Aggregation of sodium dodecylbenzene sulfonate: Weak molecular interactions modulated by imidazolium cation of short alkyl chain length

https://doi.org/10.1016/j.colsurfa.2020.124435Get rights and content

Highlights

  • Aggregation of SDBS is studied in imidazolium IL aqueous solution.

  • SDBS aggregation process involves two stages.

  • The SDBS aggregate stability is determined by the IL structure.

  • The IL changes the size of the SDBS aggregates.

Abstract

Ionic liquids (ILs) can modify cooperative process in aqueous solutions to a large extent, including anionic surfactant aggregation. Here, the micellization of sodium dodecylbenzene sulfonate (SDBS) was evaluated in low concentrations of 1-alkyl-3-methylimidazolium chloride (CnmimCl, n = 0, 2, and 4) aqueous solutions through fluorescence spectroscopy, isothermal titration calorimetry, dynamic light scattering, and conductometry. The thermodynamic stability of SDBS aggregates strongly depended on the IL structure and concentration, following the order C4mim+ > C0mim+ ≈ C2mim+. At 1.0 mmol L−1 of the ILs, the increase of the hydrophobicity of the imidazolium cation decreased the enthalpic favorableness, changing ΔHmico from −3.75 ± 0.07 kJ mol−1, for C0mim+, to −2.69 ± 0.01 kJ mol−1, for C4mim+. On the other hand, the entropic feasibility showed an opposite trend, i.e., the higher hydrophobicity of C4mim+ overcame the kosmotropic effect of IL cations in the bulks. We suggested that the imidazolium cations interact with the SDBS monomers on the micellar surface, mainly through hydrophobic, π-π, and electrostatic interactions for C4mim+ and C2mim+, and through electrostatic interactions and hydrogen bonds for C0mim+.

Introduction

The self-assembly of chemical compounds in aqueous solution is a fundamental physicochemical process, both in theorical and practical aspects. Among several molecules that can aggregate through self-assembly, surfactants have gained considerable attention because of the wide range of their potential applications in fields such as pharmaceutical formulations [1], environmental remediation [2], foam generation [3], personal care products [4], cytotoxicity of metal nanoparticles [5], nanomaterial synthesis [6], reactivity in micellar solutions [7], and drug delivery [8]. Surfactants can form aggregates above a narrow concentration range, known as the critical micellar concentration (cmc), at fixed temperature and pressure. [[9], [10], [11]] The cmc depend on several properties, specially the molecular properties of the solvent; [[12], [13], [14]] therefore, it can be modulated by the addition of cosolutes or cosolvents.

The effect of cosolutes and cosolvents, such as methanol, propanol, acetamide, urea, NaBr, and ethylene glycol [[15], [16], [17]] on the surfactant micellization has been widely studied. In the last fifteen years, ionic liquids (ILs), electrolytes with bulky organic cation and inorganic (or organic) anions [18], have attracted attention as potential cosolutes [[19], [20], [21]] because of their ability to simultaneously modify, electrostatic and hydrophobic interactions in aqueous systems, thereby modulating different cooperative processes [[22], [23], [24], [25]]. The ILs containing the imidazolium cation are widely studied due to their negligible vapor pressure, high ionic conductivity, and because their polarity can be modulated by the number of carbons in the lateral chain on the imidazolium cation ring [26].

