Next Article in Journal
Preparation and Characterization of a Novel Mucoadhesive Carvedilol Nanosponge: A Promising Platform for Buccal Anti-Hypertensive Delivery
Next Article in Special Issue
Incorporation of Natural and Recombinant Collagen Proteins within Fmoc-Based Self-Assembling Peptide Hydrogels
Previous Article in Journal
Circular Economy of Coal Fly Ash and Silica Geothermal for Green Geopolymer: Characteristic and Kinetic Study
Previous Article in Special Issue
Effective Carbon/TiO2 Gel for Enhanced Adsorption and Demonstrable Visible Light Driven Photocatalytic Performance
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synergistic Interaction and Binding Efficiency of Tetracaine Hydrochloride (Anesthetic Drug) with Anionic Surfactants in the Presence of NaCl Solution Using Surface Tension and UV–Visible Spectroscopic Methods

1
Center of Excellence for Advanced Materials Research, King Abdulaziz University, Jeddah 21589, Saudi Arabia
2
Chemistry Department, Faculty of Science, King Abdulaziz University, Jeddah 21589, Saudi Arabia
*
Author to whom correspondence should be addressed.
Gels 2022, 8(4), 234; https://doi.org/10.3390/gels8040234
Submission received: 15 March 2022 / Revised: 2 April 2022 / Accepted: 7 April 2022 / Published: 11 April 2022
(This article belongs to the Special Issue Advance in Composite Gels)

Abstract

:
Surfactants are ubiquitous materials that are used in diverse formulations of various products. For instance, they improve the formulation of gel by improving its wetting and rheological properties. Here, we describe the effects of anionic surfactants on an anesthetic drug, tetracaine hydrochloride (TCH), in NaCl solution with tensiometry and UV–visible techniques. Various micellar, interfacial, and thermodynamic parameters were estimated. The outputs were examined by using different theoretical models to attain a profound knowledge of drug–surfactant mixtures. The presence of attractive interactions among drug and surfactant monomers (synergism) in mixed micelle was inferred. However, it was found that sodium dodecyl sulfate (SDS) showed greater interactions with the drug in comparison to sodium lauryl sarcosine (SLS). The binding of the drug with surfactants was monitored with a spectroscopic technique (UV–visible spectra). The results of this study could help optimize the compositions of these mixed aggregates and find the synergism between monomers of different used amphiphiles.

1. Introduction

It is often observed that the surfactant mixtures (e.g., surfactant–co-polymer, surfactant–drug, and surfactant–surfactant) exhibit better performance than single surfactants [1,2,3,4,5,6]. It is also common to use mixtures of surfactants and polymers to formulate gels that are used in drug-dosage forms to improve their properties or to improve their physical stability [7]. The anionic surfactant used in this study, sodium dodecyl sulfate, has been used to synthesize nanogels [8]. SDS has shown better activity in the formation of microgels based on poly(N-isopropylacrylamide) [8]. The synergistic or antagonistic effects of binary mixtures are produced by attraction or repulsion between surfactant monomers. Synergism is observed when the molecular interaction between the monomers of a mixture is greater than before mixing. The strength of synergism between different types of surfactants follows the order of anionic–cationic > nonionic–ionic > ionic–ionic > nonionic–nonionic. The interaction between oppositely charged head groups and the hydrophobic interaction between chains of amphiphiles are the two main factors that are responsible for strong synergistic effects inside cationic–anionic mixtures [9,10,11]. Ionic–anionic mixtures become turbid (precipitation) at some mole fractions, producing lamellar phases and rod-like morphologies.
A lesser water solubility and the dissolution characteristics of a drug usually limit its bioavailability and therapeutic efficacy. The poor water-solubility of drugs may also lead to disappointing and inconstant ingesting, which aggravates the complications of bioavailability and scarcity in the delivery of drugs. In addition, excessive dosages of drugs cause side effects such as vomiting, nausea, dizziness, and fatigue [12,13]. The development of increasing water solubility and improvements in encapsulation efficiency can enhance absorption, enhance bioavailability, and lower the required therapeutic dose [1,14,15]. Researchers have often studied different ways to increase solubilities, such as using small drug carriers, preparing nanoparticles, and using self-emulsifying formulations or amorphous formulations based on water-soluble polymers. A surfactant is a most-capable drug transporter in biomedical applications since it can be easily fabricated into different formulations such as micelles, hydrogels, and nanoparticles to enclose bioactive agents at several points of hydrophobicity [16,17,18,19]. Surfactants are polar molecules and contain both hydrophilic and hydrophobic components orientated at the surface to diminish the surface tension of water [20,21,22]. A micelle will only form when the concentration of the amphiphile is higher than a specific concentration (called the critical micelle concentration or cmc) that can be determined using diverse methods (surface tension, conductometry, fluorometry, UV–visible spectroscopy, cyclic voltammetry, and isothermal calorimetry) [23,24,25,26]. A valuable feature of these molecules is their cmc value. The cmc value depends on various aspects such as ionic strength, temperature, and the existence of additional compounds in the solution. Most chemical industries utilize surfactants, e.g., as pharmaceuticals, corrosion inhibitors, detergents, paints, and cosmetics [27,28,29,30,31].
Certain types of drugs, such as antidepressants, anticholinergics, antihistamines, and local anesthetics, are amphiphilic; they have surfactant-like properties and form micelles [32,33,34,35]. Invariably, their therapeutic activity is determined by how they interact with surfactants. Depending on their interactions in solution, any drug can be made more active. The mixed systems of many amphiphilic drugs have also been researched by our group using different techniques with different amphiphiles [36,37,38,39,40,41,42,43,44,45]. Tetracaine hydrochloride, TCH (Figure 1), is an amphiphilic compound that also possesses colloidal properties and is one of the most used local anesthetic drugs. It is used for stopping pain during surgery and eye infections. Since tetracaine is a poorly water-soluble compound, it is usually formulated as tetracaine hydrochloride. It has been hypothesized that the +ve charge on the drug, which is the functional component, interacts with the Na+ channels on neuronal membranes and stops the transmission of the pain sensation along the nerve [46,47]. Furthermore, the cationic form provides an amphiphilic structure to such a drug, so it can be classified as a cationic tension-active molecule. Therefore, a TCH-like cationic surfactant undergoes an abrupt change above a critical concentration (cmc) and the Krafft temperature. The aqueous dissolution of tetracaine follows the same principle as all ionic surfactants (in that it is governed by both solubility and micellization). As a result, the nature of the surfactant, its counter ions, concentration, and temperature all affect the process. As the use of high concentrations of local anesthetic in spinal anesthesia is known to occasionally result in the sudden death of patients, it is important to understand how the micellization process occurs and what its phase diagram looks like.
In this work, surface tension and UV–visible measurements were carried out to examine the effects of anionic surfactants on a cationic drug. To the best of our knowledge, the mixed micellization of tetracaine hydrochloride (TCH) with sodium lauroyl sarcosine (SLS) and sodium dodecyl sulfate (SDS) in the presence of sodium chloride (NaCl) has not been previously described. Different theoretical approaches of mixed micellization (such as those by Clint, Rubingh, Rodenas, Rosen, and Motomura) were utilized to investigate the interactions of TCH + SDS/SLS mixtures. Various interfacial, micellization, and energetic parameters were analyzed. The output of this work can support the search for a surfactant-based carrier for drug delivery.

