−6.781.52690.53−6.7
0.251.9640.62−4.122.06390.62−4.25
0.51.64820.73−3.412.0280.77−2.68
0.751.63980.88−2.241.94870.96−0.65
0.8751.3790.88−3.391.48150.89−3.27
11.4525 1.526
The relative standard uncertainty of the critical micelle concentrations 4%.
Table A2. (B).
Table A2. (B).
278.1 K283.1 K288.1 K
α1CMCx1βCMCx1βCMCx1β
062.565 59.484 59.773
0.0770.46220.47−9.080.48490.47−8.860.51810.47−8.76
0.1420.44210.51−8.210.53140.50−7.460.58620.50−7.23
0.250.22470.53−10.000.23710.53−9.890.23590.53−9.98
0.330.12040.54−12.110.11150.54−12.530.12490.54−12.15
0.3680.10840.55−12.410.12080.55−11.980.12000.54−12.19
10.9218 0.9627 1.043
293.1 K298.1 K303.1 K
α1CMCx1βCMCx1βCMCx1β
063.919 65.737 66.675
0.0770.53050.47−8.870.59450.46−8.750.60290.45−9.13
0.1420.92530.50−5.610.95210.49−5.8510.1130.48−6.06
0.250.24610.53−10.010.26560.52−10.070.30520.51−10.00
0.330.13190.54−12.130.14640.53−12.080.21290.53−11.08
0.3680.13330.54−11.960.17080.54−11.330.20510.53−11.10
110.795 12.513 15.614
308.1 K313.1 K
α1CMCx1βCMCx1β
079.129 75.430
0.0770.64990.46−9.070.65590.46−9.02
0.14210.3360.49−6.1910.3760.49−6.17
0.250.44740.53−8.640.61960.53−7.35
0.330.20580.54−11.380.26820.54−10.33
0.3680.25100.55−10.450.26820.54−10.20
114.525 1.526
The relative standard uncertainty of the critical micelle concentrations 4%.
Table A3. (C).
Table A3. (C).
278.1 K283.1 K288.1 K
α1CMCx1βCMCx1βCMCx1β
062.565 59.484 59.773
0.07710.0710.37−10.2610.4820.37−10.0010.5100.37−10.02
0.1420.98780.40−9.540.99260.39−9.3710.0650.39−9.41
0.250.97730.42−8.960.98110.42−8.790.97740.42−8.91
0.330.95330.44−8.890.99280.44−8.5910.0720.43−8.64
0.3680.87390.45−9.120.82480.45−9.210.85780.44−9.16
111.541 11.336 11.968
293.1 K298.1 K303.1 K
α1CMCx1βCMCx1βCMCx1β
063.919 65.737 66.675
0.07711.1820.37−10.1211.3600.37−10.2112.9820.36−9.67
0.14210.1120.39−9.7510.7450.39−9.6611.0880.39−9.56
0.250.97370.42−9.290.99620.42−9.3610.6930.42−9.09
0.330.97770.44−9.100.97490.43−9.2810.3970.43−9.05
0.3680.89930.44v0.86210.44−9.660.90620.44v
113.431 14.195 14.154
308.1 K313.1 K
α1CMCx1βCMCx1β
079.129 75.430
0.07713.3160.37−10.4623.6290.33−7.99
0.14211.4300.39−10.3212.1140.38−10.16
0.2510.2850.42−10.1311.1030.41−9.92
0.3310.0850.44−10.0210.2660.43−10.07
0.36810.6930.44−9.6910.0470.43−10.07
118.59 21.161
The relative standard uncertainty of the critical micelle concentrations 4%.
According to Guggenheim’s original work [20], the interaction coefficient is constant over the entire range of the molar fraction of the binary mixture. However, the geometry of the quasi-crystalline lattice, i.e., the coordination number, does not change either. For surfactants, due to packing problems [22,59], the coordination number of the micellar pseudophase can change, and thus the interaction coefficient also changes. The negative values of the interaction coefficient for the binary cationic mixed micelle (1)–(2) are probably the result of the different length of the hydrocarbon chains between the cationic surfactants, which then facilitates their packing [43]. The literature CMC value of sodium deoxycholate (3) at room temperature measured by fluorescence method with pyrene as the probe molecule is 5.3 mM [60] which is lower than our experimentally determined value. However, Subuddhi and Mishra declared CMC values of sodium deoxycholate to be 5 mM at 288 K, 6 mM at 298 K, and 8 mM at 308 K, when measured by fluorescence lifetime measurement, which trend is in consistent with our results [61]. Hexadecyltrimethylammonium bromide’s (1) CMC value was measured by conductometry and is 0.94 mM at 298 K (we measured 1.24 mM) [62], while dodecyltrimethylammonium bromide’s (2) CMC value is 15.9 at 293 K (our measurement is 13.43 mM) [63].
