Interaction and Micellar Behavior of Binary Mixture of Amino Sulfonate

May 9, 2017 - College of Chemistry and Environmental Engineering, Applied Chemistry ... For a more comprehensive list of citations to this article, us...
0 downloads 0 Views 751KB Size
Article pubs.acs.org/jced

Interaction and Micellar Behavior of Binary Mixture of Amino Sulfonate Amphoteric Surfactant with Octadecyltrimethylammonium Bromide in Aqueous Solutions of NaCl Zhao Hua Ren,*,† Jing Huang,‡ Yan Cheng Zheng,† Lu Lai,† and Lin Li Hu† †

College of Chemistry and Environmental Engineering, Applied Chemistry Research Centre for Oil and Gas Fields, Yangtze University, Jingzhou 434023, China ‡ College of Management, Yangtze University, Jingzhou 434023, China ABSTRACT: The interaction behavior and the process of mixed micellization for binary mixtures of an amino sulfonate amphoteric surfactant sodium 3-(N-dodecyl ethylenediamino)-2-hydropropyl sulfonate (C12AS) and a cationic surfactant octadecyltrimethylammonium bromide (OTAB) in aqueous solution of NaCl (0.250 M) at 313.15 K were investigated using both the tensiometry and the UV−vis spectrometry using pyrene as a probe. The mixed critical micelle concentration (cmc) of the C12AS/OTAB mixtures were determined by two techniques. Based on regular solution theory, pseudophase separation model, Clint’s model, Rosen’s model, and Rubingh’s model, some parameters (including the ideal mixed cmc, the compositions in mixed micelle, the interaction parameters between two surfactants, the activity coefficients in mixed micelle, some thermodynamic parameters, etc.) were evaluated and calculated. The deviation of mixed cmc from its ideal value indicates a nonideal mixing between two surfactants. In comparison with those in aqueous solutions, the presence of NaCl induces some differences in the dependence of compositions in mixed micelle on its content in bulk solution, and there has a stronger synergism between two surfactants. Thermodynamic parameters shows that the mixed micellization of the C12AS/OTAB mixtures is an entropically spontaneous process, while the addition of NaCl to the solution can partly suppress the contribution of entropy to the process of micellization. The electrostatic attraction between two surfactants, the steric effect of hydrophilic group of C12AS, and the interaction repulsion between head groups of individual surfactants may play vital roles on the process of micellization. These findings help with probing the effect of additives on the process of micellization and understanding the interaction behavior between surfactants so as to design surfactant formulations.



INTRODUCTION In many practical applications, mixtures of surfactants, rather than individual surfactants, are widely used since they work better than their components.1−3 Herein, it should be worthy to note that some of surfactant mixtures are really not indicative of excellent properties. For surfactant mixtures, their properties are usually dependent on the interaction between surfactant molecules. These surfactant mixtures with excellent properties are often attributed to the synergistic interaction between their components, e.g., the binary mixtures of ionic/nonionic, ionic/ ionic surfactants,4−8 and the ternary mixtures of surfactants,9,10 etc. While those mixtures composed of nonionic surfactants have been reported to exhibit ideal behavior.1 Therefore, it should take attention to the interaction between surfactants and the micellar behavior of the mixture since they offer a better understanding of the causes for the properties of surfactant mixtures. Commonly, Rosen’s model,1 Rubingh’s model,11 and other thermodynamic models were adopted to estimate the relevant properties and the interaction parameters of surfactant mixtures. © XXXX American Chemical Society

Among the investigations about surfactant mixtures, binary mixtures of ionic/ionic or ionic/nonionic surfactants have been widely reported in literatures.4−8 Also, in recent years, binary mixtures of zwitterionic/ionic or zwitterionic/nonionic surfactants12−15 have also been focused, mainly resulting from the good compatibility of zwitterionic surfactant with other ionic or nonionic surfactants. Unlike conventional zwitterionic surfactant (e.g., alkyl betaine, etc.), an amino sulfonate amphoteric surfactant possesses one or more latent cationic centers and a small range of isoelectric points. 16 Regarding this, in comparison with conventional zwitterionic surfactants and their mixtures, there may be some different properties or behaviors of surfactant mixtures containing amino sulfonate amphoteric surfactant in solutions. In enhanced oil recovery (EOR), especially related to the chemical flooding, cationic surfactants were sometimes used to serve as sacrificial agents to Received: November 19, 2016 Accepted: May 2, 2017