2. Result and Discussion

The stock solutions of numerous mole fractions ( α 1 ) of component 1 (SDS/SLS) from 0 to 1 were prepared. As shown in Figure 2, the solution was turbid at some mole fractions (which barred the experiment), and we selected the mole fractions where no turbidity was observed. The surface tension (ST) measurements were used to estimate the cmc values of pure and binary mixtures of drugs and surfactants. Measurements of surface tension are widely used to provide authentic cmc values for all types of surfactants (cationic, anionic, and non-ionic). Illustrative ST graphs for the mixtures at different mole fractions of SLS in the presence of 100 mM NaCl at 298.15 K are displayed in Figure 3. The cmc values acquired via surface tension are listed in Table 1. As the surfactant molecules were mixed, a complex, which was more deeply adsorbed at the surface than single amphiphiles, was formed, thus suggesting an enhanced surface activity. The cmc values of single and mixed amphiphiles could be evaluated by the intersection of the linear fitting of the points (Figure 3). The cmc value of TCH was found to be 79.43 mM, which was lower than the values published by Miller et al. [48], who reported a value of nearly 100 mM without any salt. The cmc values of both employed surfactants in the existence of salt were also found to be less than those with a lack of salt. The values of cmc for currently employed surfactants in the presence of NaCl were in good agreement with the literature [49,50]. The obtained value of cmc for SDS in the presence of 100 mM NaCl was much lower than the cmc value computed by Thapa et al. [51] in an aqueous solution. When NaCl was added to the drug solution, the electrical atmosphere changed. The charge between the head group in the cationic drug became neutralized. Micelles could be formed at much lower concentrations in pure water because of the reduced electrostatic repulsion among the polar head groups. The cmc values for all mixtures unified in the center of two single amphiphiles, suggesting that the micellization of a drug was preferred in the company of surfactants. The observed decline in the cmc values of the mixture was due to the enrichment in the hydrophobic interaction among drugs and surfactants.
The whole study can be divided into two parts: (A) interactions of drugs with surfactants in the solution and (B) interactions of drugs with surfactants at the surface.

2.1. Interactions of Drug with the Surfactants in the Mixed Micelle

Using Rubingh’s regular solution theory (RST) for mixtures of amphiphiles [52], the cmc of a mixed system (cmc*) can be calculated via Equation (1):
1 c m c * = α 1 f 1 c m c 1 + α 2 f 2 c m c 2  
where f1 and f2 are the activity coefficients of the surfactant (SDS/SLS) and drug in mixed micelles, respectively, and α 1 represents the mole fraction of surfactant (SDS/SLS) in the total mixed solution. The cmc values of surfactants and drugs are c m c 1 and c m c 2 , respectively. f1 = f2 = 1 if we assume ideal behavior, so Equation (1) becomes:
1 c m c * = α 1 c m c 1 + α 2 c m c 2  
Equation (2) was proposed by Clint [53]. Using the Clint equation, we could judge the ideality or non-ideality of a mixed system. Figure 4 displays a plot of cmc (experimentally determined)/cmc* (calculated with Equation (2)) vs. α1 (SDS/SLS). The cmc values of both mixtures were decreased with increases in the α1. According to one possible explanation, the mixture was more favorable than expected under an ideal condition because of the interactions among hydrophobic chains of amphiphiles.
In contrast, for non-ideal mixtures, a new theory has been established and is referred to as the Rubingh model [52]. The Rubingh model uses RST to relate the activity coefficients of components with micellar mole fractions of component 1 as follows:
f 1 R u b = e x p [ β R u b ( 1 X 1 R u b ) 2 ]
f 2 R u b = e x p [ β R u b ( X 1 R u b ) 2 ]
where β R u b and X 1 R u b are the interaction parameter and micellar mole fraction, respectively of component 1. If two variables have values of less than 1, the mixing components are not ideal. When computing the β R u b values (parameter based on the cmc values of each amphiphile and their mixtures), the nature and strength of the interactions between the two surfactants are determined. Rubingh [52] derived the relationship shown in Equation (5) by considering the phase separation model for micellization.
β R u b = ln ( α 1 c m c / X 1 R u b c m c 1 ) ( 1 X 1 R u b ) 2  
The micellar mole fraction of component 1 is represented by X 1 R u b , which is calculated by iteratively solving Equation (6):
( X 1 R u b ) 2 ln ( α 1 c m c / X 1 R u b c m c 1 ) ( 1 X 1 R u b ) 2 ln [ ( 1 α 1 ) c m c / ( 1 X 1 R u b ) c m c 2 ] = 1
It is commonly believed that the deviation from zero of the interaction parameters ( β R u b ) is due to interactions among the amphiphile head groups. Positive divergence from zero indicates antagonistic behavior, and negative deviation indicates synergistic interactions between two components. Free energy subsidies associated with amphiphile head groups have been found to be the main sources of mutual interaction. When positively and negatively charged amphiphiles are assorted in water, the most noteworthy feature of this mixture is its unusually huge drop in cmc values. A mixture of anionic and non-ionic surfactants usually yields a nonconformity from ideal behavior (less negative β R u b ) and synergistic effects in the mixed micelles of two non-ionic amphiphiles are even to a lesser extent. In most cases, experimentally computed values of β R u b for mixtures of positively and negatively charged amphiphiles are higher. According to Table 1, there were considerable interactions (synergism) between the current mixed systems. The synergism was detected because of the electrostatic interaction among +ve and –ve charged head groups. The β R u b average values were – 15.90 and – 11.25 for SDS + TCH and SLS + TCH, respectively. The positive and negatively charged amphiphiles were found to be firmly tied to one another through electrostatic and hydrophobic forces, consequently leading to ultimate attraction that promoted the growth of micellar aggregates. The synergism between two amphiphiles depends not only on the strength of the interaction but also on the individual amphiphile properties. The higher the hydrophobicity of an amphiphile, the easier it is to make micelles.
In a mixed system, the ideal micellar mole fraction of component 1 is represented by Equation (7) [54]
X 1 i d e a l = α 1 c m c 2 α 1 c m c 2 + α 2 c m c 1
The values of X 1 i d e a l are given in Table 1. The values of X 1 i d e a l display nonconformity from the values of X 1 R u b , signifying non-ideality. The higher values of X 1 i d e a l for both binary mixtures at all mole fractions confirmed that added drug molecules replace some of the surfactant molecules from the mixed micelles, so the contribution of drug molecules is greater in mixed micelles than it should be in ideally mixed systems.

