Polyoxyethylene alkyl ether carboxylic acids: An overview of a neglected class of surfactants with multiresponsive properties

Polyoxyethylene alkyl ether carboxylic acids: An overview of a neglected class of surfactants with multiresponsive properties

    Polyoxyethylene alkyl ether carboxylic acids: an overview of a neglected class of surfactants with multiresponsive properties Leonard...

2MB Sizes 276 Downloads 1095 Views

    Polyoxyethylene alkyl ether carboxylic acids: an overview of a neglected class of surfactants with multiresponsive properties Leonardo Chiappisi PII: DOI: Reference:

S0001-8686(17)30356-1 doi: 10.1016/j.cis.2017.10.001 CIS 1834

To appear in:

Advances in Colloid and Interface Science

Received date: Revised date: Accepted date:

10 August 2017 7 October 2017 10 October 2017

Please cite this article as: Chiappisi Leonardo, Polyoxyethylene alkyl ether carboxylic acids: an overview of a neglected class of surfactants with multiresponsive properties, Advances in Colloid and Interface Science (2017), doi: 10.1016/j.cis.2017.10.001

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

ACCEPTED MANUSCRIPT

T

Polyoxyethylene alkyl ether carboxylic acids: an overview of a neglected class of surfactants with multiresponsive properties

RI P

Leonardo Chiappisia,b,∗ a Technische

SC

Universität Berlin, Stranski Laboratorium für Physikalische Chemie und Theoretische Chemie, Institut für Chemie, Straße des 17. Juni 124, Sekr. TC7, D-10623 Berlin, Germany b Institut Max von Laue - Paul Langevin, Large Scale Structures Group, 71 avenue des Martyrs - 38042 Grenoble Cedex 9, France

NU

Abstract

In this work, an overview on aqueous solutions of polyoxyethylene alkyl ether carboxylic acids is given. Unique properties arise from the combination of the nonionic, temperature-responsive polyoxyethylene block with the weakly

MA

ionic, pH-responsive carboxylic acid termination in a single surfactant headgroup. Accordingly, this class of surfactant finds broad application across very different sectors. Despite their large use on an industrial and a technical scale, the literature lacks a systematic and detailed characterization of their physico-chemical properties which is provided

ED

herein. In addition, a comprehensive overview is given of their self-assembly and interfacial behavior, of their use as colloidal building blocks and for large-scale applications.

AC CE

Contents

PT

Keywords: polyoxyethylene alkyl ether carboxylic acids, surfactant science, self-assembly

1

Introduction

2

2

Synthetic procedure and availability

4

3

Physico-chemical properties

4

3.1

pH titrations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

5

3.2

Surface tension . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

8

3.3

Small-angle neutron scattering . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

12

3.4

Temperature responsiveness . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

16

4

Colloidal building blocks and technical uses

18

4.1

Building blocks for colloidal supramolecular structures . . . . . . . . . . . . . . . . . . . . . . . . .

18

4.2

Practical applications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

19

∗ Corresponding

Author. Email address: [email protected] (Leonardo Chiappisi)

Preprint submitted to Advances in Colloid and Interface Science

October 10, 2017

ACCEPTED MANUSCRIPT

5

Conclusions

21 22

RI P

T

References

1. Introduction

Surfactants are among the most versatile compounds in the colloidal playground. The properties of their aque-

SC

ous solutions, e.g., critical micelle concentration (cmc), surface activity, solubilization capacity, foamability, foam stability, depend on the molecular structure of the lyophobic and lyophilic groups. Given their amphiphilic nature,

NU

surfactants are employed as solubilizing agents and detergents, to reduce surface and interfacial tensions, to stabilize (or destabilize) foams and emulsions, to disperse or flocculate solids or liquids in liquids, etc.[1]. Optimizing surfactants for their function is a complex and tedious task, however great efforts have been pursued by the scientific and

MA

industrial communities towards this goal. Recent progress has been well reviewed[2–12]. Carb. Acid termination

Ethyleneoxide units

Alkyl Chain

ED

O

OH

O

i

j

O

PT

Figure 1: Schematic representation of the chemical formula of polyoxyethylene alkyl ether carboxylic acids, composed of a hydrophobic alkyl chain, hydrophilic ethylene oxide units and terminated by a carboxy-methyl unit.

AC CE

This work is focussed on polyoxyethylene alkyl ether carboxylic acids (chemical formula given in Fig. 1), often abbreviated as AECs. For the sake of simplicity, three main parts of the surfactant can be identified: the lipophilic alkyl chain, which can vary in length and degree of saturation; the polyoxyethylene (EO) block, and the terminal carboxylic acid, whose degree of ionization depends on the pH of the solution. Hereafter, the surfactant chemical structure will be abbreviated as Ci Ej CH2 COOH, with i being the usual nomenclature for fatty acids, j representing the number of ethylene oxide units, and CH2 COOH is the carboxymethyl termination. Depending on pH, mixed systems of acids and sodium salts are present, but for the sake of simplicity we name them "acids", keeping in mind the importance of the degree of ionization. The surfactant structure strongly resembles that of nonionic polyoxyethylene alkyl ethers Ci Ej or fatty acids (Ci COOH). Polyoxyethylene alkyl ether carboxylic acids, however, present some major advantages with respect to the above-mentioned compounds, which arise from the combination of ionic and nonionic surfactant properties in a single molecule. Fatty acids – or soaps, when in their salt form – are the oldest man-made surfactants and are easily prepared from the alkaline treatment of natural fats and oils. However, their use is essentially limited by their high Krafft point, i.e., the temperature at which the free energy of the fatty acid molecule is lower in the crystal than in the micellar 2

ACCEPTED MANUSCRIPT

aggregate, and the low solubility in neutral and acidic medium. Optically clear solutions of medium and long-chain fatty acids are obtained only in very alkaline solutions, e.g. over pH ∼ 10.5 for sodium oleate[13, 14]. A simple

T

approach for increasing the solubility of fatty acids consists of exchanging their counterion, usually sodium, with a

RI P

bulkier one with a lower tendency to crystallize[15–18]. Similarly, the Krafft point can be reduced and the tolerance to multivalent cations increased by the use of branched chain fatty acids[19, 20]. Moreover, the low solubility of fatty acid salts of multivalent cations, in particular calcium and magnesium, represents a strong limitation to their use in

SC

everyday consumer products. Although different parameters affect the solubility of fatty acids – pH, temperature, or the presence of multivalent cations – the problem has a common origin: the high stability of the fatty acid crystal and

NU

the low hydrophilicity of the nonionized species. Both factors are addressed when some ethylene oxide units (EO) are inserted between the carboxylic headgroup and the aliphatic chain. The bulky and amorphous EO units prevent AECs from crystallizing and confer a remarkable hydrophilicity even in acidic media, where solubilities and cmcs are close

MA

to the corresponding nonionic polyoxyethylene alkyl ether surfactants[21]. The strategy of introducing some ethylene oxide units between hydrophonic tail and surfactant headgroup is widely applied to alkyl sulfates. Despite they are permanently charged, alkyl sulfates are characterized by high Krafft points and a low solubility in the presence of

ED

multivalent ions[22–24]. Both limitations are, at least partially, overcome, by the insertion of few – generally two to four – ethylene oxide units.