A

DOI: 10.1021/acs.jced.6b00968 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Article

Shimadzu was used to scan the absorbance in the wavelength range of 220−400 nm, in the intermediate scanning speed and in the sample interval of 0.5 nm. In this investigation, the resulted value of surface tension or absorbance is a mean value of three experimental values measured by the tensiometry or UV−vis spectrometry. The uncertainties in the surface tension and the absorbance are 0.20 mN/m and 0.04, respectively.

reduce the loss of other surfactants and polymers through adsorption. Then, in this article the binary mixture of an amphoteric amino sulfonate surfactant sodium 3-(N-dodecyl ethylenediamino)-2-hydropropyl sulfonate (C12AS) developed by our group17,19,20 and a cationic octadecyltrimethylammonium bromide (OTAB) were investigated. Also, in EOR, it must be mentioned to the fact that there exist inorganic salts, e.g., sodium chloride (NaCl), magnesium chloride (MgCl2), etc., in the subterranean or injection water in many oil fields and the content of salt is commonly high enough, e.g., more than 10g/L even to 100g/L or more.16,17 On the basis of these considerations, this work is to investigate the micellization of C12AS/OTAB mixture in aqueous solution of NaCl (0.250 M). In this investigation, the experimental temperature was set to 313.15 K considering two aspects. First, it must be considered that an appropriate temperature can ensure the excellent solubility of surfactants (including C12AS, OTAB, and their mixtures) in aqueous solutions of NaCl, and the temperature is within the temperature range of the subterranean reservoir in many oil fields.17,18 Second, one of the goals in this work focuses on some deviations of the interaction behaviors of C12AS/OTAB mixture in aqueous solutions in the presence of salt from those in the absence of salt. Regarding this, the experimental temperature in this investigation is given as the value adopted in the previous literature.19 The main aim of this work is to survey the interaction behavior of binary surfactant mixture so as to attain some understanding or gain some relevant data to design surfactant formulation. With this purpose, the critical micelle concentration (cmc) values of binary surfactant mixture and their components were obtained by both the tensiometry and the UV−vis spectrometry using pyrene as a probe, and then these data were treated by some thermodynamic models. These obtained parameters were used to discuss the interaction behavior.



RESULTS AND DISCUSSION Micellization and Intermolecular Interaction in Mixed Micelle. The surface tension (γ) or the ratio (A/c) of absorbance (A) and surfactant concentration (c) vs the logarithm of c (log c) plots for the C12AS/OTAB mixtures are given in Figure 1. In Figure 1, with log c, there initially



EXPERIMENTAL SECTION Materials. C12AS developed by our group17,20 was treated to a purity of over 99 wt%, measured with Vario EL III Automatic Elementary Analyzer made by Germany Elementar Co. The chemical structure of C12AS is followed as n-C12H25− NH−CH 2 CH 2 −NH−CH 2 CH(OH)CH 2 −SO 3 Na. OTAB, NaCl, and anhydrous ethanol were purchased from Sinopharm Chemical Reagent Co., Ltd. and are analytical reagents with purities of over 99 wt%, all of which were used as supplied. Pyrene with a purity of 98 wt%, from American Aldrich Chemical Reagent Co., was recrystallized three times using the mixed solvent of deionized triple distilled (DTD) water and anhydrous ethanol to obtain a purity of over 99 wt%. Throughout all the experiments, the DTD water with a conductivity of about 4.01 μS/cm at 298.15 K measured by the DDSJ-318 conductometer made in China was used to prepare all the solutions of surfactants. Tensiometry and UV−vis Spectrometry. In this investigation, both the tensiometry and the UV−vis spectrometry using pyrene as a probe were adopted to determine the cmc value of individual or mixed surfactants. In the measurement of surface tension, a JK99B automatic tensiometer made in China was used to measure the experimental values of surface tension using Wilhelmy plate method21 at 313.15 ± 0.20 K. In the analysis of UV−vis spectrum, all the solutions added the stock solution of pyrene (its preparation refers to the literature22) were heated to a constant temperature of 313.15 ± 0.20 K, and then the UV-2450 UV−vis spectrophotometer made by

