Phase Separation and Thermodynamic Behavior ... - ACS Publications

Mar 28, 2019 - Department of Chemistry, Jahangirnagar University, Savar, Dhaka 1342, ... Bangabandhu Sheikh Mujibur Rahman Agricultural University, ...
0 downloads 0 Views 667KB Size
Article pubs.acs.org/jced

Cite This: J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Phase Separation and Thermodynamic Behavior of Triton X‑100 in the Occurrence of Levofloxacin Hemihydrates: Influence of Additives

J. Chem. Eng. Data Downloaded from pubs.acs.org by UNIV OF LOUISIANA AT LAFAYETTE on 04/15/19. For personal use only.

Md. Ruhul Amin,† Shamim Mahbub,† Mohammad Robel Molla,† Md. Masud Alam,‡ Md. Farhad Hossain,§ Shahed Rana,† Malik Abdul Rub,∥ Md. Anamul Hoque,† and Dileep Kumar*,⊥,# †

Department of Chemistry, Jahangirnagar University, Savar, Dhaka 1342, Bangladesh Department of Chemistry, Mawlana Bhashani Science and Technology University, Santosh, Tangail 1902, Bangladesh § Department of Plant Pathology, Bangabandhu Sheikh Mujibur Rahman Agricultural University, Gazipur 1706, Bangladesh ∥ Chemistry Department, Faculty of Science, King Abdulaziz University, Jeddah 21589, Saudi Arabia ⊥ Division of Computational Physics, Institute for Computational Science, Ton Duc Thang University, Ho Chi Minh City 700000, Vietnam # Faculty of Applied Sciences, Ton Duc Thang University, Ho Chi Minh City 700000, Vietnam ‡

S Supporting Information *

ABSTRACT: Measurements of clouding phenomena in Triton X-100 (TX-100) were carried out in the aqueous system and the presence of the antibiotic drug levofloxacin hemihydrate (LFH)/(LFH + inorganic salts/alcohols) and were explained thoroughly. In the aqueous system, the observed values of cloud point (CP) of TX-100 increased continuously with the increasing concentration of TX-100. In the occurrence of LFH, the magnitudes of cloud point of a certain concentration of TX-100 were found to increase first, and after that, the CP value declined with increasing concentration of the drug. The magnitudes of CP for the LFH + TX-100 mixed system were found to be dwindled through the enrichment of salt concentration in the presence/absence of a drug (LFH). The CP values in the presence of salts follow the order K2SO4 < KNO3 < K3PO4. On the other hand, in the case of alcohol solvent, the CP values decreased from lower alcohol to higher alcohol and increased through the enrichment of alcohol concentration in the presence and absence of a drug (LFH). Various types of thermodynamics parameters, such as standard free energy (ΔG0c), standard enthalpy (ΔH0c), and standard entropy (ΔS0c) changes of phase separation, were also determined and explained in an extent on the basis of their nature. The values of free energy changes (ΔG0c) were positive, which indicated that the clouding processes were nonspontaneous in nature.

1. INTRODUCTION Surfactants are amphiphilic substances having characteristic properties of decreasing surface tension of the water, and most of the solvents are called surfactants.1 Surfactants have been utilized widely in the pharmaceutical industry, cosmetic items, domestic detergents, and different industrial formulations/ processing such as solubilizers, wetting agent, emulsifiers, and separation of metal ions, organic molecules, enzymes, etc. since several decades.2−4 Surfactant micelles have been utilized as a simplified model of biomembranes.5 In pharmaceutical formulations, surfactants are foresighted to solubilize the poorly soluble essential organic compounds in water by accommodating them in the surfactant micelle.6−9 Triton X100 (TX-100) is one of the commonly utilized detergents in laboratories of research purposes and industries. It can also be utilized to lyse cells to extract protein/organelles.10 It is one of the common constituents in influenza vaccine. TX-100 is often used in biochemical applications, for example, to solubilize © XXXX American Chemical Society

proteins, for the recovery of membrane components under mild nondenaturing conditions.10 The solubility of the nonionic surfactant is known to be increased due to the presence of hydrogen bonding. The weak hydrogen bonds are expected to break upon heating, and thus, the solubility of the nonionic surfactant is certainly decreased in an aqueous medium and turns cloudy at a specific temperature, which is referred to as cloud point (CP).1,11 Surfactant solution exists as a single phase below and above CP, and the solubility is decreased at CP, which results into the cloudiness.12 The CP values can be applied in liquid−liquid extraction (LLE) to separate different compounds, for example, metal complexes on the basis of their relative solubility in two immiscible liquids. Recently, the nature of the Received: February 12, 2019 Accepted: March 28, 2019

A

DOI: 10.1021/acs.jced.9b00146 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Article

All solutions were prepared using distilled−deionized water having a specific conductivity of 1−2 × 10−6 S·cm−1 over the temperature range of 298.15−318.15 K. TX-100 is easily solubilized in distilled−deionized water (more than five times of employed concentration in the present study) in the temperature range studied. All phase separation/cloud point values were obtained by positioning Pyrex glass tubes (containing a surfactant solution (TX-100/ TX-100 + LFH/ TX-100 + LFH + inorganic salts/alcohols)) into a thermostatted H2O bath, the temperature of the systems was raised slowly (0.2 K/min when the temperature reached the preestimated CP values), and phase segregation was estimated through visual observation. The temperature above which clouding just began was termed CP, and the detailed procedure was reported earlier.24,25 A similar experiment was reproduced at a minimum of three times. The average temperature of vision and disappearance of phase separation was believed to be the cloud point of the respective system. The reproducibility in CP values was achieved not beyond ±0.1 K.