In contrast to their ionic counterparts, nonionic polyoxyethylene alkyl ether are generally water soluble (provided

PT

a sufficient number of EO units are present), their behavior is pH independent and they are resistant to hard water conditions, i.e., in the presence of multivalent cations[4, 25]. Polyoxyethylene alkyl ethers exhibit a rich phase diagram

AC CE

in aqueous solution as a function of temperature and concentration[26]. With increasing temperature the hydrophilicity of the EO block is reduced, resulting in a gradual dehydration of the EO units. Above a certain temperature, called the cloud point (CP), phase separation into a surfactant-rich and a surfactant-poor phase occurs[4, 27]. This phenomenon is common to all polyoxyethylene alkyl ethers, but the CP varies strongly with: the length of the surfactant tail, the number of EO units, and the presence of additives, e.g. inorganic salts or alcohols[28–30]. The complex physics underlying this phenomenon was recently reviewed[4, 27]. The continuous dehydration of the surfactant headgroup has remarkable consequences not only on the phase boundaries, but also on other properties such as the cmc, the adsorption at interfaces, the spontaneous curvature of the surfactant film, and is the basis for several surfactantmediated extraction methods[31–33]. Fatty acids and polyoxyethylene alkyl ethers are both highly relevant classes of surfactants, each exhibiting characteristic properties. However, the combination of the aqueous solubility and the thermoresponsive behavior of nonionic polyoxyethylene alkyl ethers with a pH-responsive carboxylic headgroup opens up a multitude of possibilities, as will be shown hereafter. Current literature lacks in systematic studies of the physico-chemical properties of AECs. To compensate for this lack, a comprehensive physico-chemical characterization of aqueous solutions of polyoxyethylene alkyl ether carboxylic acids is provided in this work. Experimental details are given in the supporting information. In particular, their behavior is rationalized in terms of their chemical structure and experimental conditions such as 3

ACCEPTED MANUSCRIPT

pH, ionic strength, and temperature. These new results are analyzed and discussed jointly with the available literature data, therefore providing a comprehensive overview on the self-assembly properties of AECs. The behavior of

T

polyoxyethylene alkyl ether carboxylic acids will be compared with that of the nonionic polyoxyethylene alkyl ethers

RI P

and strong anionic polyoxyethylene alkyl ether sulfates. Based on this exhaustive characterization, recent progress in fundamental and applied research, demonstrating the potential of these surfactants in different fields, is discussed.

SC

2. Synthetic procedure and availability

AECs are prepared from polyoxyethylene alkyl ethers[34–40]. Polyoxyethylene alkyl ethers are mainly prepared

NU

by the addition of ethylene oxide to fatty alcohols in the presence of a strong base, resulting in a rather broad distribution of EO units[41, 42]. Polyoxyethylene alkyl ether carboxylic acids with branched alkyl chain can be prepared from Guerbet alcohols[43–45]. Two main methods are available for the final modification leading to the carboxymethyl ter-

MA

mination. The first method consists of the reaction of the alcohol with monochloroacetic acid in the presence of sodium hydroxide, i.e. the Williamson’s ether synthesis[34, 38, 46–48].The reaction of the fatty alcohol with chloroacetic acid presents some drawbacks: the low yield of ca. 90 %, the use of the toxic chloroacetic acid, and the formation of a large

ED

amount of sodium chloride and other byproducts as diglycolic and glycolic acids. To avoid the formation of these short chain carboxylic acids, a bulky base such as a secondary or tertiary alkoxide can be used[34, 38, 46]. The second

PT

method consists of the direct oxidation of the terminal alcohol[34, 49–51]. The reaction is mostly performed using oxygen as the oxidizing agent in the presence of a platinum or palladium-based catalyst[34, 49, 50]. The efficiency of the catalyst can be improved by the introduction of Pb, Bi, or Cd salts[34, 52–54], or by the use of a Au-based

AC CE

catalyst[40, 51, 55–57]. Synthetic routes using H2 O2 as oxidizing agent are also reported[51]. Although the oxidation reaction has a yield close to 100%, the main route for the industrial preparation of AECs is the Williamson’s synthesis, as it has proven to be more cost-effective[46, 51, 55]. A further limitation to the implementation of the large-scale oxidation reaction, is the long diffusion time of oxygen into the highly viscous surfactant phase. In contrast, the higher yield and simpler reaction conditions make the oxidation reaction more suitable for academic research purposes. Technical grade polyoxyethylene alkyl ether carboxylic acids are readily available in large quantities and at low cost. The alkyl chain length generally varies between 4 and 20 carbon atoms, and the degree of ethoxylation between 2 and 20. Some of the key manufacturers of AECs are Kao Chemicals (AKYPO R series), Huntsman (EMPICOL R C series), Clariant (Emulsogen R series), Uniqema (Atpol series), Nippon Shokubai, and BASF SE[58].

3. Physico-chemical properties As mentioned previously, the surfactant can be subdivided in three main parts: the lipophilic alkyl chain, the hydrophilic polyoxyethylene part, and the ionizable carboxylic termination. This subdivision helps in rationalizing properties such as the cmc, surface tension, size and shape of the aggregates. Some studies dealing with the self-assembly and interfacial properties of AEC solutions are available in the literature[21, 36, 59–69], but a systematic study is 4

ACCEPTED MANUSCRIPT

missing. To provide a more exhaustive characterization of their properties, the behavior of aqueous solutions of a set of AECs with variable alkyl chain length and number of ethylenoxide units was probed as a function of pH, ionic

T

strength, and temperature. In particular, three technical surfactants, represented by the formulæ C12 E10 CH2 COOH,

RI P

C18:1 E9 CH2 COOH, and C12 E4.5 CH2 COOH were chosen to highlight alkyl chain length and headgroup size effects. This choice was limited to surfactants having a sufficient number of EO units to be water soluble over a large pH range, and an alkyl chain large enough to form well-defined self-assembled aggregates. The shorter alkyl chains denoted by

SC

C12 are in fact 2:1 mixtures of C12 and C14 and the longer alkyl chain denoted as C18:1 is a 3:1 mixture of C18:1 and C16:0 . The number of EO units show a roughly Gaussian distribution, with a distribution width of approximately three

NU

for the C12 E4.5 CH2 COOH surfactant[65]. Further details on the materials and on the techniques used are given in the experimental section in the appendix of this work and in the supporting information. The use of technical surfactants presents two major advantages: their low cost and availability in large quantities, and the increase in stability of the

MA

easy-to-process isotropic micellar and hexagonal phases over that of the lamellar Lα one[26]. Despite small deviation of the physico-chemical properties such as cmc, surface tension, and cloud point of nonionic polyoxyethylene alkyl ether surfactant solutions are reported between pure and non-pure components, no difference is found between the

ED

trends as a function of the average alkyl chain length and number of EO units remains unchanged[26, 70, 71]. This similarities allows generaling the findings of this work on technical surfactants also on their pure counterparts.

PT

3.1. pH titrations

Potentiometric and turbidimetric titrations of 1 wt% solutions of C12 E10 CH2 COOH, C18:1 E9 CH2 COOH, and

AC CE

C12 E4.5 CH2 COOH were conducted at different ionic strengths. All results are reported in Fig. 2. The turbidity of the solution (first row of Fig. 2) reflects the weight average molar mass of the self-assembled surfactant aggregates, the smaller the aggregate the more transparent the solution. C12 E10 CH2 COOH solutions were completely transparent throughout the titration, the C18:1 E9 CH2 COOH were slightly translucent for pH < 3.0, and the C12 E4.5 CH2 COOH turned from turbid to transparent during the titration. Clearly, the larger the number of EO units and the higher the pH, the smaller the aggregates are in solution. Similar turbidity curves were reported earlier for C12 E4.5 CH2 COOH by Kunz and coworkers[65], showing a transition from vesicles to globular micelles. Structural details will be discussed in section 3.3, based on the small-angle neutron scattering (SANS) data. Potentiometric titrations, through the charge balance, directly provide the degree of ionization α of the surfactant, i.e., the fraction of nonprotonated carboxylic units (second row of Fig. 2). The effective acidity constant is then calculated as: pKa (α) = pH − log

α . 1−α

(1)

A monotonic increase in pKa (α) with pH is observed for all surfactants at each ionic strength (third row of Fig. 2). The pKa (α) values are found to be between 3.5 and 5.0, depending on the pH and ionic strength of the solution. In all cases, the acidity of the surfactant increases upon addition of sodium chloride. The strong dependence of the pKa from the experimental conditions makes a direct comparison with literature values difficult. Extrapolating the pKa 5

ACCEPTED MANUSCRIPT

C12E10CH2COOH

C18:1E9CH2COOH

0.8

0.4

NaCl 0.0 M 0.15 M 1.0 M

0.2

NaCl 0.0 M 0.15 M 1.0 M

0.8

NaCl 0.0 M 0.15 M 1.0 M

SC

0.6 0.4 0.2

NU

0.0 5.0 4.5 4.0

MA

pKa( )

T

0.6

0.0

3.5 0 -20

ED

!0( ) / mV

C12E4.5CH2COOH

RI P

Transmission

1.0

-40 -60

2

3

4

PT

-80 5

6

2

3

4 pH

5

6

2

3

4

5

6

AC CE

Figure 2: From top to bottom: Light transmission (wavelength λ = 520 nm, path 2.5 cm), degree of ionization α, effective pKa , and electrostatic surface potential Ψ0 as a function of pH, obtained from turbidimetric and potentiometric titrations of 1 wt% solutions of (from left to right) C12 E10 CH2 COOH, C18:1 E9 CH2 COOH, and C12 E4.5 CH2 COOH with the initial addition of different amounts of NaCl. See text for full details.

values to α → 0 allows for the determination of the intrinsic acidity constant pKa (0) (reported in Table 1). The pKa (0) values are approximately constant, and slightly decrease with increasing number of EO units of the headgroup. This is a result of the electron withdrawing effect of the ethylene oxide units[60]. To a first approximation, the effective acidity constant of a generic polyacid (weak electrolytes, dendrimers, aggregated weak acids/bases, etc.) is determined by two main factors: the intrinsic binding affinity pKa (0), and the electrostatic interaction between the charged headgroups and can be described as[73–76]: ! dGel 1 pKa (α) = pKa (0) + 2.303RT dα

(2)

where Gel , R, and T are the headgroup interaction contribution to the change in Gibbs free energy of ionization, the ideal gas constant, and the temperature, respectively. More complex models, taking into account conformational changes of the colloid as a function of pH or ionic strength, are also available[77–79]. The main contribution to the free energy of ionization arises from the work needed to create an additional charge on the micellar interface, and is 6

ACCEPTED MANUSCRIPT

Table 1: Results from pH titrations obtained at different initial NaCl concentrations (in mol L−1 ).