Figure 1. Variation of surface tension or the ratio (A/c) of absorbance (A) versus surfactant concentration (c) with the logarithm of c for the C12AS/OTAB mixtures with different compositions (x1) in aqueous solution of NaCl (0.250 M) at 313.15 K: (a) tensiometry; (b) UV−vis spectrometry (A at the wavelength of about 277 nm).

shows a linear decrease in γ or A/c up to cmc, beyond which γ or A/c remains more or less constant, showing a typical signature of surfactant.1,3,8,14,16,20−22,24−28 The intersection of two straight lines in each of the γ−log c or A/c−log c plot corresponds to the formation of micelle and is identified as the cmc value of mixed surfactants. By the way, the UV−vis spectrum shows a distinct and larger absorption peak at the wavelength of about 277 nm, and the peak is increasing gradually with the concentration of surfactant in solutions.22 Then, the absorbance at the wavelength of about 277 nm was adopted to determine the cmc value of surfactant. The cmc values for the C12AS/OTAB mixtures with different B

DOI: 10.1021/acs.jced.6b00968 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Article

Table 1. cmc Values and Micellization Parameters for the C12AS/OTAB Mixture in Aqueous Solution of NaCl (0.250 M) at 313 Ka cmc/(10−4 M) experimental x1

tensiometry

spectrometry

ideal

Xideal 1

X1

β12

f1

f2

0 0.150 0.300 0.450 0.600 0.750 0.900 1

3.150 2.679 2.482 2.373 2.448 2.597 2.720 2.724

3.287 2.760 2.544 2.464 2.509 2.545 2.612 2.700

3.129 3.044 2.964 2.888 2.815 2.746

0.174 0.338 0.494 0.641 0.781 0.915

0.244 0.382 0.495 0.607 0.737 0.889

−0.882

0.604 0.714 0.798 0.873 0.941 0.989

0.949 0.879 0.806 0.722 0.619 0.498

a Standard uncertainty u are u(T) = 0.20 K, u(P) = 0.006 MPa, u(x1) = 0.002, u(cmc) = 0.007 × 10−4 mol/L for the experimental value, u(cmc) = 0.004 × 10−4 mol/L for the ideal value, u(Xideal 1 ) = 0.004, u(X1) = 0.005, u(β12) = 0.023, u( f1) = 0.008, and u( f 2) = 0.008 (0.68 level of confidence). Note: P indicates the pressure.

repulsion between the cationic head groups of OTAB, and there exists an electrostatic attraction between C12AS and cationic OTAB, promoting the formation of mixed micelle and then decreasing the value of mixed cmc. For the amphoteric C12AS, it can be viewed as a slightly charged anionic surfactant in water/NaCl solution.16 Some similar investigations have been reported in literatures.25,26 However, when the content of C12AS in the C12AS/OTAB mixture increases, the steric effect of its headgroup16,19,20,22 may be gradually unfavorable to the formation of mixed micelle. Therein, even if an electrostatic attraction between two surfactants still works partly in the process of micellization, the steric contribution is increasing with adding C12AS and can be overwhelming when a content of C12AS is enough high. It can be found from Figure 2 that throughout the entire range of compositions investigated, the experimental value of mixed cmc has always a negative deviation from its ideal value, indicating a substantial interaction between two surfactants or a nonideal mixing. For binary surfactant mixtures, the mole fraction of component in mixed micelle can help with understanding the interaction behavior between two surfactants. Based on the pseudophase separation model, the mole fraction (Xideal 1 ) of component 1 in mixed micelle on an ideal state1,20,22,23 can be evaluated by the following relationship: x1cmc 2 X1ideal = x1cmc 2 + (1 − x1)cmc1 (2)

compositions and their individual components are listed in Table 1. Besides the experimental values of cmc, Table 1 also lists the ideal values of cmc estimated by the following equation:20,22−24 x 1 − x1 1 = 1 + ideal cmc1 cmc 2 cmc12 (1) where x1 is the mole fraction of component 1 of binary surfactant mixture in bulk solution, cmc1, cmc2, and cmcideal 12 are the cmc values of the components 1 and 2 in binary surfactant mixtures and the ideal value of mixed cmc, respectively. In this investigation, C12AS and OTAB are set to the components 1 and 2 in the binary mixture of C12AS/OTAB, respectively. As depicted in Figure 2, the experimental or ideal value of mixed cmc for the C12AS/OTAB mixture is dependent on the