clouding phenomenon has been illustrated in terms of two considerations.13,14 First, the nonionic micelles formed with continuous heating up to CP,9 and the second point is that CP reveals as the critical temperature that is attained when micelles get closer jointly, and at a temperature higher than CP, the micelles tend to separate out as the insoluble phase (second phase). Alauddin et al.15 performed the experiment to estimate the CP values of TX-100 in the occurrence of an aliphatic compound and alcohols. Mudawadkar et al.16 investigated the phase segregation phenomenon and thermodynamic behavior of pure TX-100, as well as the presence of various polyvinyl pyrrolidones (PVPs), and accounted that CP values enhance slightly from 336 to 340 K with the augmentation of the concentration of TX-100 from 1 to 10 wt %. The variation of CP of TX-100 has been envisaged very slowly with the variation of surfactant concentration.17 Panchal et al.18 studied the CP of individual TX-100 along with the TX-100 + sodium lauryl sulfate (anionic surfactant) mixture in the presence and absence of electrolytes. They reported that CP of the employed surfactant enhanced in the occurrence of SLS.18 Another researcher19 also accounted the clouding phenomena in ionic liquids and the surfactant mixed system as well as their thermodynamics parameters. In addition, Mahajan et al.20 studied the consequence of different organic compounds with the hydroxyl functional group (−OH)/amines/salts on the clouding behavior of above-mentioned nonionic surfactants. The hydrophilic−lipophilic balance (HLB), as well as the strength of the hydrogen bonding, can be altered in the occurrence of different additives, which is the cause of the change in the neighboring atmosphere and hence the phase segregation in the occurrence of additives.9,21 The sagacity about the CP of nonionic amphiphiles is important because industrial formulations with temperatures noticeably higher than the CP temperature perhaps will perform phase separation of the surfactant solution. The performance of nonionic surfactants has been reported to be much efficient close to and beneath CP for a specific surfactant.22 Thus, the investigation of the consequence of different additives on the phase segregation of the nonionic surfactant is very important. The obvious significance of the CP can be obtained in the fact that suspensions, emulsions, foams, and ointments stagnated in the presence of nonionic surfactants when not fixed but heated close to CP.18,23 Considering all of the facts mentioned above, herein, we devised our study to obtain the CP values and different thermodynamics parameters associated with the clouding behavior, for example, free energy (ΔG0c), enthalpy (ΔH0c), and entropy (ΔS0c) changes of TX-100 (Scheme 1) in the absence/presence of (LFH)/(LFH + various electrolytes/ alcohols).

3. RESULTS AND DISCUSSION 3.1. CP of TX-100: Effect of Additives. The critical micelle concentration (cmc), micellar size, micelle aggregation number, and phase separation are remarkable results for each amphiphilic system.1,26 The critical micelle concentration value of TX-100 at room temperature was 0.2 to 0.31 mM, as measured by surface tension measurement technique.27 In the current system, the clouding behavior of TX-100 was investigated in different conditions. The concentration of the currently employed surfactant (TX-100) for cloud point evaluations was based on the cmc of the surfactant, not on the basis of the CP of pure TX-100. The concentration was chosen in such a way that the entire surfactant molecule exists in the aggregate form, that is, in micellized form, meaning that there is no other particular interest in choosing the concentration of TX-100 except above their cmc value.28 Different additives of dissimilar character and possessions were introduced at varying concentrations to carry out the change in phase separation of TX-100 with the variation in temperature. In this current investigation, the additives utilized were the drug (LFH) and different electrolytes (KNO3, K2SO4 and K3PO4·7H2O) as well as different alcohols (C2H5OH, C3H7OH, and C4H9OH). The process of phase segregation is an energetically regulable process; thus, an acquaintance of the outcome of different additives (possibly employed as a pharmaceutical constituent) on phase separation of TX-100 is required from the application perspective. The ethylene oxide (EO) part of the surfactant inclines on the adjuvant in both hydrophilic and hydrophobic interactions, which is really oversensitive to the existence of the additives.29 With the enhancement of the temperature, the phase separation of TX-100 happens due to the thermal movement of H2O particles that affect the interaction/salvation of micelles. Accordingly, by increasing the temperatures, surfactant aggregates (micelles) initiate to interact with each other, developing a network structure.30 The CP of TX-100 solutions in an aqueous medium was evaluated at different concentrations (9.51, 19.95, 30.76, 49.85, 69.67, 88.51, and 110.59 mmol·kg−1) of TX-100 along with their magnitudes, as profiled in Table 2. Herein, the obtained CP value of the 9.51 mmol·kg−1 surfactant in an aqueous medium was 336.99 K, exhibiting good consistency by the value obtained earlier, and

2. EXPERIMENTAL SECTION All chemicals applied in this study are of analytical grade and used without any additional purification. The provenance and purity of chemicals are shown in Table 1. Scheme 1. Molecular Structure of Nonionic Surfactant (TX100)

B

DOI: 10.1021/acs.jced.9b00146 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Article

Table 1. Origin and Purity of the Chemical Utilized chemical

source

mass fraction purity

CAS number

TX-100 LFH KNO3 K2SO4 K3PO4·7 H2O C2H5OH C3H7OH C4H9OH H2O

Merck, Germany General Pharmaceutical Ltd., Bangladesh BDH, India Merck, India Merck, Germany Merck, Germany Merck, Germany Merck, Germany double-distilled deionized

0.98 0.98 0.99 0.99 0.99 0.98 0.98 0.99

9002-93-1 138199-71-0 7757-79-1 7778-80-5 7778-53-2 64-17-5 71-23-8 71-36-3 1−2

19.95 and 30.76 mmol·kg−1 TX-100 in the occurrence of different concentrations of LFH are shown in Figure 1 and

Table 2. CP Temperature (TCP) Values for Individual TX100 in the Aqueous System at Pressure p = 0.1 MPaa cTX‑100 (103) (mol·kg−1)

TCP (K)

9.51 19.95 30.76 49.85 69.67 88.51 110.59

336.99 337.15 337.40 337.65 337.85 338.25 338.90

specific conductance (×10−6 S·cm−1)

a

Standard uncertainties (u) are u(T) = 0.1 K, u(p) = 5 kPa, and u(C) = 0.02 mmol·kg−1 (level of confidence = 0.68).

their value was 336.85 K for 1.15 × 10−2 mol·kg−1 surfactant concentration in an aqueous medium.31 In this study, we viewed that the CP value of TX-100 increases with the increment of their concentration in an aqueous solution, which is also in good agreement with results by Mudawadkar et al.16 who noticed that CP values of TX-100 increase gradually from 336 to 340 K with the augmentation of the concentration (1 to 10%, w/w). The CP of this TX-100 has been estimated and achieved to vary gradually by means of the augmentation of concentration.19 There are two principal causes to study and manage the phase segregation phenomenon: (i) remarkable progress has been made in the machinery of this phenomenon to an amount of extraction, and (ii) also in the separation application, the application of this phenomenon for a particular purpose can reluctantly have a consequence on the action of the surfactant-adopted formulation that we ought to evade. The clouding phenomenon of TX-100 in the aqueous medium occurs due to large enhancement of the aggregation number (Nagg) of the surfactant and the decrease in intermicellar repulsion, which is imminent from lessened dehydration of oxyethylene oxygen parts of surfactant with the augmentation of the temperature. The enhanced Nagg and dwindled intermicellar repulsion generate bulky micellar association at cloud point temperature, and thus, clouding appears; that is, phase segregation takes place.32 The CP values of the currently utilized surfactant increase with increasing concentration of the surfactant, and the size of the spherical micelle increases; in addition, the shape also changes from a sphere shape to a disc-like shape at an elevated concentration of the nonionic surfactant.30 This observable fact enhances the surface area of the micelles and consequently reduces dehydration, thus resulting in the increase in the CP of the system. Herein, the outcome of LFH addition to TX-100 solution was estimated from the CP measurement. The CP values of

Figure 1. Plot of TCP against [LFH] for (red boxes) 19.95 × 10−3 mol·kg−1 and (dark cyan-blue circles) 30.76 × 10−3 mol·kg−1 TX-100 systems.