0.0

3.49 ± 0.03

0.15

3.47 ± 0.01

1.0 C18:1 E9 CH2 COOH

0.0

3.68 ± 0.05

3.50 ± 0.02

1.0

3.67 ± 0.02

0.0

4.00 ± 0.03

NU

C12 E4.5 CH2 COOH

3.60 ± 0.02

SC

0.15

T

pKa (0)

RI P

C12 E10 CH2 COOH

[NaCl]

0.15

3.90 ± 0.01

1.0

3.73 ± 0.02

0.0

3.8±0.2

C12 E4.5 CH2 COOH

0.0

4.67

C12 E10 CH2 COOHc at α = 0.5

0.0

4.0

MA

C8 E5 CH2 COOHa

b

from Ref. [60]; b from Ref. [65]; c from Ref. [72]

ED

a

PT

directly related to the surface potential Ψ0 [74]:

dGel = −e0 NA Ψ0 (α) dα

(3)

AC CE

where e0 and NA are the elementary charge and the Avogadro constant, respectively. The electrostatic potential at the micellar headgroup/water interface can be obtained by combining Eqs. 2 and 3 (fourth row of Fig. 2). The surface potential decreases from 0 mV at low pH to -80 mV in neutral and basic water. Values of up to -50 mV are found in saline alkaline solutions.

Ψ0 can also be obtained by the one-dimensional Poisson-Boltzmann equation for spherical particles immersed in an electrolyte solution[74]:

! 8πe0 I e0 Ψ(r) d 2 dΨ(r) r = sinh dr ε0 εr kB T r2 dr

(4)

with r being the distance from the center of the spherical particle, I the ionic strength of the solution, ε0 εr the dielectric constant of the solvent, and kB the Boltzmann constant. The equation can be solved using the boundary conditions that: Ψ(r → ∞) = 0 and dΨ/dr = −4πσgeo /ε0 εr at r = R, with R being the distance of the headgroup-water interface from the micelle center, and σgeo the geometrical charge density at the headgroup-water interface. σgeo and the radius of the micelle can be estimated from the SANS data reported in section 3.3 of this work. In particular, the geometric charge density is obtained from the degree of ionization α, the aggregation number Nagg , and the micellar radius R as σgeo = αNagg /4πR2 . Eq. 4 was solved iteratively using the Fortran routine made available by Lund University[80] using the values of Nagg and R determined from the SANS patterns (see supporting information for further details). 7

ACCEPTED MANUSCRIPT

The surface electrostatic potential was also determined from the charged hard-sphere structure factor used for the interpretation of the SANS data. The values of the surface potential directly determined from the potentiometric

T

titrations, from the solution of the Poisson-Boltzmann equation (Eq. 4), and from the analysis of the SANS patterns, pH

RI P

S (q) respectively denoted hereafter as Ψ0 , ΨPB 0 , and Ψ0 , are reported in Fig. 3. All values and detailed calculations are

given in the supporting information. Briefly, the electrostatic potential predicted by solving the Poisson-Boltzmann

SC

0 -20

NU

-60

0

pH

/ mV

-40

-100

0

-120

PB

pH

0

S(q)

z

-120

-100

ED

-140 -140

0

MA

-80

-80 0

0

PB

0

S(q)

C18:1E9CH2COOH C18:1E9CH2COOH+NaCl C12E10CH2COOH C12E10CH2COOH+NaCl C12E4.5CH2COOH

S(q)

-60 and

0

PB

-40

-20

0

/ mV pH

Figure 3: Plot reporting the values of the surface potential at the micelle headgroup/water interface determined from pH titrations (Ψ0 ) as a

PT

function of the ones obtained by the solution of the Poisson-Boltzmann equation ΨPB 0 (empty points) and as a function of the values determined S (q)

by the analysis of the structure factor of the SANS patterns Ψ0

(solid points). Exact values used for the calculations are given in Table S3. The

AC CE

dashed line represents f (x) = x and is a guide for the eyes. If the different methods would result in the same surface potential values, all points would fall on the dashed line. The inset at the bottom-left corner schematically shows the micelle/water interface as a function of the distance from the micelle center z, with the location of the plane at which the potentials are calculated/determined. See text for further details.

equation is significantly more negative than the one found from pH titrations. In contrast, slightly less negative values are found for Ψ0S (q) . This discrepancy is thought to arise from the effect of the micelle geometry on the calculation S (q) of ΨPB 0 and Ψ0 . In both cases, the micelles are modeled as rigid, spherical particles. However, analysis of the

SANS revealed that the micelles are in fact slightly anisotropic. ΨPB 0 is calculated at the surface of a rigid sphere with a volume equivalent of that of the globular micelle; Ψ0S (q) is calculated at the surface of the sphere with a radius equal to the hard sphere radius, close to the largest dimension of the globular micelle. The ζ-potential values, i.e., the electrostatic potential at the slipping plane, are in the range -20 – -30 mV for C18:1 E9 CH2 COOH at pH 4 to 5[68], in line with the trend of decreasing potential with increasing distance from the micelle headgroup/water interface. Moreover, due to the flexibility of the ethylene oxide block, and its size distribution, the charges of the micelles are distributed in an interfacial region rather than on a sharp plane, allowing for more spatial freedom to minimize the electrostatic repulsion. Despite the substantial approximations, a good agreement is found between the different values. This demonstrates that the dependence of acidity constant on pH for the micellized AECs is predominantly governed by the electrostatic repulsion and counterion condensation, and that the shape and size of the self-assembled 8

ACCEPTED MANUSCRIPT

aggregates play only a minor role. The effects of the addition of sodium chloride to a surfactant solution at constant pH can be rationalized accordingly. Due to counterion condensation and charge screening, the electrostatic potential

T

at the micelle/water interface is reduced, essentially lowering the repulsion between the charged headgroups. This

RI P

screening is observed in the decrease in the pKa (α) and, in turn, in an increase in the degree of ionization α. 3.2. Surface tension

50 45 40 35

10-7

ED

30 25

pH = 2.5 4.1 5.0 8.0

MA

/ mN m-1

55

pH = 2.5 4.1 5.0 8.0

NU

pH = 2.5 4.1 5.0 8.0 8.0 + 0.2M NaCl

60

Surface tension

SC

65

10-6 10-5 10-4 10-3 10-2 10-7 10-6 10-5 10-4 10-3 10-2 10-7 10-6 10-5 10-4 10-3 10-2 [C12E10CH2COOH] / mol kg-1 [C18:1E9CH2COOH] / mol kg-1 [C12E4.5CH2COOH] / mol kg-1

PT

Figure 4: Equilibrium surface tension of (from left to right) C12 E10 CH2 COOH, C18:1 E9 CH2 COOH, and C12 E4.5 CH2 COOH solutions in water of varying acidity as a function of the molality of surfactant.