Thermodynamic Parameters for Drug–Surfactant Mixtures in the Mixed Micelle

Using RST, it is feasible to evaluate the free energy change for micellization in the following way [55,56,57,58,59]:
Δ G m i x = R T [ X 1 R u b l n ( X 1 R u b f 1 R u b ) + X 2 R u b l n ( X 2 R u b f 2 R u b ) ]
If the values of activity coefficients ( f 1 R u b and f 2 R u b ) for an ideal mixed system are equal to unity, then Equation (8) becomes:
Δ G m i x i d e a l = R T [ X 1 R u b l n X 1 R u b + X 2 R u b l n X 2 R u b ]  
where Δ G m i x i d e a l is the free energy change for an ideal mixed system. Interestingly, the data (Table 2) show that the values were negative, implying that the micelles were spontaneously formed and were stable. If the values of Δ G m i x i d e a l deviate from the values of Δ G m i x , rather than forming an ideal micelle, it then forms a real one. The literature confirms that previous investigators have observed the same behavior [60,61].
An excess thermodynamic function is a variation among the energetic function of the mixer for a non-ideal solution and the subsequent values for an ideal solution at a similar pressure and temperature [54]. The excess free energy of mixed micellization G m i x E for a two-amphiphile mixtures can be computed with the help of equations 8 and 9 in form of Equation (10).
G m i x E = Δ H m = R T [ X 1 R u b l n f 1 R u b + X 2 R u b l n f 2 R u b ]  
From Table 2, we can observe that the values of G m i x E were negative over the entire mole fraction range, confirming observations that the creation of the mixed micelles was thermodynamically more stable than the ideal state.
For the mixed system, Equations (9) and (10) were also used to calculate the entropy change as Equation (11):
Δ S m = Δ H m Δ G m T = R [ X 1 R u b l n X 1 R u b + X 2 R u b l n X 2 R u b ]
Moreover, both binary and mixed micellization were found to be constrained by positive entropy values, which confirmed that entropy contribution drives mixed micellization. In the literature, the same results have previously been reported [55]. When we consider SDS + TCH mixed systems, the contributions to entropy were more significant at initial fractions. It was found to be an entropically favorable process when mixed micelles were formed, as the entropy/free energy change in this process was greater than 0.
Equation (12) was utilized to compute standard Gibbs free energy per mole of micellization using the mass-action model without considering counterion binding [58]:
Δ G m o = R T l n X C M C  
In the above equation, X C M C is the cmc value at mole fraction unit while R and T have their basic scientific meaning. The values of Δ G m o listed in Table 2 are negative for single and mixed amphiphiles. The negative values show that the micellization spontaneously occurred in the aqueous NaCl solution. The Δ G m o values of the drug were less than the single surfactants (SDS or SLS) and mixtures, confirming that mixed micelle formation of a drug with surfactants is more spontaneous compared to a drug alone. It is interesting to note here that the β R u b values and Δ G m o values were directly proportional with respect to α 1 , confirming that the higher interactions between amphiphile monomers cause more spontaneity in the process; the same results were reported by Bagheri et al. [54].

2.2. Interfacial Properties of TCH + SDS/SLS Mixed System

When amphiphiles are dissolved in water, the amphiphile monomers are adsorbed at the surface and the surface tension of water decreases, mainly due to the hydrophobic effects. The thermal motion and dynamic equilibrium determine the adsorption or desorption of monomers. Electrostatic interactions, hydrogen bonding, van der Waals interactions, and solvation/desolvation are factors that are less responsible for adsorption. Gibb’s adsorption equation can be used to quantify the amount of amphiphiles adsorbed per unit area of the interface (surface excess, Γ m a x ) [62]:
Γ max = 1 2.303 n R T ( d γ d l o g C )
In Equation (13), d γ d l o g C is the maximum slope, T is the absolute temperature in K, and R = 8.314 J mol–1 K–1. Based on literature, the value of n was taken as 2 for pure amphiphiles and was calculated for mixtures with the following expression [62,63]
n = X 1 s n 1 X 2 s n 2  
The Γ max values can be used to calculate the values of minimum area per molecule ( A m i n ) with Equation (15) [64]
A m i n = 10 20 N A Γ max
where NA = 6.02214 × 1023 (Avogadro’s number). The minimum area per molecule of an amphiphile suggests the packing (loose or close) and orientation of the amphiphile molecule at the surface. The low A m i n (high Γ max ) values of the mixture at all mole fractions confirmed strong electrostatic interactions between cationic drugs and anionic surfactants (Table 3). This fact was also reflected in the negative interaction parameter values for the mixture. If there is no interaction between two amphiphiles in a mixed adsorbed film at the surface, the minimum area per molecule can be calculated with the following equation [62]:
A i d e a l = α 1 A m i n ,   1 + α 2 A m i n ,   2
The observed values ( A m i n ) were lower than ideal values ( A i d e a l ), indicating significant attractive interactions between the two components (Table 3). Water became 84–99% saturated following the adsorption of amphiphiles, which reduced its surface tension by approximately 20 dyn/cm. Adding an amphiphile to the water decreased the surface tension of H2O by 20 mNm−1, indicating the efficiency of its adsorption. Hence, it has the lowest concentration required to achieve saturation adsorption. By using Equation (17), we could calculate the adsorption efficiency ( p C 20 ) as:
p C 20 = l o g C 20
where C20 is a measure of the adsorption efficiency of surfactants at the interface. The values of C20 are also listed in Table 3. It was concluded that the C20 values of SDS decreased with the addition of TCH. Decreasing C20 values of SDS with TCH were also shown by an earlier study [51]. In the case of SLS, the values of C20 only decreased at higher mole fractions. The C20 value of SDS in the presence NaCl has been found to be lower than in its absence [51], confirming that the surface activity of SDS is enhanced in the presence of NaCl.
Rosen and Hua modified Equations (5) and (6) for amphiphile adsorption to calculate the X 1 S and β s with the following equations [64]
( X 1 s ) 2 ln ( α 1 C m i x / X 1 s C 1 ) ( 1 X 1 S ) 2 ln [ ( 1 α 1 ) C m i x / ( 1 X 1 S ) C 2 ] = 1
β s = l n ( α 1 C m i x / X 1 S C 1 ) ( 1 X 1 S ) 2
The interpretation of interaction parameter at the surface ( β s ) is the same as in the case of bulk ( β R u b ) , with negative and positive β s values that suggest synergism and antagonism, respectively. Here, the values of X 1 s were increased with the stoichiometric mole fraction (Table 4) and were always greater than X 1 R u b , showing amphiphiles contributed more to mixed monolayer formation than in the mixed micelle. Additionally, the contribution of SDS was greater than SLS in the mixed monolayer formation with the TCH. The β s values were negative for both mixed systems, suggesting attractive interaction. The activity coefficients at the surface could be calculated by the following equations
l n f 1 S = β s ( X 2 S ) 2
l n f 2 S = β s ( X 1 S ) 2
The values of f 1 S and f 2 S are listed in Table 4 and were found to be less than unity, thus indicating non-ideality at the surface.