Critical micelle concentrations of ionic surfactants, as bile acid anions (bile salts), sodium dodecylsulfate, cetyltrimethylammonium bromide, etc., show temperature dependence whose function is U-shaped. On the certain temperature, TH micelle formation is exclusively entropic in nature- entropy driven T = (293.1–298.1) K, i.e., the change in the enthalpy of micellization is zero, so on this temperature is the highest hydrophobic effect (the passage of water molecules from the hydration layer above the hydrophobic surface of the monomeric surfactant into the interior of the aqueous solution). As the temperature increases, the entropy difference between water molecules from the bulk of the aqueous solution and water molecules from the hydration layer above the hydrophobic molecular surface decreases, which means that with an increase in temperature, the change (increase) in entropy of micellization decreases (the entropic effect of micellization decreases), i.e., the tendency towards self-association decreases, which manifests itself with an increase in the critical micellar concentration with temperature (this is in agreement with the results from Table 1 and Appendix B). There is also a temperature (T > 350.1 K) TS, on which the change in the entropy of micellization is zero, the smallest hydrophobic effect, and micellization is the result of the enthalpic effect—enthalphy becomes the major driving force for aggregation (hydrophobic interaction). Hydrophobic interaction is the result of the induced dipol interactions between hydrophobic surfaces of the micelle building units in the micelle core. Deviations are possible if the aqueous solution of surfactant contains certain additives that disturb the structure of the bulk water [46,47,48,49,50,51,52,64].

Appendix C

Each tested aqueous solution of the ternary mixture of surfactants was transparent, without sedimentation or cloudiness.
Figure A1. Boltzmann function of the dependence of the ratio of the pyrene fluorescence intensities of the first and the third vibronic peaks from the concentration of the ternary mixture of surfactants (T = 293.1 K, α1 = 0.2, α2 = 0.2, α3 = 0.6).
Figure A1. Boltzmann function of the dependence of the ratio of the pyrene fluorescence intensities of the first and the third vibronic peaks from the concentration of the ternary mixture of surfactants (T = 293.1 K, α1 = 0.2, α2 = 0.2, α3 = 0.6).
Molecules 28 06722 g0a1

References

  1. Rusanov, A.I. The wonderful world of micelles. Colloid J. 2014, 76, 121–126. [Google Scholar] [CrossRef]
  2. Rusanov, A.I. The mass action law theory of micellar solutions. Adv. Colloid Interface Sci. 1993, 45, 1–78. [Google Scholar] [CrossRef]
  3. Poša, M. Self-Association of the Anion of 7-Oxodeoxycholic Acid (Bile Salt): How Secondary Micelles Are Formed. Int. J. Mol. Sci. 2023, 24, 11853. [Google Scholar] [CrossRef]
  4. Kumar, H.; Kaur, G. Scrutinizing Self-Assembly, Surface Activity and Aggregation Behavior of Mixtures of Imidazolium Based Ionic Liquids and Surfactants. A Comprehensive Review. Front. Chem. 2021, 9, 667941. [Google Scholar] [CrossRef] [PubMed]
  5. Szymczyk, K.; Jańczuk, B. The adsorption at solution–air interface and volumetric properties of mixtures of cationic and nonionic surfactants. Colloids Surf. A 2007, 293, 39–50. [Google Scholar] [CrossRef]
  6. Song, L.D.; Rosen, M.J. Surface properties, micellization, and premicellar aggregation of gemini surfactants with rigid and flexible spacers. Langmuir 1996, 12, 1149–1151. [Google Scholar] [CrossRef]