Table S1 (Supporting Information). The cloud point variation of TX-100 solution by the LFH drug is a remarkable physicochemically modified phenomenon. The cloud point magnificence of TX-100 was found to be increased first by means of the augmentation of the concentration of LFH and then decreased with the further enhancement of the concentrations of LFH. At a lower concentration of the LFH drug, the drug attaches by means of polyethylene oxide of TX100 via discharge of solvated H2O from the allied centers, raising the surface charge of micelles; therefore, the CP value of systems increases. The reduced CP value of TX-100 in the occurrence of LFH because of the drug acting as the function of salt and the salt-out result subsists, resulting in a cloud point decrease. This process makes it possible for the surfactant micelles to get together and thus decreases the CP of TX-100, which is further reduced with the increase of LFH concentration. The CP values of the surfactant do not alter noticeably if the applied additives have no effect on the characteristics of surfactant micelles, such as the growth of micellar size or varying the extent of hydration. The impact of various electrolytes, for example, KNO3, K2SO4, and K3PO4·7H2O, on the CP values of LFH + TX-100 mixtures having fixed concentrations of the drug (1.0 mmol·kg−1) and 19.95 mmol· kg−1 solutions of TX-100 was observed, and results are shown in Figure 2 and Table S2 (Supporting Information). The CP values of LFH + TX-100 mixtures decreased upon adding different salts in comparison to the aqueous medium due to the salting-out effect. In fact, the electrolytes exist in the C

DOI: 10.1021/acs.jced.9b00146 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Article

Figure 2. Graph of CP against concentrations of potassium salts (a) KNO3, (b) K2SO4, and (c) K3PO4·7H2O of the 19.95 × 10−3 mol·kg−1 TX-100 system.

structure makers and thus immobilize water molecules, for example, SO42−, are positioned at the left region of the Hofmeister series. The insertion of the electrolytes in the surfactant solution is the cause of the salting-out effect. During this phenomenon, the electrolyte ions pull the H2O molecules toward themselves and make the surfactant molecules insoluble, which is the cause of phase segregation. Thus, with the introduction of the electrolyte to the surfactant solution, CP values were found to be reduced. Again, the ions, which are H2O organization breakers and have more polarizability, lead to the interruption of structured water and are positioned at the right side of the Hofmeister series. These ions deliver more H2O particles that are efficient to form H-bonding by means of the polar group of the constituent, and thus, the salting-in effect takes place. In the alternative mechanism, it has latterly noticed that the ions interact directly with the surfactant in comparison to the innovation of water structure with the introduction of electrolytes to the surfactant solutions. The insertion of electrolytes to the surfactant solution leads the dehydration of TX-100, thus reducing the interactions between the surfactant and H2O particles. Because the H2O particles are bound to the electrolyte ions, solute−solute interactions are increased, and thus, phase segregation happens. In the case of the introduction of electrolytes, divalent anions are more effective in salting out the nonpolar chain because of their higher charge density as compared to that of the monovalent anions. The salting-out effect in the occurrence of electrolytes is the cause of reduced solubility of

micelles and the aqueous phase. Additionally, the entire calculated concentrations of the electrolytes do not cross their water solubility; thus, no solid phase formation happens. For this reason, the insertion of particular inorganic electrolytes possibly reduces the CP value as electrostatic interactions between electrolytes and H2O molecules are found to be higher in magnitude in comparison to the H-bonds existing between the polar portion of amphiphiles and H2O molecules. Herein, the CP values of the surfactant were found to be reduced in the presence of inorganic electrolytes utilized. The efficacy of the electrolytes in decreasing CP values follows the order K2SO4 > KNO3 > K3PO4·7H2O. The insertion of inorganic electrolytes to the surfactant system is necessary to increase the low-temperature surface activity of the surfactants. Again, the mixture of the electrolytes and surfactants is important when surfactants need to be utilized at a temperature beyond its CP value.26 In view of the fact that various electrolytes exist in the biological fluid, hence, their occurrence may influence the characteristics of the surfactant in the presence of drug and drug + electrolytes media. The cloud point values of employed surfactant solution lessened more sharply with K2SO4 in comparison to those in the presence of KNO3. This achieved outcome is possibly due to the nature of charge of employed electrolyte, which concludes the overall of the effect on the CP of TX-100. Two different mechanisms have been proposed to demonstrate the position of the cations and anions in the Hofmeister series.27 Normally, the mechanism is assumed to associate with the efficacy of the ions to alter the H-bond formation capability of water. The ions, which are water D

DOI: 10.1021/acs.jced.9b00146 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Article

TX-100, and earlier phase segregation happens in the presence of electrolytes. The impact of different alcohols, for example, C2H5OH, C3H7OH, and C4H9OH, on the CP of the 19.95 × 10−3 mol· kg−1 nonionic surfactant in the occurrence of 1.02 × 10−3 mol· kg−1 LFH was furthermore estimated, and their achieved outcomes are profiled in Table S3 (Supporting Information). Usually in such systems, the CP in the presence of alcohols depends on the chain length of the alcohol. In the case of C2H5OH and C3H7OH with increasing concentration of alcohol, the CP values increased (both in the absence and presence of LFH), but more increase in CP of TX-100 takes place in a solution of TX-100 + drug mixtures. However, in the presence of C4H9OH, the CP values of studied systems were so low that they could not be determined by the current technique; therefore, they barred the study. The increase in CP in the presence of C2H5OH and C3H7OH in the TX-100 solution both in the absence and presence of a drug occurs because alcohol molecules are placed at the exterior/surface of micelles formed by TX-100, which obstruct the micellar association, resulting in an increase in the CP of studied systems. 3.2. Thermodynamics of Clouding Phenomenon. The CP values of any surfactant are used to determine their limit of its solubility as surfactant solutions segregate into two different liquid phases at any temperature superior than CP. The various thermodynamics quantities of the clouding phenomenon of TX-100 were obtained from the CP measurement. The standard free energy (ΔG0c), standard enthalpy (ΔH0c), and standard entropy (ΔS0c) changes of clouding were evaluated by envisaging the limit of the solubility of the surfactant at CP temperature based on the concept of phase segregation model using the subsequent set of the equations2,27,31,33 ΔG 0 c = −RT ln Xs