AC CE

The equilibrium surface tension of C12 E10 CH2 COOH, C18:1 E9 CH2 COOH, and C12 E4.5 CH2 COOH solutions at different surfactant concentration and pH was measured and is shown in Fig. 4. The recorded curves are clearly different from typical titration curves. A clear minimum in the surface tension curve is observed, which is deeper the higher the pH of the solution, i.e., at higher degree of ionization of the surfactant. Technical surfactants, where a large spectrum of molecules with variable surface activity coexist, usually – but not necessarily – show a minimum in their surface tension vs. log C curves[81]. Few examples of surface tension curves with the same shape were reported also for similar, monodisperse surfactant systems[61, 82]. In addition to the polydisperse nature of the surfactant, the equilibrium between the charged and uncharged forms adds a certain degree of complexity to the data interpretation. Different theoretical models for the descriptions of the adsorption of ionic/nonionic surfactant at interfaces are available, some of them explicitly consider electrostatic interactions while others work within the thermodynamic concept of a by-definition-neutral Gibbs dividing interface[83]. Given the complexity of the studied system, the surface tension titration curves were interpreted within the Gibbsian framework. At constant pressure and temperature, the Gibbs adsorption isotherm is given as: dγ = −Γ s RT d ln a s − Γc RT d ln ac

9

(5)

ACCEPTED MANUSCRIPT

where γ is the surface tension, Γ s , a s , Γc , and ac are the surface excess and the activity of the surfactant and its coun-

and

χc = αχ s

(6)

RI P

Γc = αi Γ s

T

terion, respectively. To preserve the electroneutrality of interface and solution, the following relations are introduced:

where αi , χ s , and χc are the degree of ionization of the surfactant at the interface, and the molar fractions of the surfactant and the counterion, respectively. αi is not to be confused with the degree of ionization of the surfactant

SC

in solution α determined from potentiometric titrations. Small difference of 0.2[84] and 1 unit[85] are reported between the pKa values of fatty acids determined in bulk solution or at the air-water interface. This differences arises from from the electrostatic potential at the air-water interface established upon the adsorption of charged species[84].

NU

Assuming that the potential at the micelle-water interface is similar to that at the air-water interface, it is reasonable to approximate that αi ≃ α. At least for the more dilute region of the titration curves, where the surfactant is present

MA

in its nonmicellized form, ideal behavior can be assumed, such that the activities of the surfactant and its counterion may be substituted with their concentrations. Eq. 5 can be rewritten accordingly:

ED

dγ ≈ −Γ s (1 + αi )RT d ln χ s − αi Γ s d ln αi

(7)

As the titrations are performed at constant pH, it is reasonable to neglect the second term of Eq. 7. Finally, the

PT

surfactant surface excess can be obtained as:

Γs = −

dγ 1 (1 + α)RT d ln χ s

(8)

AC CE

When working in excess of an inert salt, i.e., a salt presenting no specific interaction with the surfactant, the second term of Eq. 5 is zero and Eq. 7 simplifies to that of a nonionic surfactant. The minimum area per molecule at the air-water interface Ah is calculated from Γ s as Ah = (NA Γ s )−1 , where NA is the Avogadro constant. Within the same thermodynamic framework, the Gibbs free energy of micellization ∆G0mic can be obtained. The aggregation of n partly charged surfactant molecules S α− and their counterions C + into a micelle Mn can be expressed with the following equilibrium:

nS α− + nα(1 − β)C + ⇋ Mnnαβ−

(9)

where (1-β) is the fraction of condensed counterions. The ∆G0mic can then be calculated as: ∆G0mic 1 = ln a M − ln a s − α(1 − β) ln ac . RT n

(10)

Substituting the activities with the molar fractions, the following relation is obtained: ∆G0mic   ≈ 1 + α(1 − β) ln χ s (cmc) + α(1 − β) ln α. RT

(11)

Eq. 11 reduces to the classical equation of ionic surfactants[90] for α = 1: ∆G0mic ≈ (2 − β) ln χ s (cmc) RT 10

(12)

T

ACCEPTED MANUSCRIPT

RI P

Table 2: Summary of parameters obtained from surface tension titrations and of those used for its analysis. α is the degree of ionization obtained from potentiometric titrations, 1 − β is the fraction of condensed counterions, determined from SANS experiments. The cmc is given in µmol L−1 , the headgroup area Ah in Å2 , the surface excess Γ in µmol m−2 , the surface tension dependence from the concentration dγ/d ln(C) in 10−3 N m−1 , the surface tension at high surfactant concentration γ(c → ∞) and at the cmc γ(c = cmc) in 10−3 N m−1 , the Gibbs free energy of micellization

SC

∆G0mic in 103 J mol−1 . Values for nonionic surfactants are reported for comparison at the bottom of the table. Values in italics are from literature.

a

Values determined in 0.2 M acetic acid/sodium acetate buffer, from Ref. 21. Other values are taken from: Ref. 86; Ref. 87; Ref. 61 (chemically pure polyoxyethylene alkyl ether carboxylic acid); e Ref. 88 f Ref. 89; g Ref. 24.

d

γ(c → ∞)

γ(c = cmc)

∆G0mic

1.9

-5.3

31.1

30.3

-40.7





33.5

31.9



β

cmc

Ah

Γ

2.5

0.09

0.0

20.5

86

4.0a





29.0



4.1

0.49

0.45

5.0

0.78

8.0 NaCl

C18:1 E9 CH2 COOH

102

1.6

33.5

31.1

-44.2

0.45

100

110

1.5

-6.8

37.5

31.8

-47.3

1.00

0.45

MA

-6.0

180

130

1.3

-6.2

40.7

35.5

-48.7

8.0

1.00

1.0

75

75

2.2

-5.5

37.7

35.7

-47.6

2.5

0.05

0.0

2.0

51

3.3

-8.5

31.3

31.8

-45.1



5.5







33.6

32.0



0.3

0.4

14

96

1.7

-6.0

31.1

32.7

-45.1

0.7

0.4

24

97

1.7

-7.2

34.7

33.8

-52.1

a

4.0

4.1 5.0



PT

−1

ED

54

AC CE

NU

α

+ 0.2 mol L

c

dγ/d ln(C)

pH C12 E10 CH2 COOH

b

8.0

1.0

0.4

43

124

1.3

-6.7

39.2

37.0

-56.0

2.5

0.05

0.0

5.5

58

2.8

-7.4

28.0

27.9

-42.5

4.0





15







28.8

30.5



4.1

0.30

0.15

26

68

2.4

-7.8

28.2

28.7

-45.9

5.0

0.60

0.15

150

81

2.0

-8.1

30.4

28.6

-48.5

∼ 7d





400







39.1

30.5



8.0

1.00

0.25

450

123

1.4

-6.7

35.4

34.0

-51.0

C12 E10 (at 303 K)b







12.7











-38.5

C16 E9







2.0

53





36

36

-42.5







6.6







31

30

-39.5



1.0



2200

52







43





1.0



1300













C12 E4.5 CH2 COOH

a

c

C12 E5 e C12 E3 OSO3 Na

f

C12 E10 OSO3 Nag

11

ACCEPTED MANUSCRIPT

It should be noted that by introducing the relation χc ≈ αχ s from Eq. 6, the contribution of hydroxide ions to the charge balance of the system is neglected. However, this approximation holds in the range of pH (2.5 to 8) and

T

surfactant concentration to which Eqs. 7 and 11 are applied. The critical micelle concentration is taken either at the

RI P

abrupt change in slope of the surface tension titration curves or, if present, at the minimum value. The values are summarized in Table 2.

For each surfactant, increasing the degree of ionization α causes an increase of the cmc of up to two orders of

SC

magnitude. The values of β used for the calculation of ∆G0mic were obtained from the analysis of the SANS patterns given in Fig. 5. The values of β ≃ 0.3 - 0.5 found for C12 E10 CH2 COOH and C18:1 E9 CH2 COOH are relatively

NU

high when compared to the typical values of 0.1-0.2 for weak and strong ionic surfactants in water[91–94]. Such low counterion condensation arises from the presence of the EO units, which “dilute” the charge of the surfactant headgroup[23, 60, 95], as also supported by the lower values of β found for C12 E4.5 CH2 COOH. The consequences of

MA

the presence of the EO units are also evident when comparing the cmc values found in solutions of C12 E10 CH2 COOH and C12 E4.5 H2 COOH at different pH. At low pH, the cmc is higher for the surfactant with the larger headgroup, as expected by the hydrophilic nature of EO units. In contrast, at high pH, where the ionic form of the surfactant

ED

dominates, the cmc of C12 E10 CH2 COOH is a factor 2 lower than that of C12 E4.5 CH2 COOH. A similar reduction of the cmc with increasing number of EO units was also reported for polyethylene oxide alkyl ether sulfates[24, 89, 96]. effects of the EO units.