  7. Bhattarai, A.; Rub, M.A.; Posa, M.; Saha, B.; Asiri, A.M.; Kumar, D. Studies of ninhydrin and phenylalanine in cationic dimeric gemini micellar system: Spectrophotometric and conductometric measurements. Colloids Surf. A 2022, 655, 130334. [Google Scholar] [CrossRef]
  8. Poša, M.; Bhattarai, A.; Khan, J.M.; Saha, B.; Kumar, D. Impact of double headed geminis on leucine and ninhydrin reaction in buffer solvent. Colloids Surf. A 2023, 674, 131951. [Google Scholar] [CrossRef]
  9. Azum, N.; Rub, M.A.; Asiri, A.M. Bile salt–bile salt interaction in mixed monolayer and mixed micelle formation. J. Chem. Thermodyn. 2019, 128, 406–414. [Google Scholar] [CrossRef]
  10. Rub, M.A.; Azum, N.; Khan, F.; Asiri, A.M. Aggregation of sodium salt of ibuprofen and sodium taurocholate mixture in different media: A tensiometry and fluorometry study. J. Chem. Thermodyn. 2018, 121, 199–210. [Google Scholar] [CrossRef]
  11. Ren, Z.H.; Huang, J.; Zheng, Y.C.; Lai, L.; Mei, P.; Yu, X.R.; Chang, Y.L. Micellization and interaction for ternary mixtures of amino sulfonate surfactant and octylphenol polyoxyethylene ethers in aqueous solution: 2 blending with surfactant with a longer or shorter hydrophilic chain. J. Mol. Liq. 2018, 272, 380–386. [Google Scholar] [CrossRef]
  12. Sheng, R.; Quan, X.F.; Ren, Z.H.; Huang, J.; Li, D.N.; Wang, J.R.; Qian, Z.B.; Zhang, Y.X.; Chai, L.L.; Li, B.B.; et al. Molecular interaction between sodium dodecylbenzene sulfonate and octylphenol polyoxyethylene ether and effect of hydrophilic chain. Colloids Surf. A 2021, 626, 127048. [Google Scholar] [CrossRef]
  13. Rubingh, D.N. Mixed Micelle Solutions. In Solution Chemistry of Surfactants; Mittal, K.L., Ed.; Springer: New York, NY, USA, 1979; Volume 1, pp. 337–354. [Google Scholar] [CrossRef]
  14. Holland, P.M.; Rubingh, D.N. Nonideal multicomponent mixed micelle model. J. Phys. Chem. 1983, 87, 1984–1990. [Google Scholar] [CrossRef]
  15. Holland, P.M. Nonideal mixed micellar solutions. Adv. Colloid Interface Sci. 1986, 26, 111–129. [Google Scholar] [CrossRef]
  16. Poša, M. Mixed micelles of binary surfactant mixtures: Tween 40—Na-3,12-dioxo-5β-cholanoate; Tween 80—Na-3,12-dioxo-5β-cholanoate and their thermodynamic description-characterization. Chem. Eng. Res. Des. 2014, 92, 1826–1839. [Google Scholar] [CrossRef]
  17. Shiloach, A.; Blankschtein, D. Prediction of critical micelle concentration and synergism of binary surfactans mixtures containing zwitterionic surfactants. Langmuir 1997, 13, 3968–3981. [Google Scholar] [CrossRef]
  18. Shiloach, A.; Blankschtein, D. Predicting micellar solution properties of binary surfactants mixtures. Langmuir 1998, 14, 1618–1636. [Google Scholar] [CrossRef]
  19. Nagarajan, R. Molecular theory for mixed micelles. Langmuir 1985, 1, 331–341. [Google Scholar] [CrossRef]
  20. Guggenheim, E.A. Mixtures; The Clarendon Press: Oxford, UK, 1952; pp. 23–46. [Google Scholar]
  21. Poša, M.; Pilipović, A.; Popović, K.; Kumar, D. Thermodynamics of trimethyltetradecylammonium bromide—Sodium deoxycholate binary mixed micelle formation in aqueous solution: Regular Solution Theory with mutual compensation of excess configurational and excess conformational entropy. J. Mol. Liq. 2022, 360, 119473. [Google Scholar] [CrossRef]
  22. Poša, M. Symmetry (asymmetry) of the molar excess Gibbs free energy function of the binary mixed micelles of bile acid anion and classical cationic surfactant: Influence of sterically shielded and sterically unshielded polar groups of the steroid skeleton. Symmetry 2022, 14, 2337. [Google Scholar] [CrossRef]
  23. Bagheri, A.; Khalili, P. Synergism between and cationic surfactants in a concentration range of mixed monolayers at an air–water interface. RSC Adv. 2017, 7, 18151–18161. [Google Scholar] [CrossRef]