Thermodynamic Parameters for Drug–Surfactant Mixtures at the Surface

The standard free energy of interfacial adsorption ( Δ G a d d o ) can be computed by using the following relation [58]:
Δ G a d d o = Δ G m o ( π C M C Γ m a x )  
At the cmc, surface pressure is measured with the term π C M C . Here in Equation (22), G m o is the standard Gibbs free energy previously computed with Equation (12). It was observed that the accomplished upsides of Δ G a d d o were –ve, similar to those of Δ G m o ; nonetheless, the extent was much more noteworthy, showing that adsorption was further unconstrained for this situation. f 1 S and f 2 S can be utilized to ascertain excess free energy ( G e x c s ) at surface:
G e x c s = R T [ X 1 S l n f 1 S + ( 1 X 1 S ) l n f 2 S ]
With negative values of G e x c s , stability can be attained by the stable mixing at the surface, which is possible with the monolayer of surfactants or drugs alone. Negative G e x c s values (Table 4) also indicate synergism at the surface. The degree of synergism for a mixed system can also be quantified by an energy parameter [65],
G m i n s = A m i n γ C M C N A
The energy parameters that define the work required to create an interface per mole of the solution by transferring monomers from bulk to interface can be determined by the above-described energy parameters ( G m i n s ). According to Table 4, a lower value of G m i n s indicates a more stable surface, and this in turn results in increased surface activity.

3. UV–Visible Spectroscopic Study

The interaction of TCH with SDS and SLS was monitored with UV–visible absorption spectroscopy. The absorption spectrum of TCH (0.05 mM) in a 100 mM NaCl solution showed two absorption peaks at 226 and 310 nm due to the attendance of the aminobenzoate group. π–π* and n–π* transitions were involved in the first and second ones, respectively. When increasing concentrations of SDS and SLS were added to the TCH solution, the absorbance increased but the maximum absorbance at 310 nm was not changed (Figure 5). This spectral behavior indicates the electrostatic interactions between the positive charge of TCH molecules and the negative charge of surfactant monomers.
The binding constant and stoichiometric ratio were estimated with the differential absorbance method represented by the Benesi–Hildebrand equation [66]:
1 A A 0 = 1 K ( A m a x A 0 ) [ S ] n + 1 A m a x A 0
where the concentration of SDS/SLS is represented by [S], while A, A0, and Amax represent values of absorbance due to the presence of surfactants, the absence of surfactants, and resulting absorbance due to the drug–surfactant complex, respectively. When plotting 1/(A − A0) against 1/[SDS/SLS]2, a straight line is obtained (Figure 6), specifying the creation of the 1:2 complex. For an SDS + TCH mixed system without the addition of salt, Thapa et. al. reported a 1:1 complex [51]. However, for our system, a curvilinear fit was obtained, so the SDS + TCH complex was mainly 1:2. Using the Benesi–Hildebrand equation, the binding constant could be calculated (intercept/slope). We found values of K of 1.86 × 105 (± 0.04) and 9.09 × 104 (± 0.04) mol–1 dm3 for the SDS + TCH and SLS + TCH mixed systems, respectively. The SLS + TCH mixed system had lesser binding constant values than the SDS + TCH system. In comparison, SDS has one functional group and SLS has two functional groups. The localized positive charge on the nitrogen atom on the TCH interacts with the negative charge on the sulphonic group, thus enhancing the electrostatic attraction between the guest and host. SLS, however, has methylated amide nitrogen, so the amide bond cannot be a hydrogen bond donor, which inhibits intermolecular attraction between SLS and TCH at the palisade layer. Furthermore, the steric hindrance of the N-methyl group of SLS may make it difficult to tightly align the amphiphiles. All these behaviors of SLS are responsible for its lesser binding constant compared to SDS.
By using binding constant (K) values, free energy change of binding could be attained with Equation (26):
Δ G K = R T l n K
The binding free energies were –30.08 (± 0.2) Jmol–1 for SDS + TCH and – 28.30(± 0.2) kJmol–1 for SLS + TCH. In both mixed systems, the G values were negative, indicating that the binding process was spontaneous.

4. Conclusions

The synergistic interaction of TCH (+ve charged head group) with SDS and SLS (–ve charged head group) surfactants in the presence of salt (100 mM NaCl) was analyzed with both tensiometry and UV–visible spectroscopic techniques. The following conclusions can be derived:
  • The negative deviation of experimentally determined cmc values with hypothetical values confirms the nonideality of current mixtures.
  • The interaction parameter at the interface and in solution was determined to be –ve, thus validating synergism between monomers of two species at the surface and in bulk.
  • The higher values of the ideal mole fraction of component 1 ( X 1 i d e a l ) for both binary mixtures at all mole fractions indicate the strong ability of the drug to form of mixed micelles.
  • Energetics parameters confirm the spontaneity, stability, and entropic favorability of drug–surfactant mixtures.
  • The TCH with SLS had smaller binding constant values than SDS, possibly because SLS has a methylated amide nitrogen so the amide bond cannot be a hydrogen bond donor, which inhibits the intermolecular attraction between SLS and TCH at the palisade layer. Furthermore, the steric hindrance of the N-methyl group of SLS may make it difficult to tightly align the amphiphiles. All these behaviors of SLS are responsible for its smaller binding constant in comparison to SDS.

5. Experimental

5.1. Materials

Tetracaine hydrochloride (TCH, 99%), an anesthetic amphiphilic drug, and sodium lauroyl sarcosine (SLS, >95%) were supplied by Molecules On (Switzerland) and used as received. Sodium chloride (NaCl, 99%) and sodium dodecyl sulfate (SDS, 98.5%) were acquired from Sigma-Aldrich (St. Louis, MO, USA). At 298.15 K, all experiments were performed using ultra-pure, double-distilled de-ionized water with a conductivity between 1 and 2 µScm–1. To prepare standard solutions for experiments, amphiphiles (both pure and mixed) were dissolved and accurately weighed in a 100 mM NaCl solution. The stock solutions for both techniques (surface tension and UV–vis spectrophotometer measurements) were prepared in aqueous 100 mM NaCl solutions.

5.2. Methods

5.2.1. Surface Tension Measurements

The surface tension experiments were conducted with a digital tensiometer (Sigma 700, Attention, Darmstadt, Germany) by using a platinum ring. The instrument was occasionally calibrated with ultra-pure distilled water. In tensiometric titration, an amphiphile stock solution was titrated into a static volume of H2O. Throughout all experiments, water was circulated from a thermostatically controlled water bath through the outer jacket to keep the temperature at 298.15 K.

5.2.2. UV–Vis Spectrophotometer Measurements

We measured the spectra of the aqueous solutions of the drug and the drug–surfactant binary mixtures to determine the level of the binding of the drug with surfactants. As a first step, TCH in water was prepared as a stock solution in a volumetric flask. The desired concentration of surfactant solution was prepared from the aqueous TCH solution. Finally, a suitable volume of surfactant solution was added to the H2O solution of TCH in a quartz cell. We measured the absorption spectra of TCH solutions with surfactants and plotted them against the wavelengths. For the measurement of the absorption spectrum of TCH solutions over the range of 200–400 nm, an Evolution 300 spectrophotometer from Thermo Scientific, Tokyo, Japan was used to record UV–visible spectra (Figure 2).