Moreover, the imidazolium ILs have been extensively used to investigate the thermodynamics of micellization of model ionic surfactants, such as sodium dodecyl sulfate (SDS). In 2009, Behera et al. [27,28] studied the effects of hydrophilic 1-butyl-3-methylimidazolium tetrafluoroborate (C4mimBF4) and hydrophobic 1-butyl-3-methylimidazolium hexafluorophosphate (C4mimPF6) ILs on SDS micellization, by conducting fluorescence spectroscopy and conductivity. Both these ILs showed two distinct effects: for C4mimBF4, the cmc decreases when IL concentration is increased up to 2 % wt, but increases when IL concentration is raised from 2 to 30 % wt; for C4mimPF6, the cmc of SDS decreases when IL concentration is increased up to 0.1 % wt and, remains constant thereafter even when the IL concentration is increased up to 30 % wt. These results confirm that ILs act as simple electrolytes at lower concentrations and are partitioned from the bulk to the SDS micelles core at higher concentrations, promoting the micelle growth. In 2009, Smirnova et al. [29] studied the SDS micellization in 1-alkyl-3-methylimidazolium salts + water mixtures through conductivity and calorimetry measurements. The increase in concentration and carbon number in the lateral chain of the imidazolium cation promoted a sharper decrease in the cmc and micellization enthalpy change values (ΔHmico) when compared with those of the simple electrolyte NaCl. This decrease was attributed to the changes in the electrostatic and hydrophobic interactions in the system caused by IL cations. In 2013, Pal et al. [30] examined the behavior of SDS solutions in the presence of 3-methyl-5-methylimidazolium hexafluorophosphate through conductivity and fluorescence spectroscopy. They showed that the IL causes an increase in the cmc values of the SDS due to the presence of solvophobic interactions between the IL and the surfactant chains, facilitating the formation of mixed micelles. In 2015, Ferreira et al. [24] investigated the effect of 1-butyl-3-methylimidazolium halides on the SDS micellization by performing calorimetry, fluorescence, and conductivity and showed that the imidazolium cation introduces its small carbonic chain into the hydrophobic part of the micelle.

Micellar systems based on more complex anionic surfactants, such as sodium dodecylbenzene sulfonate (SDBS), have also been studied in the presence of imidazolium ILs. In 2010, Rai et al. [31] studied the size of SDBS micelles in the presence of the C4mimPF6 IL using dynamic light scattering (DLS). They showed that for a SDBS concentration of 300 mmol L−1 and an IL concentration of approximately 100 mmol L−1 (3:1 stoichiometric ratio), the diameter of the micelles increases drastically by approximately 20 nm due to the incorporation of the imidazolium cations in the micellar aggregate. In 2015, Chabba et al. [32] studied the formation of SDBS micelles in aqueous solutions in the presence of CnmimCl (n = 8, 10 or 12) using tensiometry, fluorescence spectroscopy, and dynamic light scattering. Their results suggested that the formation of mixed micelles and vesicles occur due to the balance between cation-π, hydrophobic, strong electrostatic, and H-bonding interactions.

Despite the extensive information about the effect of imidazole-based ILs on the micellization of model ionic surfactants, systematic thermodynamic studies involving the effect of the hydrophobic/hydrophilic balance of these ILs on the micellization of atypical anionic surfactants, particularly in low concentrations of the IL, are scarce.

Here, we study the effect of low concentrations of 1-alkyl-3-methylimidazolium chlorides (Fig. 1S) on the micellization thermodynamics of SDBS in aqueous solution using isothermal titration calorimetry, fluorescence spectroscopy, electrical conductivity, and dynamic light scattering. The thermodynamic parameters of micellization (ΔGmico, ΔHmico, and TΔSmico) have been evaluated for ILs with distinct concentrations and hydrophobicity.

Section snippets

Materials

Sodium dodecylbenzene sulfonate (technical degree), 1-methylimidazolium chloride (95.0 %), 1-ethyl-3-methylimidazolium chloride (≥ 95.0 %), and 1-butyl-3-methylimidazolium chloride (≥ 98.0 %) were obtained from Sigma-Aldrich (USA). Sodium chloride (99.0 %) was purchased from Isofar (Brazil). All reagents were used without further purification. Deionized water (18 MΩ cm at 298.2 K) from Milli-QII (Millipore, USA) was used to prepare all solutions.

Fluorescence measurements

Fluorescence emission spectra of pyrene in SDBS

Fluorescence spectroscopy

To evaluate the effect of ILs on the self-assembly of SDBS in aqueous solution, it is necessary to first determine its aggregation properties in pure water. The SDBS cmc values reported in the literature vary in a large range, depending on the technique used to access this property and the intrinsic nature of the sample, and considering that commercially available SDBS can be a mixture of isomers. [33] A method commonly used to measure the cmc of surfactants in aqueous solutions is the

Conclusions

In this work, the thermodynamics of SDBS micellization was investigated in aqueous solutions with low concentrations of 1-alkyl-3-methylimidazolium chlorides (CnmimCl, n = 0, 2, or 4). Fluorometric and calorimetric measurements showed that the aggregation process of the surfactant occurred in two stages: first, a transition of monomers to dimers, trimmers or small aggregates, followed by the formation of micelles.