PT

This counterintuitive relation of increasing hydrophilicity and decreasing cmc can be ascribed to the charge-dilution Moreover, the values for the cmc and the ∆G0mic are close to those reported for similar nonionic Ci Ej [61, 86–88].

AC CE

With increasing pH a decrease in the free energy of micellization is observed. This trend is in contrast to an expected increase due to the electrostatic repulsion between the charged headgroups, which should become more important the higher the degree of ionization of the surfactant. One explanation lies in high sensitivity of the calculation of ∆G0mic via Eq. 11 on the values of β. The β values used were determined from the analysis of the SANS patterns at surfactant concentrations much higher than the cmc and may be underestimated, in particular when compared with the high values of 0.7-0.85, determined by conductometric experiments close to the cmc for the sodium salts of several AECs in pure water (resulting pH of 8)[60, 69]. In addition to strongly increasing the cmc of the surfactant, the ionization of the carboxylic headgroup causes a decrease of the maximal surfactant surface excess Γ s and, accordingly, an increase of the minimal area occupied by the surfactant at the air-water interface. This effect is particularly pronounced for C12 E4.5 CH2 COOH, the surfactant with the smallest number of EO units. However, the reported values must be treated with caution given the technical nature of the surfactant used and the uncertainty arising from the approximation of α ≃ αi . All surfactants show a minimum in the surface tension curves, at least at higher pH values. The technical nature of the surfactant can only partially explain this behavior, as the polydispersity in alkyl chain length and number of EO units is pH independent. However, the difference in surface activity between the main component and impurities, e.g., the non-carboxymethylated polyoxyethylene alkyl ether, increases with increasing pH. In addition, at pH 2.5 12

ACCEPTED MANUSCRIPT

and 8.0 the surfactant can be regarded as a pure nonionic or pure ionic compound, respectively. Sakai et al., however, recorded the surface tension for the chemically pure C12 E4 CH2 COOH surfactant at pH ∼ 7, reporting a similar, well-

T

pronounced minimum in the surface tension close to the cmc[61]. The authors justify the observed minimum with

RI P

a vesicle-to-micelle structural transition in the proximity of the cmc. Zana et al. reported a continuous increase in surfactant aggregation number for C12 E5 CH2 COOH, C12 E7 CH2 COOH, and C12 E9 CH2 COOH close to the cmc[60]. A similar effect was found in another pure, weakly-anionic surfactant[82]. In both cases, the authors assumed structural

SC

rearrangements to justify the shape of the surface tension curves. A common assumption is that the concentration and therefore the activity of the free surfactant remains constant or increases slightly after the cmc. However, it was

NU

demonstrated that the activity of free surfactant can decrease with increasing total surfactant concentration above the cmc[97, 98], due to the increased ionic strength which reduces the repulsion between the surfactant headgroups and the micelles. To address the effect arising from the increasing ionic strength, a surface tension curve of C12 E10 CH2 COOH

MA

at a pH value of 8 in the presence of 0.2 mol L−1 NaCl was recorded. The minimum in the surface tension curve is reduced but is still present, in line with the findings by Svanedal et al.[82]. In summary, the uncommon shape of the surface tension curves arises from a number of different factors: the polydisperse nature of the surfactant, a

ED

likely change in the micelle geometry at concentrations higher than the cmc and, more importantly, a reduction of the activity coefficient of the surfactant, resulting in a well-pronounced minimum in the surface tension curve at high pH.

100

C18:1E9CH2COOH pD = 2.0 2.8 4.4 4.1 6.1 11.7

104

102 101

0.1

104 103 102 101

100 10-1

C12E4.5CH2COOH pD = 2.0 2.7 4.3 4.7 5.0 5.3 6.6 11.7

106

I(q) / cm-1

103 I(q) / cm-1

101

C12E10CH2COOH pD = 2.0 2.8 4.2 4.8 5.2 5.9 11.4

AC CE

I(q) / cm-1

102

PT

3.3. Small-angle neutron scattering

100

10-1 1 q / nm-1

10-1 0.1

q / nm-1

1

0.1

q / nm-1

1

Figure 5: SANS patterns arising from ca. 1 wt% solution of (from left to right) C12 E10 CH2 COOH, C18:1 E9 CH2 COOH, and C12 E4.5 CH2 COOH. Solid lines show best fits according to different models described in detail in the main text and supporting information. The models are schematically represented by the sketches, not to scale, in the figure.

The size and shape of three representative surfactants, C18:1 E9 CH2 COOH, C12 E10 CH2 COOH, and C12 E4.5 CH2 COOH were studied in D2 O as a function of pD via small-angle neutron scattering (SANS). The scattering patterns could be 13

ACCEPTED MANUSCRIPT

described using different models, reported in detail in the supporting information. In all cases, a homogeneous coreshell structure was assumed, where the core is formed by the anhydrous hydrophobic blocks, and the shell is formed

T

by the hydrated EO units and the carboxymethyl unit. Depending on the nature of the surfactant and on the acidity

RI P

of the solution, the data were described using vesicle, rod or ellipsoid models[99]. A charged hard-sphere structure factor was used for the ellipsoids[100]. The obtained parameters are given in Tables 3 and 4, and the analytical expressions used for the description of the SANS patterns are given in the ESI. The fits were constrained by imposing

SC

the condition that at least one characteristic length of the aggregate core should be 1.6 and 2.1 nm, as determined by the maximum length of the C14 and C16 alkyl chains of the surfactants, respectively (see supporting information for

NU

further details). The thickness of the shell depends on the conformation of the ethylene oxide units[101–103], and its limits can be estimated assuming different conformations: 0.35 nm per EO unit for the fully extended chain as in polyoxyethylene crystals, 0.19 nm per EO unit for the meander conformation.

MA

Scattering patterns of ca. 1 wt% aqueous solutions are shown in Fig. 5, additional curves determined in a 150 mM NaCl solution are given in the supporting information. At high pD, the scattering behavior of the different surfactants is similar, and a peak is found at q p = 0.3−0.4 nm−1 due to the electrostatic repulsion between the micelles. This value

ED

corresponds to a most probable distance of 2π/q p ∼ 15 − 20 nm and a dehydrated micelle volume of ν = 8π3 φ/q p of 40-100 nm3 , with φ being the surfactant volume fraction. At lower pD, the surfactants self-assemble into very different structures.

PT

C12 E10 CH2 COOH is the surfactant with the largest head-to-tail size ratio investigated in this work and, accordingly, it forms the aggregates with the largest curvature. The micelles formed are slightly anisotropic at low pD but

AC CE

become increasingly spherical as the pD is increased. With increasing pD, the shell thickness decreases from ca. 1.6 to 1.4 nm with an approximately constant hydration of 60-70 vol% D2 O in the shell. The micelles at low pD are structurally very similar to those formed by the nonionic C12 E10 [104]. The addition of 150 mM NaCl appears to have no effect on the micelle size and shape. C18:1 E9 CH2 COOH has a similar headgroup to C12 E10 CH2 COOH but a longer alkyl chain and the self-assembled structures vary with pD. In particular, at low pD, when the surfactant is found in its nonionic form, cylindrical micelles are found, as evidenced by the large q−1 region in the SANS patterns. Increasing the degree of ionization of the surfactant causes a transition from rodlike to ellipsoidal micelles. The transition from rods to ellipsoids causes a reduction of the shell thickness from 1.8 to 1.5 nm, while the shell hydration increases from 50 to 70 vol% D2 O. The structures show only minimal changes upon addition of 150 mM NaCl or 1 mM CuCl2 [66]. C12 E4.5 CH2 COOH features the smallest headgroup of the investigated surfactants. The surfactant self-assembles into planar structures up to pD ∼ 5 as evidenced by the large q−2 region in the SANS spectrum. The SANS patterns can be described using a model of large polydisperse unilamellar vesicle at pD < 5. The overall size of the vesicle is outside of the probed q-range and the value given for the radius (200 nm) has no physical meaning. The addition of 150 mM NaCl at pD 4.3 triggers the formation of small vesicles with low polydispersity, and a core radius of ca. 18 nm. Globular oblate ellipsoidal micelles appear at pD > 4.3, which become increasingly spherical with increasing pD. 14

ACCEPTED MANUSCRIPT

Table 3: Summary of fit parameters used for the description with an ellipsoidal model of the scattering curves reported in Fig. 5. χ is the fraction of surfactant in the globular micellar state; the ionic strength I is given in 10−3 mol L−1 , the ellipsoidal rotational semi-axis A, equatorial semi-axis B,