Author Contributions

N.A., M.A.R. & A.K. were involved in experimental planning, interpreting data and writing the manuscript; M.M.A., A.M.A. reviewed and edited the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This research work was funded by Institutional Fund Projects under grant no (IFPRC-174-130-2020). Therefore, the authors gratefully acknowledge technical and financial support from the Ministry of Education and King Abdulaziz University, Jeddah, Saudi Arabia.

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.

References

  1. Van Eerdenbrugh, B.; Vermant, J.; Martens, J.A.; Froyen, L.; Van Humbeeck, J.; Van den Mooter, G.; Augustijns, P. Solubility Increases Associated with Crystalline Drug Nanoparticles: Methodologies and Significance. Mol. Pharm. 2010, 7, 1858–1870. [Google Scholar] [CrossRef]
  2. Ruso, J.M.; Attwood, D.; Rey, C.; Taboada, P.; Mosquera, V.; Sarmiento, F. Light Scattering and NMR Studies of the Self-Association of the Amphiphilic Molecule Propranolol Hydrochloride in Aqueous Electrolyte Solutions. J. Phys. Chem. B 1999, 103, 7092–7096. [Google Scholar] [CrossRef]
  3. Awang, N.; Ismail, A.F.; Jaafar, J.; Matsuura, T.; Junoh, H.; Othman, M.H.D.; Rahman, M.A. Functionalization of polymeric materials as a high performance membrane for direct methanol fuel cell: A review. React. Funct. Polym. 2015, 86, 248–258. [Google Scholar] [CrossRef]
  4. Azum, N.; Rub, M.A.; Khan, A.; Alotaibi, M.M.; Asiri, A.M.; Rahman, M.M. Mixed Micellization, Thermodynamic and Adsorption Behavior of Tetracaine Hydrochloride in the Presence of Cationic Gemini/Conventional Surfactants. Gels 2022, 8, 128. [Google Scholar] [CrossRef] [PubMed]
  5. Rub, M.A.; Azum, N.; Kumar, D.; Asiri, A.M. Interaction of TX-100 and Antidepressant Imipramine Hydrochloride Drug Mixture: Surface Tension, 1H NMR, and FT-IR Investigation. Gels 2022, 8, 159. [Google Scholar] [CrossRef]
  6. Ahmed, M.F.; Abdul Rub, M.; Joy, M.T.R.; Molla, M.R.; Azum, N.; Anamul Hoque, M.; Rub, M.A.; Azum, N.; Kumar, D.; Asiri, A.M.; et al. Influences of NaCl and Na2SO4 on the Micellization Behavior of the Mixture of Cetylpyridinium Chloride + Polyvinyl Pyrrolidone at Several Temperatures. Gels 2022, 8, 62. [Google Scholar] [CrossRef] [PubMed]
  7. Alvarez-Lorenzo, C.; Concheiro, A. Effects of Surfactants on Gel Behavior. Am. J. Drug Deliv. 2003, 1, 77–101. [Google Scholar] [CrossRef]
  8. Wedel, B.; Brändel, T.; Bookhold, J.; Hellweg, T. Role of Anionic Surfactants in the Synthesis of Smart Microgels Based on Different Acrylamides. ACS Omega 2017, 2, 84–90. [Google Scholar] [CrossRef] [PubMed]
  9. Maiti, K.; Mitra, D.; Mitra, R.N.; Panda, A.K.; Das, P.K.; Rakshit, A.K.; Moulik, S.P. Self-Aggregation of Synthesized Novel Bolaforms and Their Mixtures with Sodium Dodecyl Sulfate (SDS) and Cetyltrimethylammonium Bromide (CTAB) in Aqueous Medium. J. Phys. Chem. B 2010, 114, 7499–7508. [Google Scholar] [CrossRef] [PubMed]
  10. Jafari-Chashmi, P.; Bagheri, A. The strong synergistic interaction between surface active ionic liquid and anionic surfactant in the mixed micelle using the spectrophotometric method. J. Mol. Liq. 2018, 269, 816–823. [Google Scholar] [CrossRef]
  11. Mal, A.; Bag, S.; Ghosh, S.; Moulik, S.P. Physicochemistry of CTAB-SDS interacted catanionic micelle-vesicle forming system: An extended exploration. Colloids Surf. A Physicochem. Eng. Asp. 2018, 553, 633–644. [Google Scholar] [CrossRef]
  12. Tozuka, Y.; Imono, M.; Uchiyama, H.; Takeuchi, H. A novel application of α-glucosyl hesperidin for nanoparticle formation of active pharmaceutical ingredients by dry grinding. Eur. J. Pharm. Biopharm. 2011, 79, 559–565. [Google Scholar] [CrossRef]
  13. Shen, S.; Ng, W.K.; Chia, L.; Dong, Y.; Tan, R.B.H. Stabilized Amorphous State of Ibuprofen by Co-Spray Drying With Mesoporous SBA-15 to Enhance Dissolution Properties. J. Pharm. Sci. 2010, 99, 1997–2007. [Google Scholar] [CrossRef] [PubMed]
  14. Sigfridsson, K.; Lundqvist, A.J.; Strimfors, M. Particle size reduction for improvement of oral absorption of the poorly soluble drug UG558 in rats during early development. Drug Dev. Ind. Pharm. 2009, 35, 1479–1486. [Google Scholar] [CrossRef]
  15. Sugano, K.; Okazaki, A.; Sugimoto, S.; Tavornvipas, S.; Omura, A.; Mano, T. Solubility and Dissolution Profile Assessment in Drug Discovery. Drug Metab. Pharmacokinet. 2007, 22, 225–254. [Google Scholar] [CrossRef] [PubMed]
  16. Schreier, S.; Malheiros, S.V.P.; de Paula, E. Surface active drugs: Self-association and interaction with membranes and surfactants. Physicochemical and biological aspects. Biochim. Et Biophys. Acta (BBA)-Biomembr. 2000, 1508, 210–234. [Google Scholar] [CrossRef] [Green Version]
  17. YOKOYAMA, S.; FUJINO, Y.; KAWAMOTO, Y.; KANEKO, A.; FUJIE, T. Micellization of an Aqueous Solution of Piperidolate Hydrochloride in the Presence of Acetylcholine Chloride. Chem. Pharm. Bull. 1994, 42, 1351–1353. [Google Scholar] [CrossRef] [Green Version]
  18. Attwood, D.; Tolley, J.A. Self-association of analgesics in aqueous solution: Association models for codeine, oxycodone, ethylmorphine and pethidine. J. Pharm. Pharmacol. 2011, 32, 761–765. [Google Scholar] [CrossRef]
  19. Kumar, D.; Azum, N.; Rub, M.A.; Asiri, A.M. Interfacial and spectroscopic behavior of phenothiazine drug/bile salt mixture in urea solution. Chem. Pap. 2021, 75, 3949–3956. [Google Scholar] [CrossRef]