The feasibility of the micellization process strongly depends on the cation

Credit author statement

Álvaro Javier Patiño Agudelo, Guilherme Max Dias Ferreira and Gabriel Max Dias Ferreira carried out the ITC, fluorescence, conductometric, and UV–vis experiments, performed the analysis and interpretation of data, and contributed to the preparation and writing of this article; Yara Luiza Coelho and Eliara Acipreste Hudson performed the fluorescence and conductometric experiments, as well as the interpretation of the obtained parameters; Ana Clarissa S. Pires participated in the writing and

Declaration of Competing Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

Acknowledgments

We are thankful for financial support from Fundação de Apoio à Pesquisa de Minas Gerais (FAPEMIG), Financiadora de Estudos e Projetos (FINEP), Instituto Nacional de Ciências e Tecnologias Analíticas Avançadas (INCTAA), and Conselho Nacional de Desenvolvimentos Científico e Tecnológico (CNPq). The authors are grateful to Dr. Alvaro Vianna Novaes de Carvalho Teixeira of Federal University of Viçosa for DLS measurements. This study was financed in part by the Coordenação de Aperfeiçoamento de

References (63)

  • K. Behera et al.

    Modulating properties of aqueous sodium dodecyl sulfate by adding hydrophobic ionic liquid

    J. Colloid Interface Sci.

    (2007)
  • N.A. Smirnova et al.

    Self-assembly in aqueous solutions of imidazolium ionic liquids and their mixtures with an anionic surfactant

    J. Colloid Interface Sci.

    (2009)
  • A. Pal et al.

    Ionic liquid induced alterations in the physicochemical properties of aqueous solutions of sodium dodecylsulfate (SDS)

    Colloids Surf. A

    (2013)
  • S. Chabba et al.

    Interfacial and aggregation behavior of aqueous mixtures of imidazolium based surface active ionic liquids and anionic surfactant sodium dodecylbenzenesulfonate

    Colloids Surf. A

    (2015)
  • M. Aoudia et al.

    Excimer-monomer emission in alkylbenzenesulfonate dispersions: effect of the surfactant structure on aggregation

    J. Colloid Interface Sci.

    (1991)
  • A. Graciaa et al.

    HLB, CMC, and phase behavior as related to hydrophobe branching

    J. Colloid Interface Sci.

    (1982)
  • F. Comelles et al.

    Micellization of sodium laurylethoxysulfate (SLES) and short chain imidazolium ionic liquids in aqueous solution

    J. Colloid Interface Sci.

    (2014)
  • J. Jiao et al.

    Electrolyte effect on the aggregation behavior of 1-butyl-3-methylimidazolium dodecylsulfate in aqueous solution

    J. Colloid Interface Sci.

    (2013)
  • S. Chauhan et al.

    Effect of temperature and additives on the critical micelle concentration and thermodynamics of micelle formation of sodium dodecyl benzene sulfonate and dodecyltrimethylammonium bromide in aqueous solution: a conductometric study

    J. Chem. Thermodyn.

    (2014)
  • P.D. Galgano et al.

    Micellar properties of surface active ionic liquids: a comparison of 1-hexadecyl-3-methylimidazolium chloride with structurally related cationic surfactants

    J. Colloid Interface Sci.

    (2010)
  • H. Zhang et al.

    Study on the dimer-monomer equilibrium of a fluorescent dye and its application in nucleic acids determination

    Anal. Chim. Acta

    (1998)
  • W. Loh et al.

    Use of isothermal titration calorimetry to study surfactant aggregation in colloidal systems

    Biochim. Biophys. Acta, Gen. Subj

    (2016)
  • J. Aguiar et al.

    On the determination of the critical micelle concentration by the pyrene 1:3 ratio method

    J. Colloid Interface Sci.

    (2003)
  • M. Pal et al.