  24. Bagheri, A.; Jafari-Chashmi, P. Study of aggregation behavior between N-lauryl sarcosine sodium and Dodecyltrimethylammonium bromide in aqueous solution, using conductometric and spectrophotometric techniques. J. Mol. Liq. 2019, 282, 466–473. [Google Scholar] [CrossRef]
  25. Maeda, H. Phenomenological approaches in the thermodynamics of mixed micelles with electric charge. Adv. Colloid Interf. Sci. 2010, 156, 70–82. [Google Scholar] [CrossRef]
  26. Hoffmann, H.; Pössnecker, G. The mixing behavior of surfactants. Langmuir 1994, 10, 381–389. [Google Scholar] [CrossRef]
  27. Lozar, J. Thermodynamique des Solutions et des Mélanges; Elipses: Paris, France, 2013; pp. 100–116. [Google Scholar]
  28. Ganguly, J. Thermodynamic modelling of solid solutions. EMU Notes Mineral. 2001, 3, 37–69. [Google Scholar]
  29. Straikov, B.E.; Norden, B. Enthalpy-entropy compensation: A phantom or something useful? J. Phys. Chem. B 2007, 111, 14431–14435. [Google Scholar] [CrossRef]
  30. Kumar, D.; Poša, M. Linear hydrophobic congeneric groups of bile acid anion derivatives based on the self-association (micellization) process and the phenomenon of enthalpy-entropy compensation. J. Mol. Liq. 2023, 382, 121925. [Google Scholar] [CrossRef]
  31. Poša, M.; Pilipović, A.; Tepavčević, V.; Obradović, S. Micellisation in water solutions of binary mixture of surfactants Na-deoxycholate—Na-decyl-sulphate and Na-hyodeoxycholate—Na-decyl-sulphate: Rational development (selection) of the thermodynamic model for describing GE. J. Chem. Eng. Data 2018, 63, 691–701. [Google Scholar] [CrossRef]
  32. Pilipović, A.; Ocokoljić, M.; Janev, M.; Poša, M. The ternary mixed micelle of Tween 20- Sodium Deoxycholate- Sodium Cholate: The molar excess thermodynamic potencials. J. Chem. Thermodyn. 2022, 167, 106695. [Google Scholar] [CrossRef]
  33. Schulz, J.E.P.; Rodriguez, J.L.M.; Minardi, R.M.; Miraglia, D.B.; Schulz, P.C. On the applicability of the regular solution theory to multicomponent systems. J. Surfact Deterg. 2013, 16, 795–803. [Google Scholar] [CrossRef]
  34. Dongowski, G.; Fritzsch, B.; Giessler, J.; Hartl, A.; Kuhlmann, O.; Neubert, R.H.H. The influence of bile salts and mixed micelles on the pharmacokinetics of quinine in rabbits. Eur. J. Pharm. Biopharm. 2005, 60, 147–151. [Google Scholar] [CrossRef] [PubMed]
  35. Wiedmann, T.S.; Kamel, L. Examination of the solubilization of drugs by bile salt micelles. J. Pharm. Sci. 2002, 91, 1743–1764. [Google Scholar] [CrossRef]
  36. Enache, M.; Toader, A.M.; Neacsu, V.; Ionita, G.; Enache, M.I. Spectroscopic investigation of the interaction of the anticancer drug mitoxantrone with sodium taurodeoxycholate (NaTDC) and sodium taurocholate (NaTC) bile salts. Molecules 2017, 22, 1079. [Google Scholar] [CrossRef]
  37. Toader, A.M.; Dascalu, I.; Neacsu, E.I.; Enache, M. Binding interactions of actinomycin D anticancer drug with bile salts micelles. J. Serb. Chem. Soc. 2023, 88, 367–379. [Google Scholar] [CrossRef]
  38. Poša, M. Conformationally rigid and flexible surfactant binary (pseudoternary) mixed micelle with mutual synergistic interaction. J. Mol. Liq. 2021, 334, 116158. [Google Scholar] [CrossRef]
  39. Poša, M.; Pilipović, A.; Torović, L.; Cvejić Hogervorst, J. Co-solubilisation of a binary mixtures of isoflavones in a water micellar solution of sodium cholate or cetyltrimethylammonium bromide: Influence of the micelle structure. J. Mol. Liq. 2019, 273, 134–146. [Google Scholar] [CrossRef]
  40. Kamrath, R.F.; Franses, E.I. Mass-action model of mixed micellization. J. Phys. Chem. 1984, 88, 1642–1648. [Google Scholar] [CrossRef]