(1)

ΔH 0 c = RT 2 (∂ ln Xs)/∂T

(2)

ΔS 0 c = (ΔH 0 m − ΔG 0 m)/T

(3)

Figure 3. Graph of ln Xs against TCP of 19.95 mmol·kg−1 TX-100 solution in the occurrence of different concentrations of the LFH drug.

The obtained diverse thermodynamics parameter values of the current studied system are summarized in Table 3. The Table 3. Various Thermodynamics Parameters of Phase Separation in Pure TX-100 in Aqueous Solution at Pressure p = 0.1 MPaa ΔH0c (10−4) (kJ·mol−1)

ΔS0c (10−6) (J·mol−1·K−1)

9.51 19.95 30.76 49.85 69.67 88.51 110.59

24.30 22.24 21.04 19.70 18.78 18.13 17.54

24.19 24.23 24.28 24.33 24.38 24.46 24.60

39.02 39.08 39.16 39.25 39.32 39.46 39.68

Standard uncertainties (u) are u(T) = 0.1 K, u(c) = 0.02 mmol·kg−1, and u(p) = 5 kPa (level of confidence = 0.68). Relative standard uncertainties (ur) are ur(ΔG0c) = ±3%, u(ΔH0c) = ±3%, and ur(ΔS0c) = ±4%.

obtained magnitudes of ΔG0c were found to be positive, implying the nonspontaneous clouding phenomenon. The positive magnitudes of ΔG0c were found to dwindle with the increase in the concentration of the surfactant. The appearance of phase segregation discloses the increased desolvation of the polar parts of the surfactant. In this phenomenon, the H2O molecules spontaneously segregate, meaning that H2O molecules are taken away from micelles and are set apart, showing insolubility in the system at and above the particular temperature; thus, clouding phenomena is identified at that temperature for the particular system, viewing the maximum solubility of solutes at that fixed temperature.35 Therefore, the highest solubility of the clouding ingredient can be achieved at CP temperature; hence, the ΔG0c associated with clouding illustrates the phase changes from a single phase to two different phases. The estimated magnitudes of both ΔH0c and ΔS0c were found to be positive for pure TX-100 at almost all concentrations, which increased with the augmentation of the amount of surfactant. In the case of the LFH (drug) and TX-100 (19.95 × 10−3 and 30.76 × 10−3 mol·kg−1, respectively) mixed system in aqueous solution, the values of ΔG0c were found to be positive (Table 4), illustrating that clouding phenomena are still nonspontaneous. It was observed that the estimated values of

(4)

In the above equation, A, B, and C are constants that were calculated through the regression exploration of the least squares. The schematic graph of ln Xs versus TCP was drawn to calculate the enthalpy change (ΔH0c) of phase segregation phenomena, as shown in Figure 3. The plot of ln Xs versus T are shown in Figure 3 and found to be nonlinear. The obtained values of A, B, and C are reported in Table S4 (Supporting Information). Standard enthalpy change of the clouding phenomenon can be calculated accurately by inserting the value of eq 4 in eq 2 ΔH 0 c = RT 2[B + 2CT ]

ΔG0c (kJ·mol−1)

a

In the above equations, XS, R, and T are used to expound the concentration of the electrolyte present in the surfactant solution in mole fraction, universal gas constant, and CP temperature in Kelvin, respectively. The variation of CP with the Xs (solubility of the solute in mole fraction) can be expressed as a nonasymmetrical parabolic curve with the help of the following (eq 4)34 ln Xs = A + BT + CT 2

cTX‑100 (103) (mol·kg−1)

(5) E

DOI: 10.1021/acs.jced.9b00146 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Article

Table 4. Values of Thermodynamics Parameters of Phase Separation Phenomena for 19.95 and 30.76 mmol·kg−1 Surfactant Solutions plus Varying Concentration of LFH Medium at Pressure p = 0.1 MPaa cTX‑100 (103) (mol·kg−1)

cLFH (103) (mol·kg−1)

ΔG0c (kJ·mol−1)

ΔH0c (104) (kJ·mol−1)

ΔS0c (105) (J·mol−1·K−1)

ΔG0c,t (kJ·mol−1)

ΔH0c,t (105) (kJ·mol−1)

19.95

0.50 1.01 4.01 7.71 12.02 18.01 0.50 1.01 4.01 7.71 12.02 18.01

33.60 31.60 4.01 25.67 24.28 23.01 32.84 31.20 27.62 25.68 24.21 22.95

−72.34 −72.62 −73.30 −72.24 −71.28 −70.31 −134.03 −138.54 −144.18 −143.35 −140.16 −138.35

−20.80 −20.85 −20.98 −20.78 −20.60 −20.41 −39.43 −40.31 −41.39 −41.23 −40.62 −40.27

11.36 9.37 5.46 3.44 2.04 0.77 11.80 10.15 6.58 4.64 3.17 1.91

−9.66 9.68 −9.75 −9.65 −9.55 −9.45 −15.83 −16.28 −16.85 −16.76 −16.44 −16.26

30.76

Standard uncertainties (u) are u(T) = 0.1 K, u(c) = 0.02 mmol·kg−1, and u(p) = 5 kPa (level of confidence = 0.68). Relative standard uncertainties (ur) are ur(ΔG0c) = ±3%, u(ΔH0c) = ±3%, ur(ΔS0c) = ±4% , ur(ΔG0c,t) = ±4%, and ur (ΔH0c,t) = ±4%. a