While, according to the regular solution theory (RST), the mole fraction (X1) of component 1 in mixed micelle on a real state can be estimated by solving iteratively the following eq 3 obtained in the Rubingh’s treatment.11 X12 ln[x1cmc12/(X1cmc1)] Figure 2. Variation of the experimental or ideal value of mixed cmc with the composition (x1) in bulk solution for the C12AS/OTAB mixture in aqueous solution of NaCl (0.250 M) at 313.15 K.

(1 − X1)2 ln{(1 − x1)cmc12/[(1 − X1)cmc 2]}

=1 (3)

Where cmc12 is the mixed cmc of binary surfactant mixture. The values of X1 and Xideal for all the C12AS/OTAB mixtures 1 calculated respectively from eqs 2 and 3 are given in Table 1. Figure 3 shows that the mole fraction of C12AS in mixed micelle varies with that in bulk solution. It can be observed from Figure 3 that on an ideal state the mole fraction (Xideal 1 ) of C12AS in mixed micelle is always larger than that in bulk solution, implying a stronger micellization of C12AS relative to that of OTAB. Unlike the case of ideal state, the value of X1 in mixed micelle on a real state is initially more than that in bulk

composition (x1) of C12AS in bulk solution. Herein, the experimental values of mixed cmc shown in Figure 2 and adopted in the calculations of these following parameters are mean values of those measured by two techniques. It can be found in Figure 2 that at x1 < 0.45 the experimental value of mixed cmc decreases with x1, and while at x1 > 0.45 a reverse tendency arises. It is that in the solutions enriched with OTAB, the addition of C12AS can effectively shield the electrostatic C

DOI: 10.1021/acs.jced.6b00968 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Article

attributed to the stronger competition of steric effect in the presence of NaCl. For binary surfactant mixtures, an interaction parameter (β12) introduced in the Rubingh’s model11,16,19,20,27 can be adopted to interpret the nature and strength of interaction between two surfactants in mixed micelle and can be estimated as β12 =

ln[x1cmc12/(X1cmc1)] (1 − X1)2

(4)

Usually, the value of interaction parameter (β12) depends on the components in bulk solution.8,19,20,22,28 To obtain a unique value of β12, a nonlinear fitting has been performed as the transformation of eq 4: x1cmc12/(X1cmc1) = exp[β12(1 − X1)2 ] Figure 3. Variation of the mole fraction of C12AS (X1 or Xideal 1 ) in mixed micelle with its composition (x1) in bulk solution for the C12AS/OTAB mixture in aqueous solution of NaCl (0.250 M) at 313.15 K. Thin dash dot line via 0 and 1 is a diagonal.

(5)

Other dimensionless parameter activity coefficient (f i) was introduced to indicate a degree of participation of component i in mixed micelle.11,19,20,22 For binary surfactant mixtures, f1 and f 2 can be written as