  20. Ghosh, S.; Krishnan, A.; Das, P.K.; Ramakrishnan, S. Determination of Critical Micelle Concentration by Hyper-Rayleigh Scattering. J. Am. Chem. Soc. 2003, 125, 1602–1606. [Google Scholar] [CrossRef]
  21. Zhu, Q.; Huang, L.; Su, J.; Liu, S. A sensitive and visible fluorescence-turn-on probe for the CMC determination of ionic surfactants. Chem. Commun. 2014, 50, 1107–1109. [Google Scholar] [CrossRef] [PubMed]
  22. Tadros, T.F. Applied Surfactants; Wiley: Hoboken, NJ, USA, 2005; ISBN 9783527306299. [Google Scholar]
  23. Chiu, Y.C.; Kuo, C.Y.; Wang, C.W. Using electrophoresis to determine zeta potential of micelles and critical micelle concentration. J. Dispers. Sci. Technol. 2000, 21, 327–343. [Google Scholar] [CrossRef]
  24. Priev, A.; Zalipsky, S.; Cohen, R.; Barenholz, Y. Determination of Critical Micelle Concentration of Lipopolymers and Other Amphiphiles: Comparison of Sound Velocity and Fluorescent Measurements. Langmuir 2002, 18, 612–617. [Google Scholar] [CrossRef]
  25. Romani, A.P.; da Hora Machado, A.E.; Hioka, N.; Severino, D.; Baptista, M.S.; Codognoto, L.; Rodrigues, M.R.; de Oliveira, H.P.M. Spectrofluorimetric Determination of Second Critical Micellar Concentration of SDS and SDS/Brij 30 Systems. J. Fluoresc. 2009, 19, 327–332. [Google Scholar] [CrossRef]
  26. Pérez-Rodríguez, M.; Prieto, G.; Rega, C.; Varela, L.M.; Sarmiento, F.; Mosquera, V. A Comparative Study of the Determination of the Critical Micelle Concentration by Conductivity and Dielectric Constant Measurements. Langmuir 1998, 14, 4422–4426. [Google Scholar] [CrossRef]
  27. Karsa, D.R. Industrial Applications of Surfactants; Elsevier: Amsterdam, The Netherlands, 1999. [Google Scholar]
  28. Atta, A.M.; Abdullah, M.M.S.; Al-Lohedan, H.A.; Ezzat, A.O. Demulsification of heavy crude oil using new nonionic cardanol surfactants. J. Mol. Liq. 2018, 252, 311–320. [Google Scholar] [CrossRef]
  29. Shaban, S.M.; Kang, J.; Kim, D.-H. Surfactants: Recent advances and their applications. Compos. Commun. 2020, 22, 100537. [Google Scholar] [CrossRef]
  30. Hegazy, M.A.; Abdallah, M.; Ahmed, H. Novel cationic gemini surfactants as corrosion inhibitors for carbon steel pipelines. Corros. Sci. 2010, 52, 2897–2904. [Google Scholar] [CrossRef]
  31. Torchilin, V.P. Structure and design of polymeric surfactant-based drug delivery systems. J. Control. Release 2001, 73, 137–172. [Google Scholar] [CrossRef]
  32. King, S.-Y.P.; Basista, A.M.; Torosian, G. Self-Association and Solubility Behaviors of a Novel Anticancer Agent, Brequinar Sodium. J. Pharm. Sci. 1989, 78, 95–100. [Google Scholar] [CrossRef]
  33. Matsuki, H.; Hashimoto, S.; Kaneshina, S.; Yamanaka, M. Surface Adsorption and Volume Behavior of Local Anesthetics. Langmuir 1994, 10, 1882–1887. [Google Scholar] [CrossRef]
  34. Atherton, A.D.; Barry, B.W. Photon correlation spectroscopy of surface active cationic drugs. J. Pharm. Pharmacol. 2011, 37, 854–862. [Google Scholar] [CrossRef] [PubMed]
  35. Sarmiento, F.; López-Fontán, J.L.; Prieto, G.; Mosquera, V.; Attwood, D. Mixed micelles of structurally related antidepressant drugs. Colloid Polym. Sci. 1997, 275, 1144–1147. [Google Scholar] [CrossRef]
  36. Rub, M.A.; Azum, N.; Khan, F.; Asiri, A.M. Surface, micellar, and thermodynamic properties of antidepressant drug nortriptyline hydrochloride with TX-114 in aqueous/urea solutions. J. Phys. Org. Chem. 2017, 30, e3676. [Google Scholar] [CrossRef]
  37. Abdul Rub, M.; Azum, N.; Asiri, A.M. Binary Mixtures of Sodium Salt of Ibuprofen and Selected Bile Salts: Interface, Micellar, Thermodynamic, and Spectroscopic Study. J. Chem. Eng. Data 2017, 62, 3216–3228. [Google Scholar] [CrossRef]
  38. Azum, N.; Naqvi, A.Z.; Rub, M.A.; Asiri, A.M. Multi-technique approach towards amphiphilic drug-surfactant interaction: A physicochemical study. J. Mol. Liq. 2017, 240, 189–195. [Google Scholar] [CrossRef]
  39. Azum, N.; Rub, M.A.; Asiri, A.M.; Bawazeer, W.A. Micellar and interfacial properties of amphiphilic drug–non-ionic surfactants mixed systems: Surface tension, fluorescence and UV–vis studies. Colloids Surf. A Physicochem. Eng. Asp. 2017, 522, 183–192. [Google Scholar] [CrossRef]
  40. Kumar, D.; Rub, M.A.; Azum, N.; Asiri, A.M. Mixed micellization study of ibuprofen (sodium salt) and cationic surfactant (conventional as well as gemini). J. Phys. Org. Chem. 2018, 31, e3730. [Google Scholar] [CrossRef]
  41. Khan, F.; Rub, M.A.; Azum, N.; Asiri, A.M. Mixtures of antidepressant amphiphilic drug imipramine hydrochloride and anionic surfactant: Micellar and thermodynamic investigation. J. Phys. Org. Chem. 2018, 31, e3812. [Google Scholar] [CrossRef]
  42. Azum, N.; Rub, M.A.; Asiri, A.M. Interaction of antipsychotic drug with novel surfactants: Micellization and binding studies. Chin. J. Chem. Eng. 2018, 26, 566–573. [Google Scholar] [CrossRef]
  43. Kumar, D.; Azum, N.; Rub, M.A.; Asiri, A.M. Aggregation behavior of sodium salt of ibuprofen with conventional and gemini surfactant. J. Mol. Liq. 2018, 262, 86–96. [Google Scholar] [CrossRef]
  44. 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]
  45. Azum, N.; Ahmed, A.; Rub, M.A.; Asiri, A.M.; Alamery, S.F. Investigation of aggregation behavior of ibuprofen sodium drug under the influence of gelatin protein and salt. J. Mol. Liq. 2019, 290, 111187. [Google Scholar] [CrossRef]
  46. Srivastava, A.; Thapa, U.; Saha, M.; Jalees, M. Aggregation behaviour of tetracaine hydrochloride with Gemini surfactants and the formation of silver nanoparticles using drug-Gemini surfactants mixture. J. Mol. Liq. 2019, 276, 399–408. [Google Scholar] [CrossRef]