T

shell thickness T, and the hard sphere radius RHS are given in nm; Z is the charge per micelle, given in elementary charge units determined from the

RI P

structure factor, Nagg is the number of molecules per aggregate; αS (q) is given by Z/Nagg ; β = Z/αNagg is the fraction of noncondensed counterions,

B

T

RHS

1

8

2.0

3.7

1.6

1.6

3.5

1

7

2.8

3.6

1.6

1.6

1

6

4.2

2.8

1.6

1.6

1

4

4.8

2.2

1.6

1

4

5.2

2.1

1.6

1

4

5.9

1.6

1.6

1

5

11.4

1.6

1

153

2.0

1

149

4.8

1

150

+NaCl

φwshell

αS (q)

β

Acs h

Ahsw

0.0

117

0.58

0.00



53

161

3.7

0.0

114

0.58

0.00

0.00

54

163

4.3

20.0

88

0.62

0.23

0.56

56

183

1

8

1

6

NU 4.4

19.0

69

0.66

0.27

0.46

58

209

1.6

4.2

19.5

66

0.67

0.29

0.43

59

215

1.6

4.2

21.3

50

0.71

0.42

0.47

64

255

1.5

1.4

4.0

20.8

44

0.67

0.47

0.47

67

244

3.8

1.6

1.6

3.5



47

0.77





82

159

2.7

1.6

1.6

3.7



85

0.63





56

187

MA

1.6

2.3

1.5

1.5

3.5



64

0.64





61

210

2.0

500

2

1.8



0.0

18000

0.49

0.00



54

103

2.8

60

2

1.7



0.0

2200

0.47

0.00

0.00

54

103

5

4.4

4.0

2.0

1.5

5.3

28.2

145

0.58

0.20

0.56

59

149

4

5.1

3.8

2.0

1.5

5.0

26.8

137

0.59

0.20

0.39

60

152

1

5

6.1

3.0

2.0

1.5

4.9

25.4

108

0.63

0.23

0.27

62

170

1

7

11.7

2.9

2.0

1.5

5.1

30.5

105

0.63

0.29

0.29

63

173

1

152

2.0

500

2.0

1.8

0.0



18000

0.49





54

103

1

149

4.4

8.5

2.0

1.7

3.6



307

0.57





56

127

1

152

11.7

3.3

2.0

1.7

4.8



119

0.68





61

179

0

8

2.0























0

6

2.7























0.6

5

4.3

1.6

6.0

0.9





710

0.06





37

52

0.85

4

4.8

1.6

3.8

0.9

5.8

24.0

285

0.28

0.08

0.12

41

69

0.925

4

5.0

1.6

3.5

0.8

5.6

27.1

241

0.21

0.11

0.14

43

70

0.985

4

5.3

1.6

3.1

0.8

4.5

27.4

189

0.27

0.14

0.17

45

76

1

5

6.6

1.6

2.3

0.8

3.8

24.4

104

0.42

0.23

0.23

51

99

1

7

11.7

1.6

2.2

0.8

3.8

25.6

95

0.44

0.27

0.27

53

103

0

151

2.0























0

148

4.3





















1

152

11.7

1.6

— 15 2.6 0.8

4.3



133

0.36





24

54

AC CE

1

C12 E4.5 CH2 COOH

Nagg

11.4

1

+NaCl

Z

SC

A

prolate ellipsoid −→ cylinder

C18:1 E9 CH2 COOH

pD

ED

+NaCl

I

PT

C12 E10 CH2 COOH

χ

oblate ellipsoid −→ bilayers/vesicles

respectively, given in Å2 .

sphere −→ prolate ellipsoid

and Ahsw are the areas per surfactant molecule at the core-shell and at the shell-water interface, φwshell is the volume fraction of D2 O in the shell; Acs h

ACCEPTED MANUSCRIPT

Table 4: Summary of fit parameters used for the description with a vesicle model of the scattering curves arising from 1 wt% solutions of C12 E4.5 CH2 COOH. 1-χ is the fraction of surfactant in the vesicular state; the ionic strength I is given in 10−3 mol L−1 ; the core radius Rc ,

mixture of large (Rc >200 nm) and small (Rc = 6.5 nm) vesicles.

I

pD

Rc

σ

1.00

0.2

The pattern recorded at pD 2.7 was described as a

TEO

TALK

hRi

φwshell

1.2

1.7

210

0.29

1.2

1.7

210

0.29

2.0

>200

0.67

6

2.7

>200

0.2

0.33(a)

6

2.7

8.1

0.2

1.2

1.7

8.3

0.29

0.28

5

4.3

>200

0.2

1.2

1.7

210

0.29

0.15

4

4.8

>200

0.2

1.2

1.7

210

0.29

0.08

4

5.0

>200

0.2

1.2

1.7

210

0.29

0.02

4

5.3

>200

0.2

1.2

1.7

210

0.29

0.00

5

6.6













0.00

7

11.7













1.00

151

2.0

>200

0.2

1.2

1.7

204

0.29

1.00

148

4.3

19

0.2

1.2

1.7

19.9

0.29

0.00

152

11.7













NU

PT

ED

+NaCl

SC

8

(a)

MA

C12 E4.5 CH2 COOH

1-χ

(a)

RI P

bution; hRi is the mean core radius in nm; φwshell is the volume fraction of D2 O in the shell.

T

thicknesses of the hydrophilic TEO and hydrophobic layer TALK are given in nm; σ is the standard deviation of the core radius lognormal distri-

Until pD < 5 they coexist with vesicular aggregates, in agreement with the previous studies of Kunz et al.[64, 65]. It

AC CE

should be remarked that, due to the coexistence of different types of structures, the SANS analysis is not univocal for intermediate pD regions, and the sizes and relative amounts of the different structures should be treated with caution. The gradual transition from large vesicular aggregates to small micelles however, is unequivocal. Aggregation number and micelle size of C12 E4 SO4 Na or C12 E6 SO4 Na are similar to those found for C12 E4.5 CH2 COOH at high pH, i.e., where the surfactant headgroup is fully charged.[105] In summary, AECs self-assemble in dilute solutions into a wide range of morphologies, depending on the pD of the solution and on the size ratio of their hydrophilic and hydrophobic parts. In particular, the spontaneous formation of vesicles by a single anionic surfactant at such mild pD values is a rare finding. By addition of appropriate amounts of dodecanoate, the pD range in which vesicles are found can be extended[64]. The structures formed by the investigated AECs are graphically summarized in Fig. 6 as a function of the volume fraction of the hydrophilic part of the surfactant and the degree of ionization. On the one hand, VEO /Vtot is directly related to the spontaneous curvature of nonionic Ci Ej surfactants and was already successfully employed to rationalize their the phase behavior[106–108]. On the other hand, the degree of ionization determined by potentiometric titrations quantifies the electrostatic contributions to the headgroup requirement and, accordingly, its influence on size and shape of the self-assembled aggregates. It should be noted here, that the choice of these two variables does not do justice to the complexity of system. Nevertheless, 16

ACCEPTED MANUSCRIPT

T

0.8

RI P

0.6 0. 6 0.4 0. 0.2

0.5

0.55

0.6

0.65 0

NU

0.0 0. 0 0.45 45

SC

degree of ioni de ionization nization on

1.0

VEO/Vtot

Figure 6: Summary of SANS-models used for the description of the self-assembled aggregates of polyoxyethylene alkyl ether carboxylic acids as

MA

a function of degree of ionization α and volume fraction of the hydrophilic part of the surfactant. For details of the model used see Tables 3 and 4. Dotted lines are guide for the eye. , N, and • represent samples with anisotropy values less than 1.5, between 1.5 and 3, and larger than 3, respectively. H represents samples forming bilayers.  represents samples where the coexistence of vesicles and globular micelle was observed.

ED

Sketches are not to scale.

it offers a simple approach to rationalize the two nonorthogonal variables to predict the shape of the self-assembled

PT

surfactant aggregates.

Few studies on concentrated AECs systems are reported in litterature[36, 69]. The investigation of the phase behavior of polydisperse C12,13 E4 CH2 COOH and C12-15 E6 CH2 COOH revealed the presence of extended gel regions[36].

