Historical perspective
Rheology of mixed solutions of sulfonated methyl esters and betaine in relation to the growth of giant micelles and shampoo applications

https://doi.org/10.1016/j.cis.2019.102062Get rights and content

Highlights

  • The mixing of sulfonated methyl esters and cocamidopropyl betaine is synergistic.

  • The addition of salt and cocamide monoethanolamine broadens the peak in viscosity.

  • Alcohols are either thinning or thickening agents depending on salt concentration.

  • The reptation-reaction model for wormlike micelles is applied to interpret the data.

  • Augmented Maxwell model is applied to systems of nonstandard rheological behavior.

Abstract

This is a review article on the rheological properties of mixed solutions of sulfonated methyl esters (SME) and cocamidopropyl betaine (CAPB), which are related to the synergistic growth of giant micelles. Effects of additives, such as fatty alcohols, cocamide monoethanolamine (CMEA) and salt, which are expected to boost the growth of wormlike micelles, are studied. We report and systematize the most significant observed effects with an emphasis on the interpretation at molecular level and understanding the rheological behavior of these systems. The experiments show that the mixing of SME and CAPB produces a significant rise of viscosity, which is greater than in the mixed solutions of sodium dodecyl sulfate and CAPB. The addition of fatty alcohols, CMEA and cationic polymer, leads to broadening of the synergistic peak in viscosity without any pronounced effect on its height. The addition of NaCl leads to a typical salt curve with high maximum, but in the presence of dodecanol this maximum is much lower. At lower salt concentrations, the fatty alcohol acts as a thickener, whereas at higher salt concentrations – as a thinning agent. Depending on the shape of the frequency dependences of the measured storage and loss moduli, G' and G”, the investigated micellar solutions behave as systems of standard or nonstandard rheological behavior. The systems with standard behavior obey the Maxwell viscoelastic model (at least) up to the crossover point (G' = G”) and can be analyzed in terms of the Cates reptation-reaction model. The systems with nonstandard rheological behavior obey the Maxwell model only in a restricted domain below the crossover frequency; they can be analyzed in the framework of an augmented version of the Maxwell model. The methodology for data analysis and interpretation could be applied to any other viscoelastic micellar system.

Graphical abstract

“Rheology of mixed solutions of sulfonated methyl esters and betaine in relation to the growth of giant micelles and shampoo applications”.

Unlabelled Image
  1. Download : Download high-res image (163KB)
  2. Download : Download full-size image

Introduction

Formation of large micellar aggregates of different morphology is most frequently observed in mixed surfactant solutions, in which the micelles are multicomponent and polydisperse in size [[1], [2], [3], [4], [5], [6], [7]]. Upon variation of solution's composition, high peaks in viscosity have been often observed [[8], [9], [10], [11], [12]]. Such concentration dependencies are of primary importance for various practical applications, e.g. for personal and household care detergency (e.g. shampoos and liquid detergents), because they allow one to control the micelle growth and formulation's viscosity [[13], [14], [15]]. The highest viscosities are observed in the presence of giant wormlike surfactant micelles. The high viscosity is due to the interplay of various processes and interactions that take place in such concentrated and internally structured solutions. First, the high aspect ratio of the long micellar aggregates gives rise to purely hydrodynamic interactions [16,17]. Second, de Gennes [18] identified the main relaxation mechanism for long linear polymers with reptation, which is related to the curvilinear diffusion of linear unbreakable molecules confined by their neighbors. Furthermore, in the case of long micellar aggregates (“living” polymers) Cates and coauthors [[19], [20], [21], [22], [23], [24], [25]] developed statistical theory, which accounts for micelle reversible scission and end-interchange processes. This theory predicts correctly the variations of zero-shear viscosity and other rheological parameters at not too high surfactant and salt concentrations [26]. Recently, Hoffmann and Thurn [27] took into account the energy of sticky contacts between micelles, which include contributions from the van der Waals, electrostatic, hydrophobic and hard-core interactions. Depending on the chemical nature of the component, whose concentration is varied, one could distinguish three types of viscosity peaks:

First, the variation of the mole fractions of the two basic surfactants (anionic and cationic, or anionic and zwitterionic) leads to a maximum in viscosity because of synergistic interactions between the two surfactants that promote growth of large self-assembled aggregates, usually – wormlike micelles, to the left of the maximum and diminishing of their size to the right of the maximum [10,[28], [29], [30], [31]]. At that, the maximal viscosity corresponds to the concentration domain with the longest entangled micelles. The synergism can be due to favorable headgroup interactions (e.g. in a catanionic pair) [32,33], as well as to a mismatch in the surfactant chainlengths [[34], [35], [36]].