Table 5. Values of Thermodynamics Parameters of Phase Separation Phenomena of Water + Salts Mixture Having 19.95 mmol·kg−1 TX-100 and 1.01 mmol·kg−1 LFH in Aqueous Solution of Salts at Pressure p = 0.1 MPaa csalt (mol·kg−1) KNO3 + H2O 0.000544 0.001039 0.003858 0.008209 0.012216 0.014936 K2SO4 + H2O 0.026839 0.031266 0.026982 0.024694 0.023507 0.022978 K3PO4·7H2O + H2O 0.000470 0.000918 0.003816 0.008292 0.012510 0.015064

ΔG0c (kJ·mol−1)

ΔH0c (104) (kJ·mol−1)

ΔS0c (104) (J·mol−1·K−1)

ΔG0c,t (104) (kJ·mol−1)

ΔH0c,t (104) (kJ·mol−1)

33.09 31.20 27.13 24.95 23.82 23.19

16.21 16.16 15.61 15.53 15.52 15.41

46.99 46.89 45.82 45.66 45.66 45.44

1.49 −0.40 −4.47 −6.66 45.66 −8.41

88.83 88.78 88.23 88.14 88.14 88.03

26.84 31.27 26.98 24.69 23.51 22.98

−84.67 −82.73 −80.56 −79.54 −79.18 −79.08

−246.22 242.46 −238.22 −236.21 −235.51 −235.31

−4.76 −0.34 −4.62 −6.91 −8.10 −8.63

−12.06 −10.11 −7.94 −6.92 −6.56 −6.46

33.63 31.36 27.17 25.13 23.81 23.12

10.06 9.93 9.66 9.58 9.38 9.29

28.98 28.73 28.20 28.04 27.66 27.48

2.02 −0.24 4.44 −6.48 −7.79 −8.49

82.68 82.55 82.27 82.19 82.00 81.91

Standard uncertainties (u) are u(T) = 0.1 K, u(c) = 0.02 mmol·kg−1 and u(p) = 5 kPa (level of confidence = 0.68). Relative standard uncertainties (ur) are ur(ΔG0c) = ±3%, u(ΔH0c) = ±3%, ur(ΔS0c) = ±4% , ur(ΔG0c,t) = ±4%, and ur(ΔH0c,t) = ±4%. a

ΔG0c dwindled with the enhancement of the concentration of a drug (LFH), signifying that the clouding phenomenon inclined to the spontaneity with the increase in the amount of LFH. At a low concentration of the LFH drug (0.50 mmol·kg−1), the clouding in TX-100 solution (19.95 mmol·kg−1) takes place at above 347.48 K, which is much higher than the cloud point temperature of pure TX-100 solution of the same concentration (337.15 K). This means that the aggregation is hindered; the process is made spontaneous by increasing the temperature (at which clouding occurs). Therefore, ΔG0c values are found to be positive at that particular temperature, meaning that this process is nonspontaneous in nature, and for achieving the CP, the temperature of the system need to be increased. For LFH + TX-100 mixed systems, the magnitudes of ΔH0c and ΔS0c were found to be negative in all cases, which decreased first with prolongation of drug concentration and

then increased with the further increase in the concentration of the drug. The negative magnitudes of ΔH0c signify that the clouding phenomenon is exothermic. Similar trends of ΔH0c and ΔS0c for clouding of the nonionic amphiphiles in the absence/presence of electrolytes were also reported in the literature.36 Rub et al.37−39 reported the negative values of both ΔH0c and ΔS0c for Tween 80 in the occurrence of a few amphiphilic drugs and different solute mixtures. In the case of drug and Tween 80 mixtures, akin result was also reported previously by Kabir-ud-Din et al.40 The clouding phenomena in the occurrence of diverse potassium salts in the solution of TX-100 + drug mixtures were also evaluated, and their various thermodynamics parameters values are viewed in Table 5. The magnitudes of ΔG0c for TX100 (19.95 × 10−3 mol·kg−1) and LFH (1.01 × 10−3 mol·kg−1) mixtures in the occurrence of different potassium salts (KNO3, F

DOI: 10.1021/acs.jced.9b00146 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Article

Table 6. Values of Thermodynamics Parameters of Phase Separation Phenomena of Water + Alcohols Mixed System Having 19.95 mmol·kg−1 TX-100 and 0 mmol·kg−1/1.01 mmol·kg−1 LFH in Aqueous Solution of Alcohols at Pressure p = 0.1 MPaa cdrug (mmol·kg−1) C2H5OH + H2O 0.00 0.00 0.00 1.01 1.01 1.01 C3H7OH + H2O 0.00 0.00 0.00 1.01 1.01 1.01 C4H9OH + H2O 0.00 1.01

calcohol (mol·kg−1)

ΔG0c (kJ·mol−1)

ΔH0c (103) (kJ·mol−1)

1.14243 2.41179 3.83049 1.14243 2.41179 3.83049

11.36 9.44 8.27 11.50 9.58 8.33

79.71 84.27 88.36 82.52 88.00 90.21

0.87588 1.84909 2.93677 0.87588 1.84909 2.93677

11.96 9.96 8.82 12.14 10.05 8.93

0.71009 0.71009

11.78 11.69

ΔS0c (104) (J·mol−1·K−1)

ΔG0c,t (kJ·mol−1)

ΔH0c,t (104) (kJ·mol−1)

22.77 23.61 24.35 82.52 24.28 24.68

−10.90 −12.83 −13.99 −20.10 −22.02 −23.27

−25.58 −25.13 −24.72 80.87 81.42 81.64

6725.84 6940.64 7294.99 7041.03 7118.16 7566.95

1948.55 1989.75 2056.78 2008.85 2023.47 2107.49

−10.31 −12.30 −13.44 −19.46 −21.56 −22.67

639.03 660.51 695.94 776.72 784.43 829.31

0.87 0.85

0.26 0.26

−10.48 −19.91

33.47 72.70

a Standard uncertainties (u) are u(T) = 0.1 K, u(C) = 0.02 mmol·kg−1, and u(p) = 5 kPa (level of confidence = 0.68). Relative standard uncertainties (ur) are ur(ΔG0c) = ±3%, u(ΔH0c) = ±3%, ur(ΔS0c) = ±4% , ur(ΔG0c,t) = ±4%, and ur(ΔH0c,t) = ±4%.