solution, and then at x1 > 0.590 a reverse tendency happens, all of which are similar to the cases in aqueous solution.19 These phenomena can be explained by the interaction attraction between two surfactant, the electrostatic shielding effect and the steric effect. However, in comparison with that in aqueous solution, the presence of NaCl results in an increase in the value of X1 in solutions enriched with OTAB, especially in solution with an enough high concentration of OTAB. For example, at x1 = 0.100 the value of X1 (0.180) in water/NaCl solution is larger than 0.115 in aqueous solution obtained by the Rubingh’s treatment;19 also, at x1 = 0.200 the former (0.300) is still larger than the latter (0.267).19 This is that the addition of NaCl makes C12AS become a slightly charged anionic surfactant,16 consequently promoting an intercalation of C12AS into the micelle of OTAB by the electrostatic attraction. While, at x1 > 0.590 the presence of NaCl seems to not obviously influence the value of X1 relative to that in aqueous solution.19 For example, at x1 = 0.900 the value of X1 (0.889) in water/NaCl solution is almost equivalent to that (0.887) in aqueous solution.19 As mentioned above, the attraction between two surfactants in aqueous solution with NaCl is stronger than that in aqueous solution without NaCl. Even so, the steric effect of C12AS16,19−22 can be also increased by the addition of NaCl. Also, the steric effect is gradually increasing with increasing C12AS. Once the steric contribution is more than the electrostatic one, the intercalation of C12AS into the mixed micelle may be difficult to work. At x1 > 0.590, the case will happen. At x1 > 0.590, the increase in the electrostatic attraction may be insufficient to compensate the increase in the steric effect. However, the intercalation of small amount of OTAB into the micelle of C12AS can partly offset an increase in the steric effect by the compatibility in conformation between two surfactants. In addition, it can be also found from Figure 3 that the optimum mixing ratio, which can obtain the maximum synergism and in which x1 = X1,1,12,16,19,20,22 for the C12AS/ OTAB in water/NaCl solution, is about 0.590, which is smaller than 0.670 in aqueous solution without NaCl.19 Relative to that in aqueous solution without NaCl, the earlier appearance of the optimum mixing ratio in water/NaCl solution should be

f1 = exp β12(1 − X1)2

(6a)

f2 = exp β12X12

(6b)

The unique value of β12 and the activity coefficients are listed in Table 1. β12 near to zero implies that there is little interaction between two surfactants or ideal mixing while a positive value indicates antagonism, a negative value of β12 accounts for synergism in mixed micelle.1,3,5,8,11−14,20,22 For the C12AS/ OTAB mixtures in water/NaCl solution, the value of β12 is negative and fulfills the criteria of |β12| > |ln(cmc1/cmc2)|( = 0.171),1 suggesting the attractive interaction between two surfactants or synergistic effect. Also, the unique value of β12 in this investigation is always larger than each one in values of β12 in aqueous solution,19 which indicates that the presence of NaCl can promote the attractive interaction between two surfactants or synergistic effect. The activity coefficients listed in Table 1 are smaller than 1, implying a nonideal mixing. Also, the value of activity coefficient is dependent on the composition (x1) of C12AS in bulk solution. It means that in the process of micellization the degree of participation of components in the C12AS/OTAB mixture is subjected to the mixing ratio, further confirming their nonideal mixing. Thermodynamic Behavior of Micellization. Based on the RST, the free energy change (ΔmicG) of micellization for a binary surfactant mixture can be estimated from the following relationship8,11,20,22,28 ΔmicG = RT (X1ln f1 X1 + X 2 ln f2 X 2)

(7)

Accordingly, since f1 = f 2 = 1 on an ideal case, the free energy change (ΔmicGideal) of micellization can be reduced as ΔmicGideal = RT (X1ln X1 + X 2 ln X 2)

(8)

The enthalpy change (ΔmicH) of micellization can be written as11,20,22,27 ΔmicH = RT (X1ln f1 + X 2 ln f2 )

(9)

Then, the entropy change (ΔmicS) of micellization can be obtained using eqs 7 and 9 ΔmicS = (ΔmicH − ΔmicG)/T D

(10) DOI: 10.1021/acs.jced.6b00968 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Article

confirmed in many investigations.1,7,19,20,22,29,30 It is found from Figure 4 that the share of TΔmicS in ΔmicG varies with x1. It can be attributed to the variation of the contributions from the electrostatic attraction between molecules, the steric effect and the molecular interaction repulsion with x1. In comparison with that in aqueous solution,19 the presence of NaCl in this investigation may result in a stronger stability and ability of mixed micellization and a restraint in the entropically controlled process of micellization. These can be interpreted respectively from the smaller values of ΔmicG and | TΔmicS/ΔmicG| in water/NaCl solution to that in aqueous solution.19 The NaCl can be served as a water structure breaker1,21 and its addition induces the dehydration on the head groups of surfactants,1,16,20,21 which seems to result in an increase of ΔmicS. As mentioned above, the presence of NaCl can make C12AS become a slightly charged anionic surfactant.16 Consequently, in the process of micellization an organized structure of mixed micelle can be obtained by the attraction between an anionic C12AS and a cationic OTAB, then favoring the stability of mixed micellization and reducing ΔmicS.