Second, in systems containing ionic surfactants the dependence of viscosity on the concentration of added salt (the so called salt curve) often exhibits a high peak [28,[37], [38], [39], [40], [41], [42], [43]]. In this case, the peak could be explained with a transition from wormlike micelles to branched micelles [[44], [45], [46], [47], [48]]. The initial growth of wormlike micelles could be explained with the screening of the electrostatic repulsion between the surfactant headgroups by the electrolyte, whereas the subsequent transition to branched aggregates can be interpreted in terms of surfactant packing parameters and interfacial bending energy [49]. At high salt concentrations, the viscosity could drop because of phase separation due to the salting out of surfactant.

Third, viscosity peaks are observed upon the addition of amphiphilic molecules – cosurfactants, typically fatty acids and alcohols, which are used as thickening agents [[50], [51], [52], [53], [54]]. In this case, the peak can be due again to transformation of the wormlike micelles into branched [52] or ribbonlike and disklike [53,55] aggregates. Alternatively, the peak could be related to the onset of a phase separation of surfactant as a precipitate from droplets and/or crystallites (see Section 3.4). Peaks in viscosity have been observed also in catanionic systems as a function of temperature [56] and in zwitterionic systems as a function of pH [57].

With respect to micelle topology, here we follow the terminology originating from Ref. [45], viz. the entangled wormlike micelle is a linear aggregate with two endcaps and no junctions with other micelles; a branched micelle consist of several connected branches, each of them beginning at a junction and ending with an endcap, and finally, the multiconnected saturated network represents a bicontinuous structure, where all endcaps have been transformed into intramicellar junctions.

In a preceding paper [58], we reported that the viscosity in mixed solutions of sulfonated methyl esters (SME) and cocamidopropyl betaine (CAPB) significantly increases with the rise of the total surfactant concentration. Our goal in the present article is to further extend this study and to examine the effects of surfactant composition, added salt and thickening agents, including the possible appearance of peaks in viscosity that could evidence synergistic growth, micelle shape transformations or phase separation.

The sulfonated methyl esters (SME) are produced from renewable palm-oil based materials [[59], [60], [61]] and have been promoted as alternatives to the petroleum-based surfactants [62]. SMEs exhibit a series of useful properties, such as excellent biodegradability and biocompatibility; excellent stability in hard water; good wetting and cleaning performance, and skin compatibility [60,[63], [64], [65], [66], [67], [68], [69]]. The SME surfactants are produced typically with even alkyl chainlengths, from C12 to C18. Here, they will be denoted CnSME, n = 12, 14, 16, 18. Information for the adsorption and micellar properties of CnSME can be found in Refs. [[70], [71], [72]]. So far, there is a single study on the rheology of micellar SME solutions without cosurfactants (like CAPB) and thickeners, but in the presence of nanoparticles [73].

A complete systematic study on all possible combinations of the basic surfactants, cosurfactants (including fatty alcohols of various chainlengths) and salts exceeds the scope of the present article. Here, we focus on the most significant and interesting effects observed in our experiments with an emphasis on the interpretation of the obtained rheological data on the basis of theoretical models, in order to achieve a better understanding of the underlying molecular processes and phenomena.

In Section 2, we briefly describe the ingredients in the investigated systems and the used experimental methods. In Section 3, we report and systematize data from many rheological experiments in steady shear regime, which demonstrate the existence of strong synergism in the SME + CAPB system with respect to the rise of viscosity; effects of various additives on the synergistic maxima; salt curves with and without added fatty alcohol, and the effect of the concentration of thickeners – fatty alcohols and cocamide monoethanolamine (CMEA). Section 4 is dedicated to rheological experiments in oscillatory regime and to theoretical interpretation of the obtained results for the storage and loss moduli, G′ and G″. Part of the data exhibit standard rheological behavior and are interpreted in terms of the Maxwell viscoelastic model and the reptation-reaction model proposed by Cates [19] and developed in subsequent studies [[20], [21], [22], [23], [24], [25]]. The latter model successfully explains both the Maxwellian behavior of micellar systems and the deviations from it. However, many of the investigated systems exhibit nonstandard rheological behavior. We have demonstrated that the rheological data for such systems can be theoretically described and analyzed in terms of an augmented version of the Maxwell model.