K2SO4 and K3PO4·7H2O) were found to be positive, illustrating that the clouding phenomena in such a system are also not spontaneous in character. These observed positive values were found to dwindle with the increase in the concentration of electrolytes, which illustrates the propensity from nonspontaneous to spontaneous phenomena with the augmentation of the concentration of electrolytes. However, the magnitudes of ΔH0c and ΔS0c are found to be positive for KNO3 and K3PO4·7H2O for different concentrations of salts at 19.95 mmol·kg −1 TX-100 plus 1.01 mmol·kg −1 LFH concentrations, but their values decrease with increasing concentration of electrolytes. In the case of K2SO4, the values of ΔH0c and ΔS0c are found to be negative but increases through an increasing amount of K2SO4 (Table 5). The obtained negative value of ΔH0c almost certainly will come about as the arrangement of water particles surrounding the polar head group portion happens with greater consideration, as compared to the destruction of water arrangement surrounding the nonpolar parts of the surfactant.31,41,42 The obtained outcomes show that the values of ΔG0c were directed mutually by ΔH0c and ΔS0c at the phase segregation of TX-100 in the presence of electrolytes. In the case of different alcohols (C2H5OH, C3H7OH, and C4H9OH), the values of ΔG0c were observed to be positive and reduced with increasing alcohol content in the absence/ presence of a drug (Table 5). ΔH0c and ΔS0c were found to be positive for each alcohol of different concentrations for pure 19.95 mmol·kg−1 surfactant and in the occurrence of 1.01 × 10−3 mol·kg−1 LFH, which increased with the augmentation of the concentration of alcohol. The development of positive ΔH0c values is due to the disruption of the H2O structure around the nonpolar parts of the surfactant monomers.43 The negative values of ΔH0c indicate the existence of attractive forces (London dispersion forces) during the association of the surfactant systems,44,45 while the positive magnitudes of ΔH0c indicate the destruction of the iceberg structure of H2O around the hydrophobic parts of the surfactants.46−49

The thermodynamics properties of transfer, for instance, standard free energy (ΔG0c,t) and enthalpy change (ΔH0c,t) of transfer, during the phase segregation of the surfactant from aqueous solution to medium with additives were obtained utilizing the following (eqs 6 and 7)50−54 ΔG 0 c,t = ΔG 0 c(aq. of additive) − ΔG 0 c(aq. )

(6)

ΔH 0 c,t = ΔH 0 c(aq. of additive) − ΔH 0 c(aq. )

(7)

clouding (ΔG0c,t) for −3 −1

The free energy of transfer of the TX100 (19.95 × 10−3 and 30.76 × 10 mol·kg ) and LFH mixed systems were found to be positive at all concentrations of LFH, which decreased with increasing concentration of LFH. In the presence of salts (KNO3 and K3PO4·7H2O), the values of free energy of transfer (ΔG0c,t) for the TX-100 + LFH mixture were found to be positive at a minimum applied concentration of electrolytes but were negative at all other concentrations (Table 5). These negative values of ΔG0c,t were then enhanced with the augmentation of the concentration of the electrolytes (KNO3 and K3PO4·7H2O). In the case of K2SO4 (Table 5), the magnitudes of free energy of transfer (ΔG0c,t) for the TX-100 + LFH mixture were found to be negative at every concentration of electrolytes, which further enhances with the augmentation of the concentration of the electrolytes. In the presence of alcohols (C2H5OH and C3H7OH), the values of free energy of transfer (ΔG0c,t) for the TX-100 + LFH mixture were found to be negative at the entire concentration of alcohol, which enhances with the augmentation of the concentration of alcohol (Table 6). In the case of C4H9OH, the free energy of transfer (ΔG0c,t) was found to be negative in the presence and absence of a drug, but more values could not be determined. The enthalpy of transfer of clouding (ΔH0c,t) for TX-100 (19.95 × 10−3 and 30.76 × 10−3 mol·kg−1) and LFH mixed systems was found to be negative at all concentrations of LFH, which increased with increasing concentration of LFH. In the presence of KNO3 and K3PO4·7H2O, the magnitudes of enthalpy transfer (ΔH0c,t) for the TX-100 + LFH mixed system were found to be positive (Table 5) in all cases, which G

DOI: 10.1021/acs.jced.9b00146 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data



decreased through growth of the concentration of the electrolytes, but in the case of K2SO4 (ΔH0c,t), the values were found to be negative and increased with increasing concentration of the electrolytes. In the presence of C2H5OH (Table 6), the values of enthalpy of transfer (ΔH0c,t) of the TX-100−water system were found to be negative in the absence of a drug and positive in the presence of a drug and at all concentrations of alcohol, and the values increased with increasing concentration of C2H5OH. Also, in the case of C3H7OH, the values of enthalpy of transfer (ΔH0c,t) were positive in the presence and absence of a drug and at all concentrations of alcohol, and the values increased with increasing concentration of C3H7OH.

ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.jced.9b00146. Data for the cloud point of some studied systems (PDF)