    Properties of aqueous micellar solutions in the presence of ionic liquid

    Colloids Surf. A

    (2016)
  • K.D. Collins

    Ions from the Hofmeister series and osmolytes: effects on proteins in solution and in the crystallization process

    Methods

    (2004)
  • H. Zhao et al.

    Hofmeister series of ionic liquids: kosmotropic effect of ionic liquids on the enzymatic hydrolysis of enantiomeric phenylalanine methyl ester

    Tetrahedron Asymmetry

    (2006)
  • M. Pisárcik et al.

    Spherical dodecyltrimethylammonium bromide micelles in the limit region of transition to rod-like micelles. A light scattering study

    Colloids Surf. A

    (1996)
  • C.T.S. Turk et al.

    Formulation and optimization of nonionic surfactants emulsified nimesulide-loaded PLGA-based nanoparticles by design of experiments

    AAPS PharmSciTech

    (2014)
  • M.K. Memon et al.

    Study of blended surfactants to generate stable foam in presence of crude oil for gas mobility control

    J. Pet. Explor. Prod. Technol.

    (2017)
  • B.A. Wilson et al.

    Effects of Three pharmaceutical and personal care products on natural freshwater algal assemblages

    Environ. Sci. Technol.

    (2003)
  • P. Kaur et al.

    Surfactant-based drug delivery systems for treating drug-resistant lung cancer

    Drug Deliv.

    (2016)
  • Cited by (6)

    • Thermodynamics of anion binding to zwitterionic sulfobetaine micelles

      2022, Journal of Colloid and Interface Science
      Citation Excerpt :

      Fluorescence emission spectra (350–600 nm) of pyrene (1.2 μmol L−1) in aqueous solution were acquired with excitation at 335 nm using slits of 2.5 nm for excitation and emission. The ratio of pyrene monomer emission intensities I1 at 373 nm and I3 at 384 nm and the ratio of the excimer emission intensity at 473 nm to that of the monomer at 373 nm were analyzed as a function of log ([SB3-14] / mol L−1) to estimate the CMC [21-23]. The values of 0.27 mmol L−1 estimated from the I1/I3 ratios and 0.28 mmol L−1 from the pyrene excimer/monomer fluorescence ratios are in good agreement with the reported CMC of 0.25 mmol L−1 for SB3-14 at 298.2 K, in water, determined from surface tension measurements [24].

    • Effect of selected monovalent salts on surfactant stabilized foams

      2021, Advances in Colloid and Interface Science
      Citation Excerpt :

      Salts can also impact surfactants foaming performance by affecting the micellar structure of the surfactant. Sood and Aggarwal [99], Cheng and Gulari [100] and more recently, Agudelo et al. [101] reported that screening of electrostatic repulsions between the sodium dodecylbenzene sulfonate (SDBS) head groups in salt solutions up to 0.5 M NaCl leads to an increase in the aggregation number for SDBS molecules and therefore a closer packing of the adsorbed molecules. In another study, Cheng and Gulari [100] measured the SDBS diffusion coefficient at different NaCl concentrations above the CMC of SDBS and showed that the presence of NaCl strongly affects the diffusivity only when the SDBS concentration is above its CMC.

    • Enhanced adsorption/extraction of five typical polycyclic aromatic hydrocarbons from meat samples using magnetic effervescent tablets composed of dicationic ionic liquids and NiFe<inf>2</inf>O<inf>4</inf> nanoparticles

      2020, Journal of Molecular Liquids
      Citation Excerpt :

      They are excellent candidates for extraction of organic chemicals due to several important properties [10]. Imidazolium-based ILs have been used as surface modifiers of polymers or functional monomers because they provide electrostatic, ion-exchange, and π-π interactions [11–13]. Especially, imidazolium-based dicationic ionic liquids (DILs), such as [Cn(MIM)2]Br2, include two imidazolium-ring cations and two anions, and thus they possess two imidazolium-ring active centers, which means they have distinct physicochemical properties that facilitate higher adsorption of target molecules than conventional monocationic ILs [14].

    View full text