Using the above eqs 7−10, thermodynamic parameters for the C12AS/OTAB mixture in water/NaCl solution at 313.15 K can be easily obtained and are listed in Table 2. Table 2. Thermodynamic Parameters for the C12AS/OTAB Mixture in Aqueous Solution of NaCl (0.250 M) at 313.15 Ka x1

ΔmicGideal (kJ/mol)

ΔmicG (kJ/mol)

ΔmicH (kJ/mol)

TΔmicS (kJ/mol)

|TΔmicS/ ΔmicG|

0.150 0.300 0.450 0.600 0.750 0.900

−1.448 −1.732 −1.804 −1.744 −1.501 −0.908

−1.872 −2.275 −2.379 −2.292 −1.947 −1.135

−0.424 −0.543 −0.574 −0.548 −0.446 −0.227

1.448 1.732 1.804 1.744 1.501 0.908

0.773 0.761 0.759 0.761 0.771 0.800

a

Standard uncertainty u are u(T) = 0.20K, u(P) = 0.006 MPa, u(x1) = 0.002, u(ΔmicGideal) = 0.027 kJ/mol, u(ΔmicG) = 0.095 kJ/mol, u(ΔmicH) = 0.017 kJ/mol, u(TΔmicS) = 0.028 kJ/mol, and u(TΔmicS/ ΔmicG) = 0.004 (0.68 level of confidence).

According to the thermodynamics,1,19,30 ΔmicG and ΔmicH are negative, but ΔmicS is positive, indicating the spontaneous process of micellization for the C12AS/OTAB mixture in water/NaCl solution. Figure 4 shows that the free energy



CONCLUSIONS



AUTHOR INFORMATION

In water/NaCl solution the micellization behavior of binary surfactant mixture of C12AS/OTAB was investigated using both the tensiometry and the UV−vis spectrometry. The regular solution theory, the pseudophase separation model, and other thermodynamic models (e.g., Clint’s model, Rubingh’s model, etc.) were adopted to estimate the compositions in mixed micelle, the interaction parameter between surfactants, the activity coefficients in mixed micelle, and other parameters of micellization. The mixed critical micelle concentration (cmc) of the C12AS/OTAB mixture shows some negative deviation from the ideal case, implying nonideal mixing. In water/NaCl solution, the dependence of the compositions in mixed micelle on its contents in bulk solution shows some differences from that in aqueous solution. In solutions enriched with OTAB, the mole fraction (X1) of C12AS in mixed micelle in water/NaCl solution is greater than that in aqueous solution while there are not some obvious differences in X1 at other compositions in bulk solution. The addition of NaCl to the solution results in a stronger synergism between two surfactants to that in aqueous solution without NaCl. Also, the presence of NaCl also induces an earlier appearance of optimum mixing. These phenomena can be explained by the electrostatic attraction between two surfactants, the steric effect of headgroup of ampheteric C12AS, and the interaction repulsion between the head groups of individual surfactants. Thermodynamic parameters indicate that the spontaneous process of micellization is entropically favorable and shows more stable than that in aqueous solution without NaCl. While, the contribution of entropy to micellization may be suppressed partly by the addition of NaCl.

Figure 4. Variation of the free energy change of micellization or the share of entropy change in the process of micellization with the mole fraction of C12AS (x1) in bulk solution for the C12AS/OTAB mixture in aqueous solution of NaCl (0.250 M) at 313.15 K.

change (ΔmicG) of micellization on a real state has a negative deviation from that (ΔmicGideal) on an ideal state, favoring the formation of mixed micelle. Also, the largest deviation of ΔmicG from ΔmicGideal appears at about 0.550 (x1), in which the mixed micelle may be most stable and there is a maximum synergism between two surfactants as described above. Therein, it must be mentioned that 0.550 should correspond to 0.590 mentioned above, but there was a large deviation between two values which is attributed to the errors in the calculation of parameters and in the fit of curve in the plots. The contributions from the electrostatic attraction between C12AS and OTAB, the steric effect of headgroup of C12AS, and the interaction repulsion between ionic head groups of individual surfactants can be used to explain these facts for the largest deviation at about 0.550 (x1) and the reducing deviation at other values of x1. For all the C12AS/OTAB mixtures in water/NaCl solution, the micellization is an entropically driving process, which can be stated that the share of TΔmicS in ΔmicG is more than 0.750. The entropically controlled process of micellization has also been