  47. Zhou, S.; Huang, G.; Chen, G. Synthesis and biological activities of local anesthetics. RSC Adv. 2019, 9, 41173–41191. [Google Scholar] [CrossRef] [Green Version]
  48. Miller, K.J.; Goodwin, S.R.; Westermann-Clark, G.B.; Shah, D.O. Importance of molecular aggregation in the development of a topical local anesthetic. Langmuir 1993, 9, 105–109. [Google Scholar] [CrossRef]
  49. Ray, G.B.; Ghosh, S.; Moulik, S.P. Physicochemical Studies on the Interfacial and Bulk Behaviors of Sodium N-Dodecanoyl Sarcosinate (SDDS). J. Surfactants Deterg. 2009, 12, 131–143. [Google Scholar] [CrossRef]
  50. Umlong, I.M.; Ismail, K. Micellization behaviour of sodium dodecyl sulfate in different electrolyte media. Colloids Surf. A Physicochem. Eng. Asp. 2007, 299, 8–14. [Google Scholar] [CrossRef]
  51. Thapa, U.; Kumar, M.; Chaudhary, R.; Singh, V.; Singh, S.; Srivastava, A. Binding behaviour of hydrophobic drug tetracaine hydrochloride used as organic counterion on ionic surfactants. J. Mol. Liq. 2021, 335, 116564. [Google Scholar] [CrossRef]
  52. Holland, P.M.; Rubingh, D.N. Nonideal multicomponent mixed micelle model. J. Phys. Chem. 1983, 87, 1984–1990. [Google Scholar] [CrossRef]
  53. Clint, J.H. Micellization of mixed nonionic surface active agents. J. Chem. Soc. Faraday Trans. 1 Phys. Chem. Condens. Phases 1975, 71, 1327. [Google Scholar] [CrossRef]
  54. Motomura, K.; Yamanaka, M.; Aratono, M. Thermodynamic consideration of the mixed micelle of surfactants. Colloid Polym. Sci. 1984, 262, 948–955. [Google Scholar] [CrossRef]
  55. Negm, N.A.; Tawfik, S.M. Studies of Monolayer and Mixed Micelle Formation of Anionic and Nonionic Surfactants in the Presence of Adenosine-5-monophosphate. J. Solut. Chem. 2012, 41, 335–350. [Google Scholar] [CrossRef]
  56. Ren, Z.H.; Huang, J.; Zheng, Y.C.; Lai, L.; Yu, X.R.; Chang, Y.L.; Li, J.G.; Zhang, G.H. Mixed micellization of binary mixture of amino sulfonate amphoteric surfactant with octadecyltrimethyl ammonium bromide in water/isopropanol solution: Comparison with that in aqueous solution. J. Dispers. Sci. Technol. 2019, 40, 1353–1359. [Google Scholar] [CrossRef]
  57. Das, S.; Ghosh, S.; Das, B. Formation of Mixed Micelle in an Aqueous Mixture of a Surface Active Ionic Liquid and a Conventional Surfactant: Experiment and Modeling. J. Chem. Eng. Data 2018, 63, 3784–3800. [Google Scholar] [CrossRef]
  58. Rosen, M.J.; Cohen, A.W.; Dahanayake, M.; Hua, X.Y. Relationship of structure to properties in surfactants. 10. Surface and thermodynamic properties of 2-dodecyloxypoly(ethenoxyethanol)s, C12H25(OC2H4)xOH, in aqueous solution. J. Phys. Chem. 1982, 86, 541–545. [Google Scholar] [CrossRef]
  59. Bagheri, A.; Abolhasani, A. Binary mixtures of cationic surfactants with triton X-100 and the studies of physicochemical parameters of the mixed micelles. Korean J. Chem. Eng. 2015, 32, 308–315. [Google Scholar] [CrossRef]
  60. Ren, Z.H.; Luo, Y.; Zheng, Y.C.; Wang, C.J.; Shi, D.P.; Li, F.X. Micellization behavior of the mixtures of amino sulfonate amphoteric surfactant and octadecyltrimethyl ammonium bromide in aqueous solution at 40 °C: A tensiometric study. J. Mater. Sci. 2015, 50, 1965–1972. [Google Scholar] [CrossRef]
  61. Ren, Z.H. Interacting behavior between amino sulfonate amphoteric surfactant and octylphenol polyoxyethylene ether (7) in aqueous solution and pH effect. J. Ind. Eng. Chem. 2014, 20, 3649–3657. [Google Scholar] [CrossRef]
  62. Zhou, Q.; Rosen, M.J. Molecular Interactions of Surfactants in Mixed Monolayers at the Air/Aqueous Solution Interface and in Mixed Micelles in Aqueous Media: The Regular Solution Approach. Langmuir 2003, 19, 4555–4562. [Google Scholar] [CrossRef]
  63. Rosen, M.J.; Hua, X.Y. Surface concentrations and molecular interactions in binary mixtures of surfactants. J. Colloid Interface Sci. 1982, 86, 164–172. [Google Scholar] [CrossRef]
  64. Ananda, K.; Yadav, O.P.; Singh, P.P. Studies on the surface and thermodynamic properties of some surfactants in aqueous and water+1,4-dioxane solutions. Colloids Surf. 1991, 55, 345–358. [Google Scholar] [CrossRef]
  65. Oida, T.; Nakashima, N.; Nagadome, S.; Ko, J.-S.; Oh, S.-W.; Sugihara, G. Adsorption and Micelle Formation of Mixed Surfactant Systems in Water. III. A Comparison between Cationic Gemini/Cationic and Cationic Gemini/Nonionic Combinations. J. Oleo Sci. 2003, 52, 509–522. [Google Scholar] [CrossRef] [Green Version]
  66. Benesi, H.A.; Hildebrand, J.H. A Spectrophotometric Investigation of the Interaction of Iodine with Aromatic Hydrocarbons. J. Am. Chem. Soc. 1949, 71, 2703–2707. [Google Scholar] [CrossRef]
Figure 1. Chemical structures of (a) tetracaine hydrochloride (TCH), (b) sodium dodecyl sulfate (SDS), and (c) sodium lauryl sarcosine (SLS).
Figure 1. Chemical structures of (a) tetracaine hydrochloride (TCH), (b) sodium dodecyl sulfate (SDS), and (c) sodium lauryl sarcosine (SLS).
Gels 08 00234 g001
Figure 2. The physical appearance of TCH + SDS/SLS mixtures at different compositions: (a) SDS + TCH and (b) SLS + TCH.
Figure 2. The physical appearance of TCH + SDS/SLS mixtures at different compositions: (a) SDS + TCH and (b) SLS + TCH.
Gels 08 00234 g002
Figure 3. Graph of surface tension versus log molar concentration for SLS + TCH mixed systems.
Figure 3. Graph of surface tension versus log molar concentration for SLS + TCH mixed systems.