  41. Jójárt, B.; Poša, M.; Fisher, B.; Szöri, M.; Farkaš, Z.; Viskolcz, B. Mixed micelles of sodium cholate and sodium dodecylsulphate 1:1 binary mixture at different temperatures—Experimental and theoretical investigations. PLoS ONE 2014, 9, e102114. [Google Scholar] [CrossRef]
  42. Zana, R. Critical micellization concentration of surfactants in aqueous solution and free energy of micellization. Langmuir 1996, 12, 1208–1211. [Google Scholar] [CrossRef]
  43. Schulz, P.C.; Rodríguez, J.L.; Minardi, R.M.; Sierra, M.B.; Morini, M.A. Are the mixtures of homologous surfactants ideal? J. Colloid Interface Sci. 2006, 303, 264–271. [Google Scholar] [CrossRef]
  44. Hore, P.J. Nuclear Magnetic Resonance; Oxford University Press: Oxford, UK, 1995. [Google Scholar]
  45. Sanders, J.K.M.; Hunter, B.K. Modern NMR Spectroscopy; Oxford University Press: Oxford, UK, 1995. [Google Scholar]
  46. Privalov, P.L.; Gill, S.J. The hydrophobic effect: A reappraisal. Pure Appl. Chem. 1989, 61, 1097–1104. [Google Scholar] [CrossRef]
  47. Garidel, P.; Hildebrand, A. Thermodynamic properties of association of colloids. J. Therm. Anal. Calorim. 2005, 82, 483–489. [Google Scholar] [CrossRef]
  48. Garidel, P.; Hildebrand, A.; Neubert, R.; Blume, A. Thermodynamic characterization of bile salt aggregation as a function of temperature and ionic strength using isotermal titration calorimetry. Langmuir 2000, 16, 5267–5275. [Google Scholar] [CrossRef]
  49. Paula, S.; Süs, W.; Tuchtenhagen, J.; Blume, A. Thermodynamics of micelle formation as a function of temperature: A high sensitivity titration calorimetry study. J. Phys. Chem. 1995, 99, 11742–11751. [Google Scholar] [CrossRef]
  50. Anderson, S.L.; Rovnyak, D.; Strein, T.G. Direct measurement of the thermodynamics of chiral recognition in bile salt micelles. Chirality 2016, 28, 290–298. [Google Scholar] [CrossRef]
  51. Vázquez-Gómez, S.; Pilar Vázquez-Tato, M.; Seijas, J.A.; Meijide, F.; de Frutos, S.; Vázquez Tato, J. Thermodynamics of the aggregation of the bile anions of obeticholic and chenodeoxycholic acids in aqueous solution. J. Mol. Liq. 2019, 296, 112092. [Google Scholar] [CrossRef]
  52. Kroflic, A.; Sarac, B.; Bester-Rogac, M. Thermodynamic characterization of 3-[(3- cholamidopropyl)-dimethylammonium]-1-propanesulfonate (CHAPS) micellization using isothermal titration calorimetry: Temperature, salt, and pH dependence. Langmuir 2012, 28, 10363–10371. [Google Scholar] [CrossRef]
  53. Small, D.M. The physical chemistry of cholanic acids. In The Bile Acids: Chemistry, Physiology and Metabolism; Nair, P.P., Kritchevsky, D., Eds.; Plenum Press: New York, NY, USA, 1971; Volume 3, pp. 249–332. [Google Scholar]
  54. Mikov, M.; Fawcett, J.P. Bile Acids; Medishet Publisher: Geneva, Switzerland, 2007. [Google Scholar]
  55. Pártai, L.B.; Sega, M.; Jedlovszky, P. Morphology of bile salts micelles as studied by computer simulation methods. Langmuir 2007, 23, 12322–12328. [Google Scholar] [CrossRef]
  56. Unpleby, H.C.; Williamson, R.C. The efficacy of agents employed to prevent anastomotic recurrence in colorectal carcinoma. Ann. R. Coll. Surg. Engl. 1984, 66, 192–194. [Google Scholar]
  57. Smith, A.R.W.; Lambert, P.A.; Hammond, S.M.; Jessup, C. The differing effects of cetyltrimethylammonium bromide and cetrimide B. P. upon growing cultures of Escherichia Coli NCIB 8277. J. Appl. Bacteriol. 1975, 38, 143–149. [Google Scholar] [CrossRef]
  58. Kasilingam, T.; Thangavelu, C. Biocidal behaviour of (dodecyltrimethylammonium bromide) on carbon steel in well water. Trans. Indian Inst. Met. 2016, 69, 793–803. [Google Scholar] [CrossRef]