REFERENCES

(1) Rosen, M. J. Surfactants and Interfacial Phenomenon, 3rd ed.; Wiley: New York, 2004. (2) 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. (3) Kumar, D.; Rub, M. A. Synthesis and characterization of dicationic gemini surfactant micelles and their effect on the rate of ninhydrin−copper-peptide complex reaction. Tenside Surfactants Deterg. 2018, 55, 78−84. (4) Kumar, D.; Rub, M. A. Studies of interaction between ninhydrin and Gly-Leu dipeptide: Influence of cationic surfactants (m-s-m type Gemini). J. Mol. Liq. 2018, 269, 1−7. (5) Buckingham, L. E.; Balasubramanian, M.; Emanuele, R. M.; Clodfelter, K. E.; Coon, J. S. Comparison of solutol HS 15, Cremophor EL and novel ethoxylated fatty acid surfactants as multidrug resistance modification agents. Int. J. Cancer 1995, 62, 436−442. (6) Attwood, D.; Florence, A. T. Surfactant Systems; Their Chemistry, Pharmacy and Biology; Chapman & Hall: New York, 1983. (7) Rub, M. A.; 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. (8) Rub, M. A.; Sheikh, M. S.; Khan, F.; Khan, S. B.; Asiri, A. M. Bile salts aggregation behavior at various temperatures under the influence of amphiphilic drug imipramine hydrochloride in aqueous medium. Z. Phys. Chem. 2014, 228, 747−767. (9) 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. (10) Khan, M. A. R.; Amin, M. R.; Mahbub, S.; Alam, M. M.; Rub, M. A.; Hoque, M. A.; Khan, M. A.; Asiri, A. M. Effect of various electrolytes on the phase separation and thermodynamic properties of p-tert-alkylphenoxy poly (oxyethylene) ether in the absence/presence of drugs. J. Surfactants Deterg. 2019, DOI: 10.1002/jsde.12248. (11) Schott, H. Effect of inorganic additives on solutions of nonionic surfactants-XVI. Limiting cloud points of highly polyoxyethylated surfactants. Colloids Surf., A 2001, 186, 129−136. (12) Iwanaga, T.; Kunieda, H. Effect of added salts or polyols on the cloud point and the liquid-crystalline structures of polyoxyethylenemodified silicone. J. Colloid Interface Sci. 2000, 227, 349−355. (13) Inoue, T.; Ohmura, H.; Murata, D. Cloud point temperature of polyoxyethylene-type nonionic surfactants and their mixtures. J. Colloid Interface Sci. 2003, 258, 374−382. (14) Myers, D. Surfactant Science and Technology, 3rd ed.; Oxford University Press: New York, 2005. (15) Alauddin, M.; Parvin, T.; Begum, T. Effect of organic additives on the cloud point of Triton X-100 micelles. J. Appl. Sci. 2009, 9, 2301−2306. (16) Mudawadkar, A. D.; Sonawane, G. H.; Patil, T. J. Thermodynamics of micellization of nonionic surfactant Triton X100 in presence of additive Poly-N-vinyl-pyrrolidone using clouding phenomenon. Orient. J. Chem. 2013, 29, 227−233. (17) Shinoda, K.; Makagawa, T.; Tanamushi, B.; Ishemushi, T. Colloidal Surfactants; Academic Press: New York, 1963. (18) Panchal, K.; Desai, A.; Nagar, T. Physicochemical behavior of mixed nonionic-ionic surfactants in water and aqueous salt solutions. J. Dispersion Sci. Technol. 2006, 27, 33−38. (19) Bhatt, D.; Maheria, K. C.; Parikh, J. Studies on surfactant-ionic liquid interaction on clouding Behaviour and evaluation of thermodynamic parameters. J. Surfactants Deterg. 2013, 16, 547−557. (20) Mahajan, R. K.; Vohra, K. K.; Kaur, N.; Aswal, V. K. Organic additives and electrolytes as cloud point modifiers in octylphenol ethoxylate solutions. J. Surfactants Deterg. 2008, 11, 243−250. (21) Huang, Z.; Gu, T. The effect of mixed cationic-anionic surfactants on the cloud point of non-ionic surfactant. J. Colloid Interface Sci. 1990, 138, 580−582.

4. CONCLUSIONS In the current investigation, the drug/(drug + salts)/(drug + alcohols)-mediated phase partition phenomena of TX-100 have been carried out. The variation of CP values of the surfactant in the occurrence of LFH/(LFH + electrolytes)/ (LFH + alcohols) reveals that a significant amount of drug/ electrolytes/alcohols interact with the surfactant molecules. The CP values of the surfactant + LFH increase in the presence of salts and obey the order K3PO4 > KNO3 > K2SO4, and in the case of alcohols, CP values increase for C2H5OH and C3H7OH but largely decrease in the case of C4H9OH. In the case of the high percentage of C4H9OH, CP value calculation was not possible because the CP values were lesser at room temperature. In the case of C2H5OH and C3H7OH, CP values increased with increasing concentration. The enthalpy change values of clouding (ΔH0c) and entropy change values of clouding (ΔS0c) were found to be positive, which discloses the occurrence of endothermic interaction during the phase segregation of pure TX-100 in an aqueous medium. The values of ΔG0c are found to be positive in the absence and presence of drug/(drug + salt)/(drug + alcohols), which signifies that the phase separation phenomenon is nonspontaneous in nature. The ΔG0c,t and ΔH0c,t values were evaluated in the presence of all additives and discussed in detail.



Article

AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. Tel.: +84 943720085. ORCID

Malik Abdul Rub: 0000-0002-4798-5308 Dileep Kumar: 0000-0003-2913-5032 Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS M.A.H. would like to acknowledge Jahangirnagar University, Savar, Dhaka, Bangladesh for the providing financial support to carry out the research work and the General Pharmaceutical Ltd., Bangladesh for supplying the standard sample of LFH as a gift item. H

DOI: 10.1021/acs.jced.9b00146 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Article