AC CE

A more detailed investigation by polarized optical microscopy and small-angle X-ray scattering was performed on the monodisperse sodium salts of C14 E4 CH2 COOH and C14 E7 CH2 COOH in water at 25 ◦ C[69]. The results are summarized in Fig. 7 together with the phase behavior of the strongly anionic C12 E2 OSO3 Na[109] and sodium dodecyl sulfate C12 OSO3 Na, often abbreviated as SDS or SLS[110], and the nonionic C14 E8 [111] and C12 E3 and C12 E5 [26] surfactants. The sodium salts of the AECs show a similar behavior, with the presence of an isotropic micellar solution L1 at concentrations . 25 wt%, followed by a hexagonal H1 phase, which is more extended for the surfactant with the larger headgroup. Both surfactants present a rather extended lamellar Lα phase at concentrations between ∼ 40 and 80 wt%, followed by a sponge L3 phase at high concentration. Only for C14 E4 CH2 COOH a reverse cubic phase V2 at high concentration is reported, although it should be noted that this phase can be easily overseen. In contrast and despite the fewer number of ethylene oxide units, the strongly anionic C12 E2 OSO3 Na forms phases with a larger curvature over the whole concentration range, with the boundaries between the L1 and Lα phases and the Lα and H1 being located at higher surfactant concentration. It is however evident, that the introduction of the 2 ethylene oxide units between the dodecyl chain and the sulfate headgroup enriches the phase behavior, mainly by preventing the crystallization of the surfactant. To be noted, this surfactant is of technical purity. Moreover, a comparison between C14 E4 CH2 COONa and C12 E5 reveals that the ionic headgroup stabilizes the hexagonal phases at low to moderate 17

ACCEPTED MANUSCRIPT

L1

S+L1

L1

H1

L1 + L

C12E3

L1

C14E4CH2COONa

L

C14E7CH2COONa

H1

25

50

V2

L2

L3

V1

SC

0

L H1

L1

C14E8

L3

L

L1 L1

L2

L H1

C12E5

L

T

C12E2OSO3Na

RI P

C12OSO3Na

L

75

L

+S 100

Surfactant concentration wt%

NU

Figure 7: Simplified representation of the phase behavior of alkyl sulfates (green color), alkyl carboxylates (blue color) and nonionic alkyl ethoxylates (red color) sorted by number of EO units, taken from different studies performed at 25 ◦ C in water: the sodium salts of C14 E4 CH2 COOH and C14 E7 CH2 COOH from Ref. 69, the strongly anionic C12 E2 OSO3 Na (technical grade) from Ref. 109, and C12 OSO3 Na from Ref. 110 and the

MA

nonionic C14 E8 from Ref. 111 and C12 E3 and C12 E5 from Ref. 26. Phase notations are: L1 for the isotropic micellar solution; H1 for the hexagonal phase; Lα for the lamellar phase; L2 for the hydrated liquid surfactant phase; V1 for the cubic phase; L3 for the sponge phase; V2 for the reverse cubic phase, and S the for solid surfactant phase.

ED

concentrations. The transition between lamellar and reverse phases takes place at similar concentrations. In addition, the comparison between C14 E7 CH2 COONa and C14 E8 shows that AECs form hexagonal and lamellar phases at lower surfactant concentration than their nonionic counterpart. This observation is further supported by a comparison with

PT

the extended phase diagram of polyoxyethylene dodecyl- and oleyl- ethers, which shows that the lowest concentration at which Ci Ej form hexagonal phases in water at 25 ◦ C is ca. 40 wt% for dodecyl ethers[108] and 25 wt% for

AC CE

oleyl ethers[107]. It should be kept in mind that the comparison between the different classes of surfactants is only indicative and a direct correlation cannot be made, as experimental parameters such as temperature, ionic strength, pH, strongly affect the behavior of the surfactants in a very different fashion. 3.4. Temperature responsiveness

The thermoresponsive properties of 1 wt% aqueous solutions of C12 E10 CH2 COOH, C18:1 E9 CH2 COOH, and C12 E4.5 CH2 COOH at different pH were probed by following the turbidity of the solution by recording the light scattering intensity at an angle of 90◦ (see Fig. 8). Solutions of C12 E10 CH2 COOH show a cloud point temperature of 68 ◦

C at pH 2.5, while no micelle growth can be observed at higher pH values, when the surfactant is partially or fully

charged. C18:1 E9 CH2 COOH, with a more hydrophobic alkyl chain, exhibits its cloud point at pH 2.5 at 47 ◦ C. At pH 4.1 (α = 0.4) a micelle growth is observed to start at ca. 55 ◦ C and terminate with phase separation at 92 ◦ C. The structures are not sensitive to temperature at higher pH values. Solutions of C12 E4.5 CH2 COOH can be seen to exhibit a different behavior. Turbid solutions are found at low pH, and a precipitate slowly forms at 25 ◦ C, 1 wt%, and pH < 3.5, as reported by Kunz et al.[65]. When heated, the solution shows constant turbidity up to 55 ◦ C, at which point all surfactant precipitates leaving a clear solution. At pH 4.1 a constant turbidity is recorded over the whole probed temperature range. Interestingly, at pH 5.0, a bluish solution at low temperature becomes clear for T > 20 ◦ C. A further 18

ACCEPTED MANUSCRIPT

pH = 2.5 4.1 5.0 8.0

RI P

103

102

101

SC

Mean count rate / 103 cps

pH = 2.5 4.1 5.0 8.0

T

pH = 2.5 4.1 5.0 8.0

104

100 20

40

60

80

100 0

20

40 60 Temperature / °C

NU

0

80

100 0

20

40

60

80

100

MA

Figure 8: Scattering intensity determined at a scattering angle of 90◦ as function of temperature for 1 wt% solutions of (from left to right) C12 E10 CH2 COOH, C18:1 E9 CH2 COOH, and C12 E4.5 CH2 COOH at different pH values.

increase in temperature causes a slight micellar growth. Finally, at high pH, C12 E4.5 CH2 COOH self-assembles into

ED

small micelles whose size is temperature independent. The cloud points determined at pH 2.5, where α ∼ 0.0, is lower by ca. 20-30 ◦ C than that of their nonionic alkyl ether analogs, as shown in Table 5. A similar reduction in cloud point temperature was found for nonionic

PT

alkyl ethers where the terminal OH group is substituted with an OCH3 moiety[112–114]. The lower cloud point temperature of methyl-capped alkyl ethers with respect to the conventional, hydroxy-terminated ones, is ascribed to

AC CE

different hydration properties, with the methyl-capped ones showing smaller hydration numbers[114]. Table 5: Cloud point temperatures recorded for 1 wt% aqueous solutions of C12 E10 CH2 COOH, C18:1 E9 CH2 COOH, and C12 E4.5 CH2 COOH at different pH values. The values for the corresponding nonionic alkyl ether surfactants, which are pH-independent, are reported for comparison and were measured in water at a concentration of 1 wt%.

pH = 2.5

pH = 4.1

pH = 5.0

pH = 8.0

C12 E10 CH2 COOH

68◦ C







C18:1 E9 CH2 COOH

47◦ C

>92◦ C





C12 E4.5 CH2 COOH

<5◦ C

<5◦ C





C12 E10 from Refs. 1, 33, and 115:

77-88 ◦ C

C18:1 E10 from Refs. 116, 31, and 117:

72-85 ◦ C

C12 E5 from Refs. 118, 119, and 27:

30-32 ◦ C

The molecular mechanism underlying the clouding phenomenon in dilute nonionic surfactant solutions has been intensively discussed[4, 25, 27, 120]. For what concerns nonionic, polyethylene oxide-based surfactants, the phaseseparation is ascribed to the reduced hydrophilicity and progressive dehydration of the EO units with increasing 19

ACCEPTED MANUSCRIPT

temperature. However, the nature of the critical fluctuations giving rise to the turbidity of the dilute solutions of Ci Ej solution is less evident and still a matter of debate. Different mechanisms apply as a function of the surfactant

T

headgroup size: on the one hand, amphiphiles with a smaller headgroup were reported to continuously grow into long,

RI P

wormlike micelles[119, 121]; on the other hand, this growth is limited for alkyl ethers with a larger headgroup and small, globular micelles tend to associate in large clusters near the cloud point[121, 122]. However, the mechanisms are not mutually exclusive, and both micelle growth and clustering, albeit to different extents, take place in dilute Ci Ej

SC

solutions near the cloud temperature. With this in mind, it is likely that the cylindrical micelles found at 25 ◦ C and low pH in dilute solutions of C18:1 E9 CH2 COOH behave similarly to C16 E6 or C12 E5 aggregates, growing in length with

NU

increasing temperature and forming an extended network of wormlike micelles at the cloud temperature[121, 123]. Differently, the cloud behavior of C12 E10 CH2 COOH, with a larger headgroup as compared to the hydrophobic chain, is likely to resemble to that of C12 E8 , whose aggregation numbers are constant over an extended range of temperatures

MA

and micelle growth and clustering takes place only close to the cloud point[121, 122].