  59. Huang, L.; Somasundaran, P. Theoretical model and phase behavior for binary surfactant mixtures. Langmuir 1997, 13, 6683–6688. [Google Scholar] [CrossRef]
  60. Kumari, S.; Chauhan, S.; Umar, A.; Fouad, H.; Shaheer Akhtar, M. Conductometric and fluorescence probe analysis to investigate the interaction between bioactive peptide and bile salts: A micellar state study. Molecules 2022, 27, 7561. [Google Scholar] [CrossRef] [PubMed]
  61. Subuddhi, U.; Mishra, A.K. Micellization of bile salts in aqueous medium: A fluorescence study. Colloids Surf. Biointerfaces 2007, 57, 102–107. [Google Scholar] [CrossRef] [PubMed]
  62. Pal, A.; Pillania, A. Thermodynamic and micellization properties of aqueous cetyltrimethylammonium bromide solution in presence of 1-butyl-2,3-dimethylimidazolium bromide. Fluid Phase Equilib. 2016, 412, 115–122. [Google Scholar] [CrossRef]
  63. Chauhan, S.; Kaur, M.; Kumar, K.; Chauhan, M.S. Study of the effect of electrolyte and temperature on the critical micelle concentration of dodecyltrimethylammonium bromide in aqueous medium. J. Chem. Thermodyn. 2014, 78, 175–181. [Google Scholar] [CrossRef]
  64. Mahbub, S.; Rub, M.A.; Hoque, M.A.; Khana, M.A.; Asiri, A.M. Critical Micelle Concentrations of Sodium Dodecyl Sulfate and Cetyltrimethylammonium Bromide Mixtures in Binary Mixtures of Various Salts at Different Temperatures and Compositions. Russ. J. Phys. Chem. A 2019, 93, 2043–2052. [Google Scholar] [CrossRef]
Figure 1. Surfactants: hexadecyltrimethylammonium bromide (1), dodecyltrimethylammonium bromide (2), and sodium deoxycholate (3).
Figure 1. Surfactants: hexadecyltrimethylammonium bromide (1), dodecyltrimethylammonium bromide (2), and sodium deoxycholate (3).
Molecules 28 06722 g001
Figure 2. Hess’s law for formation of ternary mixed micelles: structurally different surfactants as monomers from aqueous solution (A) form monocomponent micelles as separated monocomponent micellar pseudophases (B) from which ternary mixed micellar pseudophase (C) is formed by mixing; process between states (A) and (B) corresponds to the direct association of surfactant monomers ( x 1 f g M 1 0 + x 2 f g M 2 0 + x 3 f g M 3 0 = g ( A B ); g m i x i d + g m i x E = g ( B C ) and f g m M 0 = g ( A C ) ).
Figure 2. Hess’s law for formation of ternary mixed micelles: structurally different surfactants as monomers from aqueous solution (A) form monocomponent micelles as separated monocomponent micellar pseudophases (B) from which ternary mixed micellar pseudophase (C) is formed by mixing; process between states (A) and (B) corresponds to the direct association of surfactant monomers ( x 1 f g M 1 0 + x 2 f g M 2 0 + x 3 f g M 3 0 = g ( A B ); g m i x i d + g m i x E = g ( B C ) and f g m M 0 = g ( A C ) ).
Molecules 28 06722 g002
Figure 3. Change of specific conductivity as function of total surfactant mixture concentration (T = 293.1 K, α1 = 0.1, α2 = 0.3, α3 = 0.6).
Figure 3. Change of specific conductivity as function of total surfactant mixture concentration (T = 293.1 K, α1 = 0.1, α2 = 0.3, α3 = 0.6).
Molecules 28 06722 g003
Figure 4. 2D ROESY spectrum of binary mixed micelles (1)–(3) in 1: 1 molar ratio in aqueous solution above critical micellar concentration; T = 293.1 K: there are cross-peaks between the proton group from surfactant (1) (protons from the C4-C11 hydrocarbon segment methylene groups) and the proton group from (3) surfactant from the C21 steroid skeleton side chain methyl group and protons from C18 and C19 angular methyl groups of the steroid ring system—region A; in region B, there are cross peaks between C21 protons from (3) and protons from methyl groups bound to the (1) surfactant quaternary nitrogen atom.