(22) Mahajan, R. K.; Chawla, J.; Bakshi, M. S. Depression in the cloud point of tween in the presence of glycol additives and triblock polymers. Colloid Polym. Sci. 2004, 282, 1165−1168. (23) Arai, H.; Murata, M.; Shinoda, K. The interaction between polymer and surfactant: The composition of the complex between polyvinylpyrrolidone and sodium alkyl sulfate as revealed by surface tension, dialysis, and solubilization. J. Colloid Interface Sci. 1971, 37, 223−227. (24) Rahman, M.; Khan, M. A.; Rub, M. A.; Hoque, M. A.; Asiri, A. M. Investigation of the effect of various additives on the clouding behavior and thermodynamics of polyoxyethylene (20) sorbitan monooleate in absence and presence of ceftriaxone sodium trihydrate drug. J. Chem. Eng. Data 2017, 62, 1464−1474. (25) Aktar, S.; Molla, R. M.; Mahbub, S.; Rub, A. M.; Hoque, M. A.; Islam, D. M. S. Effect of temperature and salt/alcohol on the interaction of tetradecyltrimethylammonium bromide/ Triton X-100 with moxifloxacin hydrochloride: A multitechnique approach. J. Dispersion Sci. Technol. 2018, DOI: 10.1080/01932691.2018.1472005. (26) Kumar, D.; Hidayathulla, S.; Rub, M. A. Association behavior of a mixed system of the antidepressant drug imipramine hydrochloride and dioctyl sulfosuccinate sodium salt: Effect of temperature and salt. J. Mol. Liq. 2018, 271, 254−264. (27) Muherei, M. A.; Junin, R. Investigating Synergism in Critical Micelle Concentration of Anionic-Nonionic Surfactant Mixtures: Surface versus Interfacial Tension Techniques. Asian J. Appl. Sci. 2009, 2, 115−127. (28) Rub, M. A.; Azum, N.; Asiri, A. M. Interaction of cationic amphiphilic drug nortriptyline hydrochloride with TX-100 in aqueous and urea solutions and the studies of physicochemical parameters of the mixed micelles. J. Mol. Liq. 2016, 218, 595−603. (29) Pendegrass, C. J.; Gordon, D.; Middleton, C. A.; Sun, S. N. M.; Blunn, G. W. Sealing the skin barrier around transcutaneous implants: in vitro study of keratinocyte proliferation and adhesion in response to surface modifications of titanium alloy. J. Bone Jt. Surg. Br. Vol. 2008, 90, 114−121. (30) Ferguson, M. A.; Alderson, N. L.; Trost, S. G.; Essig, D. A.; Burke, J. R.; Durstine, J. L. Effects of four different single exercise sessions on lipids, lipoproteins, and lipoprotein lipase. J. Appl. Physiol. 1998, 85, 1169−1174. (31) Khan, M. B.; Hoque, M. A.; Islam, D. M. S. Physicochemical investigation of the clouding behavior and thermodynamics of p-tertalkylphenoxy poly (oxyethylene )ether micelles in aqueous environment and in the presence of diols. J. Chem. Thermodyn. 2015, 89, 177−182. (32) Sharma, K. S.; Patil, S. R.; Rakshit, A. K. Study of the cloud point of C12En nonionic surfactants: Effect of additives. Colloids Surf., A 2003, 219, 67−74. (33) Rahman, M.; Hoque, M. A.; Khan, M. A.; Rub, M. A.; Asiri, A. M. Effect of different additives on the phase separation behavior and thermodynamics of p-tert-alkylphenoxy poly (oxyethylene) ether in absence and presence of drug. Chin. J. Chem. Eng. 2018, 26, 1110− 1118. (34) Blankschtein, D.; Thurston, G. M.; Benedek, G. B. Phenomenological theory of equilibrium thermodynamic properties and phase separation of micellar solutions. J. Chem. Phys. 1986, 85, 7268−7288. (35) Da Silva, R. C.; Loh, W. Effect of additives on the cloud points of aqueous solutions of ethylene oxide − propylene oxide−ethylene oxide block copolymers. J. Colloid Interface Sci. 1998, 202, 385−390. (36) Heusch, R. Structures in surfactant/water mixtures and their use in biotechnology. Biotech. Forum 1986, 3, 1−8. (37) Kabir-ud-Din; Rub, M. A.; Naqvi, A. Z. Cloud point modulation of an antidepressant drug imipramine hydrochloride with pharmaceutical excipients and the thermodynamics thereon. J. Dispersion Sci. Technol. 2012, 33, 1667−1673. (38) Azum, N.; Rub, M. A.; Asiri, A. M. Energetics of clouding phenomenon in amphiphilic drug imipramine hydrochloride with pharmaceutical excipients. Pharm. Chem. J. 2014, 48, 201−208.

(39) Naqvi, A. Z.; Rub, M. A.; Kabir-ud-Din. Study of phospholipidinduced phase-separation in amphiphilic drugs. Colloid J. 2015, 77, 525−531. (40) Kabir-ud-Din; Rub, M. A.; Sheikh, M. S. Cloud-Point modulation of an amphiphilic drug with Pharmaceutical excipients. J. Chem. Eng. Data 2010, 55, 5642−5652. (41) Parikh, J.; Rathore, J.; Bhatt, D.; Desai, M. Clouding behavior and thermodynamic study of nonionic surfactants in presence of additives. J. Dispersion Sci. Technol. 2013, 34, 1392−1398. (42) Kumar, D.; Rub, M. A. Effect of anionic surfactant and temperature on micellization behavior of promethazine hydrochloride drug in absence and presence of urea. J. Mol. Liq. 2017, 238, 389− 396. (43) Alam, M. S.; Kabir-ud-Din; Mandal, A. B. Evaluation of thermodynamic parameters of some amphiphilic drugs in presence of sugars at the cloud point. Colloids Surf., B 2013, 105, 236−245. (44) Hierrezuelo, J.; Molina-Bolívar, J.; Ruiz, C. An energetic analysis of the phase separation in non-ionic surfactant mixtures: The role of the headgroup structure. Entropy 2014, 16, 4375−4391. (45) Mahbub, S.; Rub, M. A.; Hoque, M. A.; Khan, M. A. Mixed micellization study of dodecyltrimethylammonium chloride and cetyltrimethylammonium bromide mixture in aqueous/urea medium at different temperatures: Theoretical and experimental view. J. Phys. Org. Chem. 2018, 31, No. e3872. (46) Clint, J. H. Surfactant Aggregation; Chapman and Hall: New York, 1992. (47) Robins, D. C.; Thomas, I. L. The effect of counterions on micellar properties of 2-dodecylaminoethanol salts: I. Surface tension and electrical conductance studies. J. Colloid Interface Sci. 1968, 26, 407−414. (48) Kumar, D.; Rub, M. A. Aggregation behavior of amphiphilic drug promazine hydrochloride and sodium dodecylbenzenesulfonate mixtures under the influence of NaCl/urea at various concentration and temperatures. J. Phys. Org. Chem. 2016, 29, 394−405. (49) 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, No. e3812. (50) Kresheck, G. C. In Water: A Comprehensive Treatise, Franks, F., Ed.; Plenum: New York, 1995. (51) Rahman, M.; Khan, M. A.; Rub, M. A.; Hoque, M. A. Effect of temperature and salts on the interaction of cetyltrimethylammonium bromide with ceftriaxone sodium trihydrate drug. J. Mol. Liq. 2016, 223, 716−724. (52) Hoque, M. A.; Ahmed, F.; Halim, M. A.; Molla, M. R.; Rana, S.; Rahman, M. A.; Rub, M. A. Influence of salt and temperature on the interaction of bovine serum albumin with cetylpyridinium chloride: Insights from experimental and molecular dynamics simulation. J. Mol. Liq. 2018, 260, 121−130. (53) Hoque, M. A.; Alam, M. M.; Molla, M. R.; Rana, S.; Rub, M. A.; Halim, M. A.; Khan, M. A.; Akhtar, F. Interaction of cetyltrimethylammonium bromide with drug in aqueous/ electrolyte solution: A combined conductometric and molecular dynamics method study. Chin. J. Chem. Eng. 2018, 26, 159−167. (54) Khan, M. A. R.; Amin, M. R.; Rahman, M.; Rub, M. A.; Hoque, M. A.; Khan, M. A.; Asiri, A. M. Influence of different additives on the clouding nature and thermodynamic behavior of tween 80 solution in the absence and presence of the amikacin sulfate drug. J. Chem. Eng. Data 2019, 64, 668−675.

I

DOI: 10.1021/acs.jced.9b00146 J. Chem. Eng. Data XXXX, XXX, XXX−XXX