Gels 08 00234 g003
Figure 4. Experimentally determined critical micelle concentration (cmc) and ideal critical micelle concentration (cmc*) against mole fraction of surfactants (SDS/SLS) in mixed systems at 298.15 K.
Figure 4. Experimentally determined critical micelle concentration (cmc) and ideal critical micelle concentration (cmc*) against mole fraction of surfactants (SDS/SLS) in mixed systems at 298.15 K.
Gels 08 00234 g004
Figure 5. Absorption spectra of tetracaine hydrochloride in the presence of increased concentrations of (a) TCH + SDS and (b) TCH + SLS.
Figure 5. Absorption spectra of tetracaine hydrochloride in the presence of increased concentrations of (a) TCH + SDS and (b) TCH + SLS.
Gels 08 00234 g005
Figure 6. Benesi–Hildebrand plots for the interaction of TCH (a) SDS and (b) SLS.
Figure 6. Benesi–Hildebrand plots for the interaction of TCH (a) SDS and (b) SLS.
Gels 08 00234 g006
Table 1. Physical parameters of TCH + SDS/SLS mixed systems in aqueous NaCl.
Table 1. Physical parameters of TCH + SDS/SLS mixed systems in aqueous NaCl.
cmc
(mM)
cmc*
(mM)
X 1 R u b X 1 i d e a l βRub f 1 R u b f 2 R u b
SDS + TCH
0.079.43------
0.050.3716.420.540.8016.040.0330.0095
0.10.319.160.560.9015.200.0550.0081
0.70.131.450.650.9915.720.1390.0014
0.80.151.270.660.9915.440.1740.0011
0.90.121.130.670.9917.110.1560.0004
1.01.02------
SLS + TCH
0.079.43------
0.054.4941.240.500.518.860.1100.1078
0.13.0527.850.530.689.090.1390.0742
0.51.247.740.620.959.960.2430.0207
0.70.795.680.640.9811.810.2120.0082
0.90.334.490.640.9916.530.1150.0012
1.04.07------
Relative standard uncertainties (ur) are ur(cmc/cmc*) = 0.03, ur ( X 1 R u b / X 1 i d e a l ) = 0.02, ur (βRu) = 0.03, and ur ( f 1 R u b / f 2 R u b ) = 0.04.
Table 2. Energetic constraints of TCH + SDS/SLS mixtures in aqueous NaCl a.
Table 2. Energetic constraints of TCH + SDS/SLS mixtures in aqueous NaCl a.
α1 G m i x E / Δ H m ( kJmol 1 ) Δ G m i x ( kJmol 1 ) Δ G m i x i d e a l ( kJmol 1 ) T Δ S m ( kJmol 1 ) | T Δ S m Δ G m i x | Δ G m o ( kJmol 1 )
SDS + TCH
0.0-----16.23
0.059.8711.781.716.400.5429.48
0.19.2711.151.696.320.5729.93
0.78.9010.691.616.010.5632.08
0.88.5410.291.585.880.5731.79
0.99.3611.121.575.890.5332.34
1.0-----27.01
SLS + TCH
0.0-----16.23
0.055.497.331.726.170.8423.34
0.15.607.441.716.160.8224.30
0.55.797.561.645.930.7826.54
0.76.768.521.625.920.6927.66
0.99.4511.261.626.070.5429.77
1.0-----23.59
a Relative standard uncertainties (ur) are ur( G m i x E / Δ H m ) = 0.03, ur( Δ G m i x / Δ G m i x i d e a l ) = 0.03, ur( Δ S m ) = 0.03, and ur( Δ G m o ) = 0.03.
Table 3. Interfacial and packing data of TCH + SDS/SLS mixed system in aqueous NaCl a.
Table 3. Interfacial and packing data of TCH + SDS/SLS mixed system in aqueous NaCl a.
α1106 Γmax
(molm−2)
Amin
2)
Aideal
2)
C20γcmc
(mNm–1)
πcmc
(mNm–1)
SDS + TCH
0.01.641.01-19.3639.5731.43
0.051.770.941.010.0327.7943.21
0.12.440.681.010.0528.5542.45
0.73.100.530.990.0329.8841.12
0.83.390.490.980.0430.1740.83
0.93.280.510.970.0330.6840.32
1.01.710.97-0.0930.6040.40
SLS + TCH
0.01.641.01-19.3639.5731.43
0.052.730.611.010.8027.8843.11
0.12.280.731.010.4127.7243.28
0.52.570.651.030.1926.8944.11
0.72.140.771.040.0927.1143.89
0.92.010.831.050.0428.0442.96
1.01.571.05-0.1823.8047.20
a Relative standard uncertainties (ur) are urmax) = 0.05, ur(Amin/Aideal) = 0.03, ur(C20) = 0.03, and ur(γcmccmc) = 0.02.
Table 4. Thermodynamic and interfacial properties of TCH + SDS/SLS mixtures in aqueous NaCl a.
Table 4. Thermodynamic and interfacial properties of TCH + SDS/SLS mixtures in aqueous NaCl a.
α1 X 1 s β s f 1 s f 2 s G e x s ( kJmol 1 ) –ΔGads
(kJmol−1)
Gmin
(kJmol−1)
SDS + TCH
0.0-----35.3424.07
0.050.5617.660.0330.00410.7753.8915.69
0.10.5914.600.0910.0068.7147.3511.71
0.70.7112.280.3700.0026.1945.339.62
0.80.7411.400.4860.0025.3243.828.88
0.90.7512.830.4540.0015.9244.619.34
1.0-----50.7417.97
SLS + TCH
0.0-----34.8324.07
0.050.5419.050.0180.00411.7139.1610.22
0.10.5716.090.0510.0059.7743.2712.14
0.50.6414.520.1570.0028.2543.7110.47
0.70.6713.480.2510.0017.2648.1712.67
0.90.7015.100.2610.0017.8351.1513.94
1.0 ----53.5815.12
a Relative standard uncertainties (ur) are ur( X 1 S ) = 0.02, ur(βs) = 0.03, ur( f 1 s / f 2 s ) = 0.04, ur( G e x s ) = 0.03, urGads) = 0.03, and ur(Gmin) = 0.03.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Azum, N.; Rub, M.A.; Khan, A.; Alotaibi, M.M.; Asiri, A.M. Synergistic Interaction and Binding Efficiency of Tetracaine Hydrochloride (Anesthetic Drug) with Anionic Surfactants in the Presence of NaCl Solution Using Surface Tension and UV–Visible Spectroscopic Methods. Gels 2022, 8, 234. https://doi.org/10.3390/gels8040234

AMA Style

Azum N, Rub MA, Khan A, Alotaibi MM, Asiri AM. Synergistic Interaction and Binding Efficiency of Tetracaine Hydrochloride (Anesthetic Drug) with Anionic Surfactants in the Presence of NaCl Solution Using Surface Tension and UV–Visible Spectroscopic Methods. Gels. 2022; 8(4):234. https://doi.org/10.3390/gels8040234

Chicago/Turabian Style

Azum, Naved, Malik Abdul Rub, Anish Khan, Maha M. Alotaibi, and Abdullah M. Asiri. 2022. "Synergistic Interaction and Binding Efficiency of Tetracaine Hydrochloride (Anesthetic Drug) with Anionic Surfactants in the Presence of NaCl Solution Using Surface Tension and UV–Visible Spectroscopic Methods" Gels 8, no. 4: 234. https://doi.org/10.3390/gels8040234

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop