Next Article in Journal
Dynamics of Opposing Polymer Brushes: A Computer Simulation Study
Next Article in Special Issue
Molecular Assembly in Block Copolymer-Surfactant Nanoparticle Dispersions: Information on Molecular Exchange and Apparent Solubility from High-Resolution and PFG NMR
Previous Article in Journal
Healing Efficiency of CNTs-Modified-UF Microcapsules That Provide Higher Electrical Conductivity and EMI Shielding Properties
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Mixed Micellization and Spectroscopic Studies of Anti-Allergic Drug and Non-Ionic Surfactant in the Presence of Ionic Liquid

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.
Submission received: 6 July 2021 / Revised: 9 August 2021 / Accepted: 9 August 2021 / Published: 17 August 2021

Abstract

:
In drug delivery, surfactants are used to reduce side effects and to increase drug efficiency. The present work aimed to study the interaction of diphenhydramine hydrochloride (anti-allergic drug) with TX–45 (non-ionic surfactant) in the absence and presence of ionic liquid (1-hexyl-3-methylimidazolium chloride). The physicochemical parameters were estimated by the surface tension measurement. Various theoretical models (Clint, Rubingh, Motomura, and Maeda) were applied to determine the attractive behavior between drug and surfactant mixtures at the surface and in bulk. The drug and surfactant mixtures exhibit synergistic behavior in the absence and presence of ionic liquid. Several energetic parameters were also estimated with the assistance of regular solution approximation and pseudo phase separation model that indicate micelle formation and adsorption of surfactant at the surface is thermodynamically advantageous. The morphology of pure and mixture of amphiphiles has been estimated by the Tanford and Israelachvili theories. UV-visible spectroscopy was used to quantify the attractive behavior of the drug with surfactant with the help of a binding constant (K).

Graphical Abstract

1. Introduction

Drug delivery is the process of administrating drugs to achieve therapeutic effects. From time to time, the researchers try to improve and refine the drug delivery system to increase drug bioavailability, lessen drug deprivation and loss, and thus prevent destructive side effects [1,2,3,4,5]. Numerous drug carriers (surfactants, liposomes, polymers, nanomaterials, and microspheres) have been utilized for drug delivery systems [6,7,8]. Low bioavailability due to the poor solubilization of hydrophobic drugs is the main problem associated with drug therapy. Due to this poor availability, a large quantity of drugs is required to complete the therapy. As a result, accumulation of drug at non-target sites increases the harmful side effects [9]. Surfactants can act as a delivery agent that not only solubilizes the drug but also raises the bioavailability of drug molecules. Surfactants or surface-active agents are amphipathic molecules having polar and non-polar parts in the same body. Due to this dual character, they have one of the exiting features of self-aggregation or micelle formation. The surfactants aggregate into a globular form (micelles) at a specific concentration known as critical micelle concentration (CMC) [10]. Micelles are labile entities formed by the noncovalent association of individual surfactant monomers. Micelles are structurally like lipids having a hydrophilic interface and hydrophobic inner core in a polar solvent. Drugs that cannot normally cross cell membrane can easily pass by being encapsulated in a micelle [11,12,13]. Micelles have several advantages in being used as a drug carrier. (i) They are small in size; (ii) they mostly have low toxicity (iii) they can rest for an extended period in the system (iv) they can solubilize poorly soluble drugs and lessen the harmful side effects of the drugs; (v) they can be arranged in huge quantities effortlessly; and (vi) micelles can form naturally and are energetically steadier regarding association and dissociation.
The mixture of one or more surfactants is most important to improve self-assembly and their application. In drug delivery, a mixture of surfactants is used to reduce the side effects and to increase the drug efficiency [14]. This enhanced performance of mixed surfactants is due to the interaction among the two or more surfactants. To interpret the effect of mixed surfactants systems on the physicochemical properties of a drug, various research papers have been published [15,16,17,18,19,20,21,22,23]. Earlier, we also investigated the mixed system of numerous amphiphilic drugs with different surfactants or amphiphiles using different techniques [24,25,26,27,28,29,30,31,32,33].
An antihistamine drug (Diphenhydramine hydrochloride, DPH) used in this study can block the proclamation of histamine by histamine-1 receptors and the action of histamine in the body. It is a first-generation anti-allergic drug utilized in cough syrup and tablets [34]. Being first-generation, it has great efficiency in curing allergic problems as compared to a second-generation anti-allergic drug (desloratadine) [35]. Likewise, DPH has anticholinergic, sedative, antiemetic, and antitussive properties. It also acts as a local anesthetic agent because it is an intracellular sodium channel blocker. Like many other surface-active compounds, DPH also shows surface-active properties. The monomers of DPH molecules aggregate to form a micelle but having a low aggregation number [36]. The association of these types of molecules is mainly due to the hydrophobic interaction.
In recent years, a new class of solvents known as ionic liquids (ILs) has gained much attention in the scientific field. The ionic liquid can be defined as the salt in a liquid state that melts at a temperature below 100 °C. These compounds are stable over a wide range of temperatures and have low melting and vapor pressures [37,38,39,40,41]. ILs have been extensively used in the areas of electrochemistry [41], chemical separation [42], polymer chemistry [43], organic synthesis [44], nanomaterials [45], bio-catalysis, and solar cells [46]. ILs having a long alkyl chain show surface activity and behave like a surfactant. ILs, being environmentally benign and amphiphilic, have good scope in pharmaceutical science since they can enhance the permeability of drugs across biological membranes [47]. Based on their physicochemical properties and structural aspects, they are also used as drug reservoirs, excipients and co-solvent.
The main objective of the present work is to investigate the mixed micellization and absorption behavior of antiallergic drug (DPH) in the presence of nonionic surfactant (TX–45). For the future use of such micelles as drug carriers (they are cost-effective and less toxic in comparison to other alternatives), the physicochemical properties of these mixtures are required before their application. Therefore, the present work is an initiative toward amphiphiles-based drug delivery. The pharmaceutical actions of drugs appear to be connected to the drug-membrane interaction. The physicochemical properties of drugs with surfactant micelles can be envisaged as an estimation for their interactions with biological surfaces; thus, amphiphiles micelles have been utilized as mimetics for biomembrane. The literature presents wide understanding of amphiphiles with drugs but there few attempts have been made to date to study the interactions between the drug and amphiphiles in the company of IL. The work on IL has inspired us to discover the effects of IL on the interaction between drug and amphiphiles. In surface science, the novel IL with a 1-alkyl-3-methyl imidazolium cation is used [48,49,50]. This IL has several advantages compared to conventional surfactant; the broadhead group is sturdily striking through π–π interactions and has a solubilizing capacity for a broader range of solutes. Therefore, here, we performed tensiometry and UV-visible spectroscopic measurements to explore the interaction between an anti-allergic drug (DPH) and a non-ionic surfactant (TX–45) in the absence and presence of IL. Various physicochemical parameters were evaluated by using different hypothetical theories. The interaction between the drug and surfactant was also studied by UV-visible spectroscopy.

2. Materials and Methods

The anti-allergic drug, diphenhydramine hydrochloride, was the product of Molecule-on (New Lynn, Auckland, New Zealand). The non-ionic surfactant, TX–45, and ionic liquid, 1-hexyl-3-methylimidazolium chloride (IL), utilized in this study were purchased from Sigma (St. Louis, MO, USA). The chemicals were utilized as obtained from the supplier, without of additional purification. De-ionized and double distilled water was utilized to prepare all solutions. Specific conductivity was used to check the purity of water (1–6 µS cm−1). The pH values were measured by a pH meter (7). The structure of the drug, surfactant, and ionic liquid are given in Figure 1.

2.1. Surface Tension Measurements

The surface tension of DPH and the mixture with TX–45 in aqueous solution were recorded by utilizing an attention tensiometer (Sigma 701, Darmstadt, Germany) over the temperature 298.15 K (Figure 2). A Citizen CX 220 analytical lab balance was used to weigh compounds for the preparation of the solutions. The instrument was calibrated at regular intervals by ultrapure water. The ring was heated on the ethanol flame until it glowed red before each experiment and, after that, washed with ultra-pure water. In our experiment, we follow the Du Nouy ring method, in which a platinum ring is utilized for the measurement of surface tension.

2.2. UV-Visible Spectroscopy

A Thermo Scientific, Waltham, MA, USA, Evolution 300 UV-visible spectrophotometer was used to carry out the experiments. All spectra were noted in the spectral range of 230 to 350 nm (Figure 3). The stock solution of DPH (5 mM) was constructed in water. TX–45 solution (5 mM) was prepared in an already prepared drug solution to keep the drug concentration constant throughout. The spectrophotometer was used to analyze these already prepared solutions to observe the effect of TX–45 concentrations on the UV-visible spectra of DPH. To observe the ionic liquid effects, all the solutions were prepared in 25 mM aqueous IL solution instead of water.

3. Results and Discussion

3.1. Determination of Critical Micelle Concentration (CMC)

The CMC of pure and mixed systems was computed by plotting the values of surface tension (γ) against the concentration (Figure 2). Surface tension is one of the most prominent techniques used to understand the surface and bulk properties of a surface-active compound. The decrease in surface tension after the addition of an amphiphile is due to the displacement of some water molecules by amphiphiles. A plateau region is obtained after the addition of an adequate amount of amphiphiles, which depicts micelle formation. After a critical point, the surface tension values are not constant but slightly decrease due to the existence of impurities. The CMC values of pure and mixed amphiphiles were estimated from the juncture of the two lines manually drawn indicating the high and low concentration regions (Figure 2). The CMC values computed from these curves are given in Table 1. The DPH has higher CMC values than the TX–45, due to the complex structure and ionic nature of DPH. The values of CMC of pure constituents are well-matched with reported values [51,52]. The values of CMC of pure amphiphiles in the presence of 25 mM IL are shown in Table 1. The CMC values of pure amphiphiles (DPH and TX–45) decrease in the presence of 25 mM IL. The IL and DPH have the same charge on the head group, which results in two amphiphiles not interacting appropriately. The counter ions of IL and DPH also lead to decrease in the head group repulsion so that molecules can originate nearer to each other and thus micelle formation can occur at a lower concentration of the drug. Rajput et al., and Pal et al., also described similar behavior [53,54]. However, ion–dipole and hydrogen bonding are responsible for the TX–45 being non-ionic [55].

3.2. Clint Theory of Mixed Micellization

The ideal values of CMC (CMC*) were computed by considering the ideal behavior of mixed current systems by Equation (1).
1 CMC * = α 1 CMC 1 + α 2 CMC 2
The ideal values along with experimental values at numerous mole fractions of surfactant are listed in Table 1. The nonideal behavior and a constitutional interaction between two amphiphiles are clear from lower values of the experimental CMC than the ideal one, CMC* (Table 1). The experimental CMC values decline with the molar fraction of surfactant (α1). As the molar fraction of surfactant in binary mixture increases, the electrostatic self-repulsion between the ionic head group of DPH decreases results in a decline in the CMC values of mixed systems. The viewpoint can be supported by the previous investigations [56]. The negative deviation of experimental CMC and ideal CMC in the whole fraction range indicates interaction between the mixed components. In attendance of 25 mM IL, the experimental CMC values were also found to be lower than ideal.

3.3. Rubingh Theory of Mixed Micellization

For a non-ideal mixed micellar system, the CMC of a mixture can be calculated by the Rubingh theory.
1 CMC = α 1 f 1 CMC 1 + α 2 f 2 CMC 2
The activity coefficients (f1 and f2) of amphiphiles can be calculated by the following equations
f 1 = e x p [ β ( 1 X 1 ) 2 ]
f 2 = e x p [ β ( X 1 ) 2 ]
The interaction between two amphiphiles in a mixed micelle can be well understood by analyzing regular solution theory. The micellar mole fraction of amphiphile 1 can be computed by the Rubingh model:
( X 1 ) 2 ln ( α 1 CMC / X 1 CMC 1 ) ( 1 X 1 ) 2 ln [ ( 1 α 1 ) CMC / ( 1 X 1 ) CMC 2 ] = 1
The values of micellar mole fraction of component 1 ( X 1 ) are listed in Table 1. On the other hand, the ideal involvement of TX–45 in the DPH + TX–45 mixed micelle can be computed with the help of the micellar mole fraction of surfactant in the ideal state ( X 1 ideal ) as follows [57]:
X 1 ideal = α 1 CMC 2 α 1 CMC 2 + α 2 CMC 1
The X 1 ideal values are given in Table 1. It is clear from the data that the values X 1 deviate negatively from the ideal values ( X 1 ideal ). The higher values of X 1 ideal designate that the DPH + TX–45 mixture has a high involvement of drug components than the surfactant. The same results were obtained in the presence of IL. The Rubingh model utilizing the regular solution theory (RST) confirms the attractive or repulsive interaction between two amphiphiles by an interaction parameter β, calculated by Equation (7):
β = ln ( CMC α 1 / CMC 1 X 1 ) ( 1 X 1 ) 2
The physicochemical interaction between the components before mixing and after mixing at the same condition can be easily understood by the interaction parameter (β). When the β values are close to zero, means there is no attractive behavior while positive and negative values are a sign of repulsive and attractive behavior of components in a mixed system. According to data in Table 1, values of β are negative for the current mixed system at all mole fractions. The non-ionic surfactant (TX–45) has a huge quantity of oxygen atoms wearing unpaired electrons. Therefore, TX–45 molecules tend to interact with positive charge drug molecules coulombically. Table 1 also listed the β value in the attendance of IL. In the attendance of IL, β is also negative, confirm synergistic behavior. The requirement of synergism for a mixed system is as follows:
(a) β = negative value
(b) β > ln (CMC1/CMC2)
It is clear from Table 1 that both the above conditions are followed by our current mixed systems in the nonattendance and attendance of IL. The extent of interaction can also be judged by the most effective parameters and the activity coefficient (Equations (3) and (4)). The computed values of activity coefficients are less than one, showing the nonideality of the current mixed system.

3.4. Interfacial Properties of Drug/Surfactant Mixture

When an amphiphile solution is added into the water, the monomers are oriented at the surface and the surface tension of water decreases. The reduction in surface tension is due to the rupturing of hydrogen bonding between water molecules at the surface. If we add more solution, then the further addition gives rise to more reduction in the surface tension values until saturation point. The quantity of amphiphile adsorbed at an interface per unit area can be computed by using Gibbs’ adsorption isotherm. The value of surface excess for the DPH + TX–45 mixture can be calculated by using Equation (10):
Γ m a x = 1 2.303 n R T ( d γ d l o g C )
where C, R, and T show the molar concentration, gas constant, and absolute temperature, respectively. n stands for the number of constituents at the interface that differ from the surfactant bulk concentration. The values for n are taken as three, two and one for the mixture, drug, and TX–45, respectively. Γ m a x is a measure of the effectiveness of the surfactant adsorption at the surface. Higher values of Γ m a x mean extreme stuffing and powerfully squeezing of amphiphiles molecules at the surface. Some surface properties like wetting, foaming, and emulsification are strongly dependent on the effectiveness of adsorption. With the help of Γ m a x values from Equation (10), the minimum area per molecule was also computed:
A m i n = 10 18 N A Γ m a x   ( nm 2 )
where NA is Avogadro’s number. The values of Γ m a x and A m i n are listed in Table 2.
The Γ m a x value for TX–45 is more than DPH because the long chain has greater surface activity. The Γ m a x values for mixtures are lesser than the values of single amphiphiles. This can be attributed the lowest power of the mixed system to pump the monomers to the surface, which leads to a decline in the concentration of monomers at the surface. The increase in A m i n for a mixed system suggests a loose mixed monolayer at the surface is formed. However, in the presence of IL, the Γ m a x value slightly decreases (increase in A m i n ) for the current mixture. The upsurge or reduction in the values of Γ m a x depend upon numerous aspects, i.e., charge on the head group of the host molecule, hydrogen bonding of IL with the solvent and the occurrence of IL monomers at the interface. The decrease in surface activity of the drug can be understood with the fact that both drug and IL are positively charged, so these molecules of similar charge try to remain separate from each other at the surface monolayer. Thus, the space between monomers at the surface increases ( A m i n increases) and the surface activity decreases [58]. Another reason for the decreasing Γ m a x value after the addition of IL may be the partial adsorption of IL at the surface. IL is also surface-active molecules and have a smaller surface tension value than water. At the interface, some space is occupied by the IL molecules; as a result, the space for drug or other amphiphiles decreases [59]. Adsorption efficiency ( p C 20 ) is calculated by the equation:
p C 20 = l o g C 20
where C20 is the efficiency of adsorption of surfactants at the surface. It is the concentration that is mandatory to decrease the surface tension value by 20 mN m−1. The values of p C 20 are listed in Table 2. The current mixed system has larger values of p C 20 , demonstrating that the efficiency of the mixture is greater than the drug. The efficiency of amphiphile was also measured by parameter πCMC, computed by Equation (13):
π CMC = γ 0 γ CMC
where γ0 is the surface tension of water, and γCMC is the surface tension at CMC. Higher values of πCMC also confirm the earlier explanation. The IL enhances the efficiency of the drug and drug + TX–45 mixed system. Adsorption of IL at the surface offers an additional hydrophobic atmosphere for the drug, raising its propensity for surface adsorption.

3.5. Morphology of Drug/Surfactant Mixture

Israelachvili et al. described a relationship for computing the packing parameter (P) by using the surface area of amphiphiles [60]:
P = V 0 A m i n l c
where lC stands for the extreme operative length of the hydrophobic tail of the monomer, and V0 is the hydrophobic chain volume computed by using Tanford equations [61]
l C = ( 0.154 + 0.1265 ( C n 1 ) )   nm
V C = ( 0.0274 + 0.0269 ( C n 1 ) )   nm 3
The values of P are listed in Table 3. The measurement of the head group (A) is quite difficult, so in our calculations we used A m i n values instead of A. The morphology of a system can be predicted by the value of P. For a spherically shaped micelle, the value of P should be ≤0.333. The cylindrical (0.333 < P < 0.5), vesicle and bilayer (0.5 < P < 1) shapes of micelle also exist [60,62,63]. The current TX–45 + DPH mixed system shows that the values for mixed and pure amphiphiles are less than 0.333 (except pure TX–45) in the absence and presence of IL, which confirms spherical micelle formation.

3.6. Thermodynamics of the Drug/Surfactant Mixture

The capacity of an amphiphile to form a micelle is reflected in the change in standard Gibbs free energy ( Δ G m o ). For single and mixed surfactant systems, the standard Gibbs free energy of micellization is given by the following relationship [64].
Δ G m o = R T l n X CMC
The mixing of one amphiphile (TX–45) with the other (DPH) makes the mixed micelle more facile, confirmed by the obtained negative values of Δ G m o for the current mixed system. On increasing the TX–45 concentration, Δ G m o values become more negative, specifying the more spontaneous process of mixing. The Δ G m o values also are calculated by the Maeda model [65]. According to Maeda, the interaction parameter values (β) obtained by the Rubingh model incorporated alone into the head group-head group interaction; nonetheless, chain–chain interaction occurs, predominantly when the chains of two or more surfactants are different in length. The CMC of TX–45 in mixture is habitually considerably inferior to the CMC of participated components. This was credited to the decrease in the head group’s repulsion due to the attendance of non-ionic surfactants amid the head group of ionic surfactants. Maeda describes the chain–chain interaction with a parameter B1, which also describes the stability of the mixture. The free energy of micellization in the presence of ionic surfactant can be obtained by the relation:
Δ G M = R T ( B 0 + B 1 X 1 + B 2 X 1 2 )
where, B0 = ln CMC2, B1 + B2 = ln (C1/C2) and B2 = −β
The estimated values of all parameters are given in Table 3.
The changes in free energy values calculated by Equation (17) and by Maeda’s Equation (18) agree well. The attained value of Δ G M of current was found to be negative, considering the spontaneity of the mixed micellation process (Table 3). This result demonstrates the negligible contribution of a degree of counterion binding in the mixed micelle. The drug and TX–45 have different chain lengths; there should be chain-chain interaction stabilizing the micelle. The values of B1 are higher and negative in the current work, which indicates that the chain–chain interaction plays a foremost character in the stabilization of a micelle. However, the values of B1 appear to be a function of the composition of a system as well as the head groups.
The values Δ G m o can be transformed into the standard free energy of adsorption, Δ G a d d o , at the interface from the relationship (19)
Δ G a d d o = Δ G m o ( π CMC Γ m a x )
here, π CMC is the surface pressure at CMC. The obtained values of Δ G a d d o are negative as with Δ G m o but the magnitude is greater in Δ G a d d o , revealing that the adsorption is more spontaneous and a primary process while micellization phenomena is a secondary process (Table 3).
The excess free energy of micellization, Δ G e x , can be computed with help of activity coefficients values [66,67,68,69,70]:
Δ G e x = R T [ X 1 l n f 1 + ( 1 X 1 ) l n f 2 ]
The negative values of Δ G e x suggest the creation of steady mixed micelle rather than micelle formation by the single amphiphiles (Table 3). The negative values also correspond to synergism. The extent of synergism after the mixing of two more amphiphiles can also be evaluated by a thermodynamic parameter, G m i n s [71]:
G m i n s = A m i n γ CMC N A
The above energetic parameter can be defined as the work required to create a surface area per mole complemented by the changeover from the bulk phase to the surface phase of the solution. The low values of G m i n s listed in Table 3 suggest a more energetically steady interface is created and extra surface activity is achieved.

3.7. UV-Visible Spectroscopy

DPH absorption spectra were explored in the attendance of TX–45 to examine the interaction. We also studied the effect of IL on the absorption spectra of DPH + TX–45. Only one characteristic peak at 257 nm was obtained for the DPH (Figure 3). The hyperchromic effect (absorption intensity increases) was observed when we added the TX–45 to the solution of the drug. Thus, the creation of a new complex between DPH and TX–45 was confirmed by the hyperchromic result. The quantitative estimation of the binding can be computed by the Hildebrand Equation (22) [72]:
1 A A 0 = 1 K ( A m a x A 0 ) [ S ] n + 1 A m a x A 0
where [S] is the concentration of TX–45, and A, A0 and Amax are the absorbance values of DPH in the presence of TX–45, in the absence of TX–45, and the absorbance owing to the creation of the drug–surfactant complex. A straight line is obtained on plotting 1/(AA0) against 1/[S] (Figure 4) that also indicates the formation of the 1:1 complex. The binding constant is calculated (intercept/slope) from the Bensei-Hildebrand equation. The obtained binding constant values (K) are 2.46 × 102 and 9.38 × 103 M−1 for DPH + TX–45 and DPH + TX–45 + IL, respectively. The DPH shows more binding affinity toward TX–45 in the presence of IL, confirmed by the K values. It is clear from Figure 4 that the linear plot is obtained when n = 1 for DPH + TX–45 mixtures in the presence and absence of IL, indicating 1:1 stoichiometry. The equilibrium processes can be expressed as follows:
DPH + TX 45 K DPH :   TX 45
K = [ DPH : TX 45 ] [ DPH ] [ TX 45 ]
The K values are utilized to compute the values of free energy change ( Δ G K ) by Equation (24)
Δ G K = R T l n K
The values of free energy change are −13.65 and −22.67 kJmol−1 for DPH + TX–45 and DPH + TX–45 + IL, respectively. The values of ∆GK were found to be negative in both cases to confirm the spontaneity of the process of complexation.

4. Conclusions

The interaction between DPH and TX–45 is noteworthy in estimation the part of surfactant in biological progressions. Different theoretical models are applied here to know the physicochemical behavior of a drug in the attendance of amphiphile. The experimentally calculated value of CMC is lower than ideally calculated CMC values and the values of CMC decline along with the rise in the mole fraction of surfactant. The interaction parameter values that are negative confirm the synergistic behavior of drug and surfactant mixtures in the nonattendance and attendance of IL. The surface excess ( Γ m a x ) values for mixtures are less than the values of single amphiphiles. That can be attributed lowest power of the mixed system to pump the molecules to the interface, which leads to a decrease in the surface concentration. The Δ G m o and Δ G a d d o values are negative, indicating that the micelle formation and adsorption of surfactants at the surface is thermodynamically advantageous, while the stability of the mixed micelle is validated by the negative value of Δ G e x . For our current mixtures, the values for mixed and single amphiphiles in the attendance and non-attendance of IL are less than 0.333 (except for pure TX–45), confirming spherical micelle formation. The UV-visible spectra of DPH report an enhancement in the absorbance by the addition of surfactant. Electronic absorption spectra in the attendance of IL suggest that, in the presence of IL, the binding of the drug with the surfactant is more efficient, also confirmed by the K value.

Author Contributions

Conceptualization, N.A. and M.A.R.; methodology, formal analysis and investigation and writing—original draft preparation, N.A.; writing—review and editing, N.A., M.A.R. and A.M.A.; supervision and project administration, N.A.; funding acquisition, N.A. All authors have read and agreed to the published version of the manuscript.

Funding

This Project was funded by the Deanship of Scientific Research (DSR) at King Abdulaziz University, Jeddah, under grant no. (G: 341-130-1441). The authors, therefore, acknowledge with thanks DSR for technical and financial support.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

This Project was funded by the Deanship of Scientific Research (DSR) at King Abdulaziz University, Jeddah, under grant no. (G: 341-130-1441). The authors, therefore, acknowledge with thanks DSR for technical and financial support.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Allen, T.M.; Hansen, C.B.; de Menezes, D.E.L. Pharmacokinetics of long-circulating liposomes. Adv. Drug Deliv. Rev. 1995, 16, 267–284. [Google Scholar] [CrossRef]
  2. Canto, G.S.; Dalmora, S.L.; Oliveira, A.G. Piroxicam encapsulated in liposomes: Characterization and in vivo evaluation of topical anti-inflammatory effect. Drug Dev. Ind. Pharm. 1999, 25, 1235–1239. [Google Scholar] [CrossRef]
  3. Gref, R.; Minamitake, Y.; Peracchia, M.; Trubetskoy, V.; Torchilin, V.; Langer, R. Biodegradable long-circulating polymeric nanospheres. Science 1994, 263, 1600–1603. [Google Scholar] [CrossRef] [Green Version]
  4. Jones, M.-C.; Leroux, J.-C. Polymeric micelles—A new generation of colloidal drug carriers. Eur. J. Pharm. Biopharm. 1999, 48, 101–111. [Google Scholar] [CrossRef]
  5. Torchilin, V.P.; Trubetskoy, V.S.; Whiteman, K.R.; Ferruti, P.; Veronese, F.M.; Caliceti, P. New synthetic amphiphilic polymers for steric protection of liposomes in vivo. J. Pharm. Sci. 1995, 84, 1049–1053. [Google Scholar] [CrossRef] [PubMed]
  6. Lin, J.; Zhang, S.; Chen, T.; Lin, S.; Jin, H. Micelle formation and drug release behavior of polypeptide graft copolymer and its mixture with polypeptide block copolymer. Int. J. Pharm. 2007, 336, 49–57. [Google Scholar] [CrossRef]
  7. Torchilin, V.P. Structure and design of polymeric surfactant-based drug delivery systems. J. Controll. Release 2001, 73, 137–172. [Google Scholar] [CrossRef]
  8. Abildskov, J.; O’Connell, J.P. Molecular thermodynamic modeling and design of microencapsulation systems for drug delivery. J. Chem. Eng. Data 2011, 56, 1229–1237. [Google Scholar] [CrossRef]
  9. Noori, S.; Naqvi, A.Z.; Ansari, W.H.; Kabir-ud-Din. Experimental and theoretical approach to cationic drug-anionic gemini surfactant systems in aqueous medium. Colloids Surf. B Biointerfaces 2014, 115, 71–78. [Google Scholar] [CrossRef]
  10. Carmona-Ribeiro, A.; Vieira, D.; Lincopan, N. Cationic surfactants and lipids as anti-infective agents. Anti Infect. Agents Med. Chem. 2006, 5, 33–51. [Google Scholar] [CrossRef]
  11. Khan, A.M.; Shah, S.S. A UV-visible study of partitioning of pyrene in an anionic surfactant sodium dodecyl sulfate. J. Dispers. Sci. Technol. 2008, 29, 1401–1407. [Google Scholar] [CrossRef]
  12. Taboada, P.; Ruso, J.M.; Garcia, M.; Mosquera, V. Surface properties of some amphiphilic antidepressant drugs. Colloids Surf. A Physicochem. Eng. Asp. 2001, 179, 125–128. [Google Scholar] [CrossRef]
  13. 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. Biophys. Acta (BBA) Biomembr. 2000, 1508, 210–234. [Google Scholar] [CrossRef] [Green Version]
  14. Torchilin, V.P. Targeted polymeric micelles for delivery of poorly soluble drugs. Cell. Mol. Life Sci. 2004, 61, 2549–2559. [Google Scholar] [CrossRef]
  15. Naqvi, A.Z.; Noori, S.; Kabir-ud-Din, K.-D. Mixed micellization of dimeric surfactant-amphiphilic drug systems: Effect of surfactant structure. RSC Adv. 2016, 6, 20324–20336. [Google Scholar] [CrossRef]
  16. Trawiſska, A.; Hallmann, E.; MĿdrzycka, K. The effect of alkyl chain length on synergistic effects in micellization and surface tension reduction in nonionic gemini (S-10) and anionic surfactants mixtures. Colloids Surf. A Physicochem. Eng. Asp. 2016, 506, 114–126. [Google Scholar] [CrossRef]
  17. Malik, N.A. Solubilization and interaction studies of bile salts with surfactants and drugs: A review. Appl. Biochem. Biotechnol. 2016, 179, 179–201. [Google Scholar] [CrossRef]
  18. Turovsky, T.; Portnaya, I.; Kesselman, E.; Ionita-Abutbul, I.; Dan, N.; Danino, D. Effect of temperature and loading on the structure of β-casein/ibuprofen assemblies. J. Colloid Interface Sci. 2015, 449, 514–521. [Google Scholar] [CrossRef]
  19. Xiao, H.; Wang, L. Effects of X-shaped reduction-sensitive amphiphilic block copolymer on drug delivery. Int. J. Nanomed. 2015, 10, 5309. [Google Scholar] [CrossRef] [Green Version]
  20. Mahajan, S.; Sharma, R.; Mahajan, R.K. Interactions of new 1,8-diazabicyclo[5.4.0]undec-7-ene (DBU) based surface active ionic liquids with amitriptyline hydrochloride: Micellization and interfacial studies. Colloids Surf. A Physicochem. Eng. Asp. 2013, 424, 96–104. [Google Scholar] [CrossRef]
  21. 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]
  22. Saha, U.; Banerjee, A.; Das, B. Drug-surfactant comicellization: Propranolol hydrochloride-surface active ionic liquid systems in aqueous medium. J. Mol. Liq. 2020, 309, 113164. [Google Scholar] [CrossRef]
  23. Ghasemi, A.; Bagheri, A. Effects of alkyl chain length on synergetic interaction and micelle formation between a homologous series of n-alkyltrimethylammonium bromides and amphiphilic drug propranolol hydrochloride. J. Mol. Liq. 2020, 298, 111948. [Google Scholar] [CrossRef]
  24. 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]
  25. 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. Thermodynam. 2018, 121, 199–210. [Google Scholar] [CrossRef]
  26. 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]
  27. 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]
  28. 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]
  29. 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]
  30. 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]
  31. 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]
  32. 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. [Google Scholar] [CrossRef]
  33. 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]
  34. Raj, S. Simultaneous determination of pseudoephedrine hydrochloride and diphenhydramine hydrochloride in cough syrup by gas chromatography (GC). Talanta 1998, 46, 221–225. [Google Scholar] [CrossRef]
  35. Raphael, G.D.; Angello, J.T.; Wu, M.-M.; Druce, H.M. Efficacy of diphenhydramine vs desloratadine and placebo in patients with moderate-to-severe seasonal allergic rhinitis. Ann. Allergy Asthma Immunol. 2006, 96, 606–614. [Google Scholar] [CrossRef]
  36. Taboada, P.; Attwood, D.; Ruso, J.M.; Sarmiento, F.; Mosquera, V. Self-association of amphiphilic penicillins in aqueous electrolyte solution: A light-scattering and NMR study. Langmuir 1999, 15, 2022–2028. [Google Scholar] [CrossRef]
  37. Welton, T. Room-temperature ionic liquids—Solvents for synthesis and catalysis. Chem. Rev. 1999, 99, 2071–2084. [Google Scholar] [CrossRef] [PubMed]
  38. Visser, A.E.; Rogers, R.D. Room-temperature ionic liquids: New solvents for f -element separations and associated solution chemistry. J. Solid State Chem. 2003, 171, 109–113. [Google Scholar] [CrossRef]
  39. Anderson, J.L.; Armstrong, D.W. Immobilized ionic liquids as high-selectivity/high-temperature/high-stability gas chromatography stationary phases. Anal. Chem. 2005, 77, 6453–6462. [Google Scholar] [CrossRef] [Green Version]
  40. Mo, J.; Xu, L.; Xiao, J. Ionic liquid-promoted, highly regioselective heck arylation of electron-rich olefins by Aryl Halides. J. Am. Chem. Soc. 2005, 127, 751–760. [Google Scholar] [CrossRef]
  41. Boxall, D.L.; Osteryoung, R.A. Switching potentials and conductivity of polypyrrole films prepared in the ionic liquid 1-butyl-3-methylimidazolium hexafluorophosphate. J. Electrochem. Soc. 2004, 151, E41. [Google Scholar] [CrossRef]
  42. Scurto, A.M.; Aki, S.N.V.K.; Brennecke, J.F. Carbon dioxide induced separation of ionic liquids and water. Chem. Commun. 2003, 572–573. [Google Scholar] [CrossRef] [PubMed]
  43. Kubisa, P. Ionic liquids in the synthesis and modification of polymers. J. Polym. Sci. Part A Polym. Chem. 2005, 43, 4675–4683. [Google Scholar] [CrossRef]
  44. Chiappe, C.; Pieraccini, D.; Saullo, P. Nucleophilic displacement reactions in ionic liquids: Substrate and solvent effect in the reaction of NaN 3 and KCN with alkyl halides and tosylates. J. Org. Chem. 2003, 68, 6710–6715. [Google Scholar] [CrossRef] [PubMed]
  45. Nakashima, T.; Kimizuka, N. Interfacial synthesis of hollow TiO2 microspheres in ionic liquids. J. Am. Chem. Soc. 2003, 125, 6386–6387. [Google Scholar] [CrossRef]
  46. Grätzel, M. Conversion of sunlight to electric power by nanocrystalline dye-sensitized solar cells. J. Photochem. Photobiol. A Chem. 2004, 164, 3–14. [Google Scholar] [CrossRef]
  47. Mahajan, S.; Sharma, R.; Mahajan, R.K. An Investigation of drug binding ability of a surface active ionic liquid: Micellization, electrochemical, and spectroscopic studies. Langmuir 2012, 28, 17238–17246. [Google Scholar] [CrossRef] [PubMed]
  48. Srinivasa Rao, K.; Singh, T.; Trivedi, T.J.; Kumar, A. Aggregation behavior of amino acid ionic liquid surfactants in aqueous media. J. Phys. Chem. B 2011, 115, 13847–13853. [Google Scholar] [CrossRef] [PubMed]
  49. Blesic, M.; Marques, M.H.; Plechkova, N.V.; Seddon, K.R.; Rebelo, L.P.N.; Lopes, A. Self-aggregation of ionic liquids: Micelle formation in aqueous solution. Green Chem. 2007, 9, 481. [Google Scholar] [CrossRef]
  50. Jiao, J.; Han, B.; Lin, M.; Cheng, N.; Yu, L.; Liu, M. Salt-free catanionic surface active ionic liquids 1-alkyl-3-methylimidazolium alkylsulfate: Aggregation behavior in aqueous solution. J. Colloid Interface Sci. 2013, 412, 24–30. [Google Scholar] [CrossRef] [PubMed]
  51. Pal, A.; Pillania, A. Modulations in surface and aggregation properties of non-ionic surfactant Triton X-45 on addition of ionic liquids in aqueous media. J. Mol. Liq. 2017, 233, 243–250. [Google Scholar] [CrossRef]
  52. Attwood, D. Micelle formation by some antihistamines in aqueous solution. J. Pharm. Pharmacol. 1972, 24, 751–752. [Google Scholar] [CrossRef]
  53. Rajput, S.M.; Kumar, S.; Aswal, V.K.; El Seoud, O.A.; Malek, N.I.; Kailasa, S.K. Drug-induced micelle-to-vesicle transition of a cationic gemini surfactant: Potential applications in drug delivery. ChemPhysChem 2018, 19, 865–872. [Google Scholar] [CrossRef]
  54. Pal, A.; Pillania, A. The effect of hydrophilic ionic liquid 1-butyl-2,3-dimethylimidazolium bromide on the aggregation behavior of tetradecyltrimethylammonium bromide in aqueous media. J. Mol. Liq. 2015, 209, 6–13. [Google Scholar] [CrossRef]
  55. Behera, K.; Pandey, M.D.; Porel, M.; Pandey, S. Unique role of hydrophilic ionic liquid in modifying properties of aqueous Triton X-100. J. Chem. Phys. 2007, 127, 184501. [Google Scholar] [CrossRef]
  56. Sharma, R.; Mahajan, R.K. An investigation of binding ability of ionic surfactants with trifluoperazine dihydrochloride: Insights from surface tension, electronic absorption and fluorescence measurements. RSC Adv. 2012, 2, 9571. [Google Scholar] [CrossRef]
  57. Motomura, K.; Yamanaka, M.; Aratono, M. Thermodynamic consideration of the mixed micelle of surfactants. Colloid Polym. Sci. 1984, 262, 948–955. [Google Scholar] [CrossRef]
  58. More, U.; Vaid, Z.; Bhamoria, P.; Kumar, A.; Malek, N.I. Effect of [C n mim][Br] based ionic liquids on the aggregation behavior of tetradecyltrimethylammonium bromide in aqueous medium. J. Solut. Chem. 2015, 44, 850–874. [Google Scholar] [CrossRef]
  59. Javadian, S.; Ruhi, V.; Asadzadeh Shahir, A.; Heydari, A.; Akbari, J. Imidazolium-based ionic liquids as modulators of physicochemical properties and nanostructures of CTAB in aqueous solution: The effect of alkyl chain length, hydrogen bonding capacity, and anion type. Ind. Eng. Chem. Res. 2013, 52, 15838–15846. [Google Scholar] [CrossRef]
  60. Israelachvili, J.N.; Mitchell, D.J.; Ninham, B.W. Theory of self-assembly of hydrocarbon amphiphiles into micelles and bilayers. J. Chem. Soc. Faraday Trans. 1976, 72, 1525. [Google Scholar] [CrossRef]