Corresponding Author

*E-mail: [email protected]; Tel(Fax): +86-716-8060650. ORCID

Zhao Hua Ren: 0000-0002-9796-8039 Lu Lai: 0000-0002-1254-7711 E

DOI: 10.1021/acs.jced.6b00968 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Article

Funding

(18) Nilsen, R. K.; Beeder, J.; Thorstenson, T.; Torsvik, T. Distribution of thermophilic marine sulfate reducers in north sea oil field waters and oil reservoirs. Appl. Environ. Microb. 1996, 62, 1793− 1798. (19) 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. (20) Ren, Z. H.; Huang, J.; Luo, Y.; Zheng, Y. C.; Mei, P.; Lai, L.; Chang, Y. L. Micellization behavior of binary mixtures of amino sulfonate amphoteric surfactant with different octylphenol polyoxyethylene ethers in aqueous salt solution: Both cationic and hydrophilic effects. J. Ind. Eng. Chem. 2016, 36, 263−270. (21) Ren, Z. H.; Chen, D. J.; Luo, Y.; Huang, J. Investigation of influence of inorganic salt on the critical micelle concentration of sodium octylphenol polyoxyethylenated ethylsulfonate. Acta Chim. Sinica 2010, 68, 1771−1775. (22) Ren, Z. H.; Luo, Y.; Zheng, Y. C.; Shi, D. P.; Mei, P.; Li, F. S. Interacting behavior between amino sulfonate surfactant and octylphenol polyoxyethylene ether (10) in aqueous solution. J. Solution Chem. 2014, 43, 853−869. (23) Clint, J. H. Micellisation of mixed nonionic surface active agents. J. Chem. Soc., Faraday Trans. 1 1975, 71, 1327−1334. (24) Aoudia, M.; Al-Maamari, T.; Al-Salmi, F. Intramolecular and intermolecular ion−dipole interactions in sodium lauryl ether sulfates (SLES) self-aggregation and mixed micellization with Triton X-100. Colloids Surf., A 2009, 335, 55−61. (25) Kroflic, A.; Sarac, B.; Bester-Rogac, M. Thermodynamic characterization of 3-[(3-Cholamidopropyl)-dimethylammonium]-1propanesulfonate (CHAPS) micellization using isothermal titration calorimetry: Temperature, salt, and pH dependence. Langmuir 2012, 28, 10363−10371. (26) Qin, X.; Liu, M.; Zhang, X.; Yang, D. Proton NMR based investigation of the effects of temperature and NaCl on micellar properties of CHAPS. J. Phys. Chem. B 2011, 115, 1991−1998. (27) Cui, Z. G.; Canselier, J. P. Interfacial and micellar properties of some anionic/cationic binary surfactant systems. 1. Surface properties and prediction of surface tension. Colloid Polym. Sci. 2000, 278, 22−29. (28) Rub, M. A.; Kumar, D.; Azum, N.; Khan, F.; Asiri, A. M. Study of the interaction between promazine hydrochloride and surfactant (conventional gemini) mixtures at different temperatures. J. Solution Chem. 2014, 43, 930−949. (29) Shimizu, S.; Pires, P. A. R.; Loh, W.; Seoud, O. A. E. Thermodynamics of micellization of cationic surfactants in aqueous solutions: consequences of the presence of the 2-acylaminoethyl moiety in the surfactant head group. Colloid Polym. Sci. 2004, 282, 1026−1032. (30) Sikorska, E.; Wyrzykowski, D.; Szutkowski, K.; Greber, K.; Lubecka, E. A.; Zhukov, I. Thermodynamics, size, and dynamics of zwitterionic dodecylphosphocholine and anionic sodium dodecyl sulfate mixed micelles. J. Therm. Anal. Calorim. 2016, 123, 511−523.