Figure 4. 2D ROESY spectrum of binary mixed micelles (1)–(3) in 1: 1 molar ratio in aqueous solution above critical micellar concentration; T = 293.1 K: there are cross-peaks between the proton group from surfactant (1) (protons from the C4-C11 hydrocarbon segment methylene groups) and the proton group from (3) surfactant from the C21 steroid skeleton side chain methyl group and protons from C18 and C19 angular methyl groups of the steroid ring system—region A; in region B, there are cross peaks between C21 protons from (3) and protons from methyl groups bound to the (1) surfactant quaternary nitrogen atom.
Molecules 28 06722 g004
Figure 5. Cross-section of the mixed micelle: deoxycholic acid anion (3) replaces globular conformations of cationic surfactant; region A and B present segments of hydrocarbon chain whose protons give cross peaks in 2D ROESY spectra.
Figure 5. Cross-section of the mixed micelle: deoxycholic acid anion (3) replaces globular conformations of cationic surfactant; region A and B present segments of hydrocarbon chain whose protons give cross peaks in 2D ROESY spectra.
Molecules 28 06722 g005
Figure 6. Intermolecular interactions in ternary mixed micelles that do not exist in monocomponent micelles: A- cation–anion attractive interaction and B- dipole–cation attractive interaction.
Figure 6. Intermolecular interactions in ternary mixed micelles that do not exist in monocomponent micelles: A- cation–anion attractive interaction and B- dipole–cation attractive interaction.
Molecules 28 06722 g006
Table 1. Critical micelle concentrations, composition, coefficients of activity, and thermo-dynamic parameters of ternary mixed micelle hexadecyltrimethylammonium bromide (1)—dodecyltrimethylammonium bromide (2)—sodium deoxycholate (3).
Table 1. Critical micelle concentrations, composition, coefficients of activity, and thermo-dynamic parameters of ternary mixed micelle hexadecyltrimethylammonium bromide (1)—dodecyltrimethylammonium bromide (2)—sodium deoxycholate (3).
T
/K
CMC123
/mmolkg−1
x1x2x3lnf1lnf2lnf3 g m i x E
/kJmol−1
f g m M 0 (17)/kJmol−1 g m i x E f g m M 0
α1 = 0.05; α2 = 0.35; α3 = 0.6
278.10.51280.330.280.39−2.64−2.56−2.85−6.25−17.520.35
283.10.53620.320.290.39−2.53−2.53−2.81−6.22−17.730.35
288.10.54710.310.290.39−2.55−2.56−2.76−6.31−17.990.35
293.10.53620.310.290.40−2.64−2.66−2.73−6.54−18.350.36
298.10.55950.300.310.39−2.58−2.74−2.81−6.74−18.560.36
303.10.57040.300.310.39−2.78−2.72−2.81−6.99−18.820.37
308.10.61560.310.290.39−2.78−2.85−2.82−7.22−18.940.38
313.10.73070.320.310.38−2.62−2.86−3.03−7.41−18.800.39
α1 = 0.1; α2 = 0.3; α3 = 0.6
278.10.22510.330.150.52−2.93−3.08−1.72−5.36−19.420.27
283.10.23920.300.170.53−2.73−3.06−1.51−5.02−19.630.26
288.10.26070.290.180.53−2.63−3.06−1.49−5.03−19.760.25
293.10.28850.200.250.56−1.90−3.30−1.31−4.69−19.860.23
298.10.38560.210.270.52−1.80−3.10−1.56−5.02−19.480.26
303.10.42960.220.270.51−1.95−2.99−1.59−5.16−19.540.26
308.10.51070.220.290.49−1.72−3.10−1.93−5.68−19.420.29
313.10.58810.230.290.48−1.68−3.06−2.00−5.82−19.360.30
α1 = 0.2; α2 = 0.2; α3 = 0.6
278.10.18110.580.110.31−1.71−1.57−4.58−5.96−19.920.30
283.10.14250.570.140.29−1.59−1.45−4.75−5.84−20.840.25
288.10.18580.570.110.33−1.83−1.66−4.32−6.28−20.580.31
293.10.21290.510.020.46−2.82−2.65−2.86−6.91−20.600.34
298.10.24550.490.010.50−3.24−3.06−2.51−7.14−20.600.35
303.10.28430.460.050.49−3.26−2.99−2.49−7.22−20.570.35
308.10.32250.390.120.48−2.66−3.13−2.24−6.44−20.590.31
313.10.37720.350.140.51−2.23−3.15−1.83−5.61−20.520.27
α1 = 0.3; α2 = 0.1; α3 = 0.6
278.10.22080.350.160.48−3.33−2.35−2.92−6.88−19.460.35
283.10.23120.350.160.49−3.53−2.32−2.95−7.18−19.710.36
288.10.24440.320.190.48−3.42−2.30−2.88−7.04−19.920.35
293.10.2607n.d.n.d.n.d.n.d.n.d.n.d.n.d.−20.11n.d.