  61. Tanford, C. The Hydrophobic Effect: Formation of Micelles and Biological Membranes; John Wiley & Sons.: Hoboken, NJ, USA, 1980; ISBN 0471048933. [Google Scholar]
  62. Myers, D. Surfactant Science and Technology; John Wiley & Sons: Hoboken, NJ, USA, 2005; ISBN 0471746061. [Google Scholar]
  63. Israelachvili, J.N. Intermolecular and Surface Forces; Academic Press: Cambridge, MA, USA, 2011; ISBN 0080923631. [Google Scholar]
  64. Tsubone, K.; Arakawa, Y.; Rosen, M.J. Structural effects on surface and micellar properties of alkanediyl-α,ω-bis(sodium N-acyl-β-alaninate) gemini surfactants. J. Colloid Interface Sci. 2003, 262, 516–524. [Google Scholar] [CrossRef]
  65. Maeda, H. A simple thermodynamic analysis of the stability of ionic/nonionic mixed micelles. J. Colloid Interface Sci. 1995, 172, 98–105. [Google Scholar] [CrossRef]
  66. Khan, F.; Sheikh, M.S.; Rub, M.A.; Azum, N.; Asiri, A.M. Antidepressant drug amitriptyline hydrochloride (AMT) interaction with anionic surfactant sodium dodecyl sulfate in aqueous/brine/urea solutions at different temperatures. J. Mol. Liq. 2016, 222, 1020–1030. [Google Scholar] [CrossRef]
  67. Azum, N.; Rub, M.A.; Asiri, A.M. Self-association and micro-environmental properties of sodium salt of ibuprofen with BRIJ-56 under the influence of aqueous/urea solution. J. Dispers. Sci. Technol. 2017, 38, 96–104. [Google Scholar] [CrossRef]
  68. Azum, N.; Rub, M.A.; Asiri, A.M. Interaction of triblock-copolymer with cationic gemini and conventional surfactants: A physicochemical study. J. Dispers. Sci. Technol. 2017, 38, 1785–1791. [Google Scholar] [CrossRef]
  69. Azum, N.; Rub, M.A.; Asiri, A.M. Bile salt–bile salt interaction in mixed monolayer and mixed micelle formation. J. Chem. Thermodynam. 2019, 128, 406–414. [Google Scholar] [CrossRef]
  70. Azum, N.; Rub, M.A.; Asiri, A.M.; Kashmery, H.A. Synergistic effect of an antipsychotic drug chlorpromazine hydrochloride with pluronic triblock copolymer: A physicochemical study. J. Mol. Liq. 2018, 260, 159–165. [Google Scholar] [CrossRef]
  71. 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]
  72. 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. Structure of diphenhydramine hydrochloride (DPH), 1-hexyl-3-methylimidazolium chloride (IL), and triton X–45 (TX–45).
Figure 1. Structure of diphenhydramine hydrochloride (DPH), 1-hexyl-3-methylimidazolium chloride (IL), and triton X–45 (TX–45).
Polymers 13 02756 g001
Figure 2. Tensiometric plots of surface tension (γ) vs. log amphiphile concentration [C] at 298.15.
Figure 2. Tensiometric plots of surface tension (γ) vs. log amphiphile concentration [C] at 298.15.
Polymers 13 02756 g002
Figure 3. Absorbance spectra of DPH along with rising [TX–45].
Figure 3. Absorbance spectra of DPH along with rising [TX–45].
Polymers 13 02756 g003
Figure 4. Benesi-Hildebrand graph by utilizing the change in absorption spectra of DPH.
Figure 4. Benesi-Hildebrand graph by utilizing the change in absorption spectra of DPH.
Polymers 13 02756 g004
Table 1. Micellar parameters of TX–45 and DPH mixtures in the attendance and non-attendance of 25 mM IL at T = 298.15 K.
Table 1. Micellar parameters of TX–45 and DPH mixtures in the attendance and non-attendance of 25 mM IL at T = 298.15 K.
α1CMC (mM)CMC* (mM)X1 X 1 ideal βf1f2 ln ( CMC 1 CMC 2 )
TX–45 + DPH
0.0112 −4.89
0.10.698.390.590.9312.450.1280.012
0.30.492.940.650.9811.410.2490.008
0.50.391.790.680.9911.360.3170.005
0.70.231.280.690.9914.090.2580.001
0.90.210.990.710.9914.720.2950.001
1.00.90
TX–45 + DPH + 25 mM IL
093.50 −7.71
0.50.050.080.820.999.660.7360.001
1.00.04
Table 2. Surface parameters of TX–45 and DPH mixtures in the attendance and non-atendance of 25 mM IL at T = 298.15 K.
Table 2. Surface parameters of TX–45 and DPH mixtures in the attendance and non-atendance of 25 mM IL at T = 298.15 K.
α1106 Γmax (mol m−2)Amin (nm2)pC20γCMC (mN m−1)πCMC (mN m−1)
TX–45 + DPH
0.01.8090.9170.94049.7620.24
0.11.0561.5713.32448.0721.93
0.31.3181.2593.36742.8127.19
0.51.2791.2973.96136.9433.06
0.70.8791.8874.42834.7135.29
0.90.9551.7384.59132.6937.31
1.03.8270.4334.91027.4942.51
TX–45 + DPH + 25 mM IL
01.2561.3210.95650.2719.73
0.51.3151.2625.14332.2337.77
1.03.2490.5105.51128.3141.69
Table 3. Packing and energetic constraints of TX–45 & DPH mixtures in the attendance and non-attendance of 25 mM IL at T = 298.15 K.
Table 3. Packing and energetic constraints of TX–45 & DPH mixtures in the attendance and non-attendance of 25 mM IL at T = 298.15 K.
α1PB1−ΔGM (kJ mol−1) Δ G m o   ( kJ   mol 1 ) −ΔGads (kJ mol−1)Gmin (kJ mol−1)Gex
TX–45 + DPH
0.00.23 15.2026.3827.49
0.10.137.6329.9227.9948.7545.497.44
0.30.176.5929.5828.8549.4732.466.42
0.50.166.5429.6329.4055.2328.856.10
0.70.119.2731.0930.7170.8239.447.46
0.90.129.9031.3630.9369.9934.227.47
1.00.49 27.3238.437.18
TX–45 + DPH + 25 mM IL
00.16 15.8231.5240.00
0.50.171.9535.0234.4463.1524.493.50
1.00.41 34.9247.758.71
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.; Asiri, A.M. Mixed Micellization and Spectroscopic Studies of Anti-Allergic Drug and Non-Ionic Surfactant in the Presence of Ionic Liquid. Polymers 2021, 13, 2756. https://0-doi-org.brum.beds.ac.uk/10.3390/polym13162756

AMA Style

Azum N, Rub MA, Asiri AM. Mixed Micellization and Spectroscopic Studies of Anti-Allergic Drug and Non-Ionic Surfactant in the Presence of Ionic Liquid. Polymers. 2021; 13(16):2756. https://0-doi-org.brum.beds.ac.uk/10.3390/polym13162756

Chicago/Turabian Style

Azum, Naved, Malik Abdul Rub, and Abdullah M. Asiri. 2021. "Mixed Micellization and Spectroscopic Studies of Anti-Allergic Drug and Non-Ionic Surfactant in the Presence of Ionic Liquid" Polymers 13, no. 16: 2756. https://0-doi-org.brum.beds.ac.uk/10.3390/polym13162756

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