The paper could be useful for a broad audience of researchers, who are interested in synergistic effects in mixed micellar solutions and in their theoretical interpretation.

Section snippets

Materials

In our study, we used two kinds of sulfonated methyl esters (α-sulfo fatty acid methyl ester sulfonates, sodium salts, denoted also α-MES), which are products of the Malaysian Palm Oil Board (MPOB) and KLK OLEO. The first one is myristic sulfonated methyl ester (C14SME) 98%, M = 344.34 g/mol, with critical micelle concentration CMC = 3.68 mM [70].

The second one, which will be denoted C16,18SME, represents a mixture of 85 wt% palmitic (C16SME) and 15 wt% stearic (C18SME) sulfonated methyl esters

Types of flow curves

Fig. 2 illustrates the existence of two types of η-vs.-γ̇ dependencies (flow curves): regular (Fig. 2a and b) and irregular (Fig. 2c and d). The experimental data are obtained with mixed solutions of C14SME + CAPB and C16,18SME + CAPB at two total surfactant concentrations, 8 and 12 wt%, in the presence of various additives: fatty alcohols, CMEA and NaCl. In our experiments, the weight fraction of CAPB, w, in the mixed surfactant solutions has been varied; w is defined as follows:w=WCAPBWCAPB+W

Experimental results for G' and G"; comparison with the Maxwell model

In the case of experiments with rotational rheometer in oscillatory regime, sinusoidal oscillations of the strain are imposed:γt=γasinωtwhere γa is amplitude, t is time and ω is angular frequency. Our experiments were carried out at γa = 0.02, whereas ω was varied. As a rule, the measured stress, σ(t), is phase-shifted, so that it can be expressed in the form:σtγa=Gsinωt+Gcosωtwhere G′ and G″ are the storage and loss moduli, respectively; G′ and G″ are independent of t, but they depend on ω;

Conclusions

The mixing of anionic and zwitterionic surfactants in aqueous solutions is known to promote the synergistic growth of giant micelles, which is detected as a significant rise of viscosity [1,10,29,31,39,42,[50], [51], [52], [53], [54]]. In the present paper we investigated this phenomenon for mixed solutions of sulfonated methyl esters, SME, and CAPB, potential ingredients for personal-care formulations. The effect of SME is compared with those of a standard anionic surfactant, SDS. Moreover,

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

All authors gratefully acknowledge the support from KLK OLEO. VY, KD and PK acknowledge the support from the Operational Programme “Science and Education for Smart Growth”, Bulgaria, grant number BG05M2OP001-1.002-0012. GR acknowledges the financial support received from the program “Young scientists and postdoctoral candidates” of the Bulgarian Ministry of Education and Science, MCD No 577/17.08.2018.

References (89)

  • K.D. Danov et al.

    Analytical modeling of micelle growth. 1. Chain-conformation free energy of binary mixed spherical, wormlike and lamellar micelles

    J Colloid Interface Sci

    (2019)
  • K.D. Danov et al.

    Analytical modeling of micelle growth. 2. Molecular thermodynamics of mixed aggregates and scission energy in wormlike micelles

    J Colloid Interface Sci

    (2019)
  • W. Zhang et al.

    Development of a sulfonic gemini zwitterionic viscoelastic surfactant with high salt tolerance for seawater-based clean fracturing fluid

    Chem Eng Sci

    (2019)
  • M. Pleines et al.

    Molecular factors governing the viscosity peak of giant micelles in the presence of salt and fragrances

    J Colloid Interface Sci

    (2019)
  • S.E. Anachkov et al.

    Disclike vs. cylindrical micelles: generalized model of micelle growth and data interpretation

    J Colloid Interface Sci

    (2014)
  • Z. Mitrinova et al.

    Control of surfactant solution rheology using medium-chain cosurfactants

    Colloids Surf A

    (2018)
  • E.S. Basheva et al.

    Properties of the micelles of sulfonated methyl esters determined from the stepwise thinning of foam films and by rheological measurements

    J Colloid Interface Sci

    (2019)
  • D. Martinez et al.

    Simulation and pre-feasibility analysis of the production process of α-methyl ester sulfonates (α-MES)

    Bioresour Technol

    (2010)
  • K.D. Danov et al.

    Sulfonated methyl esters of fatty acids in aqueous solutions: interfacial and micellar properties

    J Colloid Interface Sci

    (2015)
  • V.I. Ivanova et al.