298.10.27390.390.110.50−3.34−2.49−2.87−7.47−20.330.37
303.10.25870.390.120.49−3.15−2.60−2.68−7.19−20.810.35
308.10.4113n.d.n.d.n.d.n.d.n.d.n.d.n.d.−19.97n.d.
313.10.47570.310.230.46−2.31−2.22−2.97−6.75−19.920.34
α1 = 0.35; α2 = 0.05; α3 = 0.6
278.10.22720.550.090.35−3.45−3.10−4.76−8.98−19.390.46
283.10.19910.550.030.42−2.95−2.55−3.93−7.88−20.060.39
288.10.21490.600.020.38−3.14−2.80−4.63−8.86−20.230.44
293.10.25570.610.020.37−2.99−2.72−4.67−8.79−20.150.44
298.10.25940.590.030.38−2.80−2.63−4.24−8.30−20.460.41
303.10.28650.500.110.39−3.36−3.07−3.73−8.75−20.560.43
308.10.27370.540.030.43−2.99−3.01−3.28−7.98−21.010.38
313.10.38050.540.010.45−3.03−2.57−3.08−7.93−20.490.39
Molar Gibbs energy of formation of ternary micellar pseudophase ( f g m M 0 ) is calculated from Equation (17); αi presents molar fraction of surfactants in starting mixtures of surfactants dissolved in aqueous solution; the relative standard uncertainty of the critical micelle concentrations 4%.
Table 2. The origin and information of used chemicals.
Table 2. The origin and information of used chemicals.
CompoundOriginCAS NumberPurity
Hexadecyltrimethylammonium bromideAlfa Aesar57-09-0>0.980
Dodecyltrimethylammonium bromideAlfa Aesar1119-94-4>0.980
Sodium DeoxycholateAlfa Aesar302-95-4>0.980
PyreneAldrich129-00-0>0.980
Table 3. Molar ratios of measured mixtures of surfactants hexadecyltrimethylammonium bromide (1)—dodecyltrimethylammonium bromide (2)—sodium deoxycholate (3); ternary mixtures and corresponding secondary mixtures.
Table 3. Molar ratios of measured mixtures of surfactants hexadecyltrimethylammonium bromide (1)—dodecyltrimethylammonium bromide (2)—sodium deoxycholate (3); ternary mixtures and corresponding secondary mixtures.
Ternary Mixture
1:2:3
Corresponding Binary Mixtures
1:21:32:3
0.5:3.5:0.60.125:0.8750.077:0.9230.368:0.632
1:3:60.25:0.750.143:0.8570.33:0.67
2:2:60.5:0.50.25:0.750.25:0.75
3:1:60.75:0.250.33:0.670.25:0.75
3.5:0.5:60.875:0.1250.368:0.6320.125:0.875
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Pilipović, A.; Vapa, I.; Tepavčević, V.; Puača, G.; Poša, M. Ternary Mixed Micelle Hexadecyltrimethylammonium Bromide—Dodecyltrimethylammonium Bromide—Sodium Deoxycholate: Gibbs Free Energy of Mixing and Excess Gibbs Energy of Mixing. Molecules 2023, 28, 6722. https://doi.org/10.3390/molecules28186722

AMA Style

Pilipović A, Vapa I, Tepavčević V, Puača G, Poša M. Ternary Mixed Micelle Hexadecyltrimethylammonium Bromide—Dodecyltrimethylammonium Bromide—Sodium Deoxycholate: Gibbs Free Energy of Mixing and Excess Gibbs Energy of Mixing. Molecules. 2023; 28(18):6722. https://doi.org/10.3390/molecules28186722

Chicago/Turabian Style

Pilipović, Ana, Ivana Vapa, Vesna Tepavčević, Gorana Puača, and Mihalj Poša. 2023. "Ternary Mixed Micelle Hexadecyltrimethylammonium Bromide—Dodecyltrimethylammonium Bromide—Sodium Deoxycholate: Gibbs Free Energy of Mixing and Excess Gibbs Energy of Mixing" Molecules 28, no. 18: 6722. https://doi.org/10.3390/molecules28186722

Article Metrics

Back to TopTop