Funding for this work was provided by the National Natural Science Foundation of China (51304029), the Natural Science Foundation of Hubei Province (2016CFB477), China, and the Training Program for Youth Scientific Research Team of College of Chemistry & Environmental Engineering, Yangtze University, China. Notes

The authors declare no competing financial interest.



REFERENCES

(1) Rosen, M. J.; Junkappu, J. T. Surfactants and interfacial phenomena, 4th ed.; Wiley-Interscience: Hoboken, NL, 2012. (2) Tzocheva, S. S.; Kralchevsky, P. A.; Danov, K. D.; Georgieva, G. S.; Post, A. J.; Ananthapadmanabhan, K. P. Solubility limits and phase diagrams for fatty acids in anionic (SLES) and zwitterionic (CPAB) micellar surfactant solutions. J. Colloid Interface Sci. 2012, 369, 274− 286. (3) Goloub, T. P.; Pugh, R. J.; Zhmud, B. V. Micellar interactions in nonionic ionic mixed surfactant systems. J. Colloid Interface Sci. 2000, 229, 72−81. (4) Bucci, S.; Fagotti, C.; Degiorgio, V.; Piazza, R. Small-angle neutron-scattering study of ionic-nonionic mixed micelles. Langmuir 1991, 7, 824−826. (5) Shiloach, A.; Blankschtein, D. Measurement and prediction of ionic/nonionic mixed micelle formation and growth. Langmuir 1998, 14, 7166−7182. (6) Shrestha, R. G.; Shrestha, L. K.; Aramaki, K. Formation of wormlike micelle in a mixed amino-acid based anionic surfactant and cationic surfactant systems. J. Colloid Interface Sci. 2007, 311, 276−284. (7) Chauhan, S.; Sharma, K. Effect of temperature and additives on the critical micelle concentration and thermodynamics of micelle formation of sodium dodecyl benzene sulfonate and dodecyltrimethylammonium bromide in aqueous solution: A conductometric study. J. Chem. Thermodyn. 2014, 71, 205−211. (8) Parekh, P.; Varade, D.; Parikh, J.; Bahadur, P. Anionic-cationic mixed surfactant systems: Micellar interaction of sodium dodecyl trioxyethylene sulfate with cationic gemini surfactants. Colloids Surf., A 2011, 385, 111−120. (9) Murphy, A.; Taggart, G. A comparison of predicted and experimental critical micelle concentration values of cationic and anionic ternary surfactant mixtures using molecular-thermodynamic theory and pseudophase separation theory. Colloids Surf., A 2002, 205, 237−248. (10) Shiloach, A.; Blankschtein, D. Prediction of critical micelle concentrations of nonideal ternary surfactant mixtures. Langmuir 1998, 14, 4105−4114. (11) Holland, P. M.; Rubingh, D. N. Nonideal multicomponent mixed micelle model. J. Phys. Chem. 1983, 87, 1984−1990. (12) Lopez-Diaz, D.; Garcia-Mateos, I.; Velazquez, M. M. Synergism in mixtures of zwitterionic and ionic surfactants. Colloids Surf., A 2005, 270−271, 153−162. (13) Mclachlan, A. A.; Marangoni, D. G. Interactions between zwitterionic and conventional anionic and cationic surfactants. J. Colloid Interface Sci. 2006, 295, 243−248. (14) Singh, K.; Marangoni, D. G. Synergistic interactions in the mixed micelles of cationic gemini with zwitterionic surfactants: The pH and spacer effect. J. Colloid Interface Sci. 2007, 315, 620−626. (15) Wydro, P. The influence of the size of the hydrophilic group on the miscibility of zwitterionic and nonionic surfactants in mixed monolayers and micelles. J. Colloid Interface Sci. 2007, 316, 107−113. (16) Ren, Z. H. Mechanism of the salt effect on micellization of an aminosulfonate amphoteric surfactant. Ind. Eng. Chem. Res. 2015, 54, 9683−9688. (17) Ren, Z. H.; Luo, Y. Dynamic interfacial tension behavior of alkyl amino sulfonate in crude oil-brine system. Tenside, Surfactants, Deterg. 2013, 50, 369−375. F

DOI: 10.1021/acs.jced.6b00968 J. Chem. Eng. Data XXXX, XXX, XXX−XXX