    Sulfonated methyl esters, linear alkylbenzene sulfonates and their mixed solutions: micellization and effect of Ca2+ ions

    Colloids Surf A

    (2017)
  • Z. Wang et al.

    The structure of alkyl ester sulfonate surfactant micelles: the impact of different valence electrolytes and surfactant structure on micelle growth

    J Colloid Interface Sci

    (2019)
  • M. Luo et al.

    Rheological behavior and microstructure of an anionic surfactant micelle solution with pyroelectric nanoparticle

    Colloids Surf A

    (2012)
  • H. Xu et al.

    Adsorption and self-assembly in methyl ester sulfonate surfactants, their eutectic mixtures and the role of electrolytes

    J Colloid Interface Sci

    (2018)
  • R.D. Stanimirova et al.

    Oil drop deposition on solid surfaces in mixed polymer-surfactant solutions in relation to hair- and skin-care applications

    Colloids Surf A

    (2019)
  • S.S. Tzocheva et al.

    Solubility limits and phase diagrams for fatty alcohols in anionic (SLES) and zwitterionic (CAPB) micellar surfactant solutions

    J Colloid Interface Sci

    (2015)
  • J.E. Moore et al.

    Wormlike micelle formation of novel alkyl-tri(ethylene glycol)-glucoside carbohydrate surfactants: structure–function relationships and rheology

    J Colloid Interface Sci

    (2018)
  • K.D. Danov et al.

    Shear rheology of mixed protein adsorption layers vs their structure studied by surface force measurements

    Adv Colloid Interface Sci

    (2015)
  • H. Hoffmann et al.

    Influence of ionic surfactants on the viscoelastic properties of zwitterionic surfactant solutions

    Langmuir

    (1992)
  • S. Hofmann et al.

    Shear-induced micellar structures in ternary surfactant mixtures: the influence of the structure of the micellar interface

    J Phys Chem B

    (1998)
  • N. Dan et al.

    Effect of mixing on the morphology of cylindrical micelles

    Langmuir

    (2006)
  • K. Imanishi et al.

    Wormlike micelles of polyoxyethylene alkyl ether mixtures C10E5+C14E5 and C14E5+C14E7: hydrophobic and hydrophilic chain length dependence of the micellar characteristics

    J Phys Chem B

    (2007)
  • F. Nettesheim et al.

    Phase behavior of systems with wormlike micelles

  • H. Rehage et al.

    Rheological properties of viscoelastic surfactant systems

    J Phys Chem

    (1988)
  • H. Rehage et al.

    Viscoelastic surfactant solutions: model systems for rheological research

    Mol Phys

    (1991)
  • N.C. Christov et al.

    Synergistic sphere-to-rod micelle transition in mixed solutions of sodium dodecyl sulfate and cocoamidopropyl betaine

    Langmuir

    (2004)
  • G. Colafemmina et al.

    Lauric acid-induced formation of a lyotropic nematic phase of disk-shaped micelles

    J Phys Chem B

    (2010)
  • S. Ezrahi et al.

    Daily applications of systems with wormlike micelles

  • H.A. Scheraga

    Non-Newtonian viscosity of solutions of ellipsoidal particles

    J Chem Phys

    (1955)
  • P.G. de Gennes

    Scaling concepts in polymer physics. Ithaca

    (1979)
  • M.E. Cates

    Reptation of living polymers: dynamics of entangled polymers in the presence of reversible chain-scission reactions

    Macromolecules

    (1987)
  • M.E. Cates

    Dynamics of living polymers and flexible surfactant micelles: scaling laws for dilution

    J Phys E

    (1988)
  • M.E. Cates

    Nonlinear viscoelasticity of wormlike micelles (and other reversible breakable polymers)

    J Phys Chem

    (1990)
  • M.E. Cates et al.

    Statistics and dynamics of worm-like surfactant micelles

    J Phys Condens Matter

    (1990)
  • M.S. Turner et al.

    Linear viscoelasticity of living polymers: a quantitative probe of chemical relaxation times

    Langmuir

    (1991)
  • Cited by (28)

    • Will biosurfactants replace conventional surfactants?

      2023, Current Opinion in Colloid and Interface Science
    View all citing articles on Scopus
    1

    Present address: Arch UK Biocides Ltd., Hexagon Tower, Crumpsall Vale, Blackley, Manchester M9 8GQ, UK

    View full text