4. Colloidal building blocks and technical uses

ED

4.1. Building blocks for colloidal supramolecular structures In the previous section, it was shown that polyoxyethylene alkyl ether carboxylic acids can self-assemble into

PT

diverse structures, depending on pH, ionic strength, and length of the hydrophilic and hydrophobic blocks. Moreover, for surfactants with a sufficiently large headgroup, e.g., C12 E10 CH2 COOH, the charge density of the micelle can be varied with minimal effects on its curvature. This is rarely the case, as the electrostatic contribution to the headgroup

AC CE

area is, for most ionic surfactants, the dominant term[41, 124, 125]. Typically, strong variations in size and shape are observed when the ionic strength of the solution is increased, or upon addition of counterions which condense at the micellar interface[126–129]. This effect is even more pronounced when the counterion takes the form of a polyelectrolyte chain[130–132], which triggers the transition from globular to cylindrical aggregates[133–136]. When AECs are admixed to the oppositely charged polycation chitosan (poly d-glucosamine) they co-assembly into hierarchical, complex structures[21, 63]. However, the shape and size of the micellar aggregate is only minimally affected by the presence of chitosan for the investigated C8 E5 CH2 COOH, C12 E10 CH2 COOH, C12 E4.5 CH2 COOH, and C18:1 E9 CH2 COOH surfactants. The presence of a non-negligible steric contribution to the headgroup area Ah therefore allows to systematically investigate the effects of charge density and curvature of the micellar aggregate on the co-assembled structures with oppositely charged polyelectrolytes. The effect of the degree of ionization was investigated in detail on chitosan complexes with C18:1 E9 CH2 COOH[63], while the effect of curvature at constant ionization degree was studied on chitosan complexes with C12 E10 CH2 COOH, C12 E4.5 CH2 COOH, and C18:1 E9 CH2 COOH at pH 4.0[21]. A structural phase diagram for the chitosan-C18:1 E9 CH2 COOH complexes, determined as a function of pH and mixing ratio Z = [surfactant molecules]/[chitosan chargeable units], is shown in Fig. 9. Three main co-assembled structures can be identified: in the bottom-left corner, at low mixing ratio 20

T

0.5

RI P

0.4 0.3 0.2 0.1 0 3.6

3.9

4.2

4.5

4.8

NU

3.3

SC

-

+

[Surf ]/[Chitosan ]

ACCEPTED MANUSCRIPT

5.1

pH

MA

Figure 9: Structural phase diagram depicting the different structures found in mixtures of C18:1 E9 CH2 COOH and chitosan as a function of pH and mixing ratio Z. Experiments were performed at a constant chitosan concentration of 0.3 wt%, in an acetic acid/sodium acetate buffer of 0.2 mol L−1 . The different symbols represent sample composition investigated by small-angle neutron scattering. See Ref. 63 for full details. Adapted with

ED

permission from Ref. 63. Copyright 2017 American Chemical Society.

and low surfactant content, the micelles are randomly distributed in solution, and the chitosan network is preserved; in the central region of the phase diagram, one-dimensional aggregates are formed, where aligned micelles are glued

PT

together by chitosan chains. An appropriate scattering form factor for such object is available[137]. Finally, in the top-right corner, at high degree of ionization and high surfactant content, the aggregates are collapsed into a core-

AC CE

corona suprastructure, with a core formed of densely packed micelles surrounded by a stabilizing chitosan corona. Similar structures are often found in complexes of polyelectrolyte-b-neutral block copolymers and oppositely charged surfactants as a consequence of microphase-separation[138–141], but only rarely when homo-polyelectrolytes are involved[142, 143].

The structural evolution observed in chitosan - C18:1 E9 CH2 COOH is also found in complexes with other alkyl ether carboxylic acids self-assembling into globular micelles, such as C8 E5 CH2 COOH or C12 E10 CH2 COOH[21]. However, when AECs assemble into vesicles, e.g., C12 E4.5 CH2 COOH at pH ∼ 4, the formation of ordered multilayered vesicles with a well-defined interlayer distance has been observed[21]. The average number of layers forming each vesicles depends on the mixing ratio. Such structures have a significant potential as drug delivery agents and for cosmetic applications, as they may be loaded with a hydrophilic cargo in the internal lumen, a charged one in the PE-surfactant layer, and a lipophilic load within the hydrophobic alkylic layer. AEC - chitosan complexes exhibit an exceptional structural variety in solution, but they also exhibit a remarkable high solubility when compared to mixtures of chitosan with strong anionic surfactants, like sodium dodecyl sulfate[144–146]. The high solubility and the mildness of the surfactant combined with the beneficial properties of chitosan make the chitosan - alkyl ether carboxylic acid complexes highly interesting systems for detergency and body care, as demonstrated by diverse patents in the field[147–149]. Similarly, highly soluble catanionic surfactant 21

ACCEPTED MANUSCRIPT

systems with unique properties can be expected from mixtures of AECs with oppositely charged, ethoxylated cationic surfactants.

T

AECs are not only very versatile with respect to their non-covalent interaction properties, but the broad reactivity

RI P

of the carboxylic acid termination allows for simple chemical modifications. For instance, nonionic, thermoresponsive surfactants with a chelating headgroup were prepared by coupling alkyl ether carboxylic acids with an amino acid derivative for the recovery of metal ions[150–152]. Amino acid surfactants are often prepared by coupling a single

SC

or a short sequence of amino acids with fatty acids, resulting in interesting, green surfactants[153–155]. The use of AECs enables the insertion of some ethylene oxide units between the hydrophobic chain and the amino acid surfactant,

NU

conferring upon them a more hydrophilic character and thermoresponsive properties. The same concept applies to covalent coupling with the carboxylic headgroup of fatty acids with polymers[156–159], nanoparticles[160–162], or surfaces[163, 164]. In particular, the covalent conjugation of carboxylic acids to an amine-functionalized surface

MA

via 1-ethyl-3-(3-dimethylaminopropyl)carbodiimide (EDC) activation, allows for a simple and very versatile surface modification[164, 165]. AEC-modified surfaces are potentially attractive as supports for tethered lipid membranes, separated from the solid surface by a soft, hydrated spacer of tunable size. Such bilayers, deposited on a hydrated,

ED

soft cushion, represent excellent model systems to study cell membranes[166–169]. They are also the active elements of highly sensitive biosensors[170, 171].

PT

4.2. Practical applications

In the previous sections, a comprehensive characterization of AECs has been given. The properties of AEC-water

AC CE

solutions depend on a large number of factors, most notably pH and temperature, as well as the size of the hydrophilic and hydrophobic blocks of the surfactant. In addition of being highly tunable, their high solubility in hard water, high salinity, and acidic as well as alkaline conditions, makes them suitable for a wide scope of applications[34, 67]. A comprehensive summary of industrial applications of polyoxyethylene alkyl ether carboxylic acids is given – in german language only – in a review by Czichocki et al.[67] and in a more recent book chapter[34]. The most notable field of application is as household and industrial cleaning agents. This is due to their remarkable mildness, resistance to hard water conditions, high temperatures, the presence of bleaching and oxidizing agents, in addition to their good foaming, wetting, and emulsification properties[34, 36, 172–180]. Moreover, their high compatibility with cationic polyelectrolytes makes them a valuable ingredient for the formulations of shampoos, and skin care, hair conditioning and strengthening component in cosmetic preparations[147–149, 173, 179]. In particular, with respect to alkyl ether sulfates, AECs were reported to cause lower swelling of human stratum corneum and a minor denaturation of the skin proteins[173, 181]. The studies also show that optimal properties are achieved by mixing of alkyl ether carboxylates and alkyl ether sulfates. These mixtures show similar mildness, but better foaming and cleansing properties[173, 181, 182]. Another important field of application of AECs is in enhanced oil recovery[67, 183–188]. In order to extract residual oil from small pores, a sufficiently low interfacial tension between water and oil must be achieved[189]. 22

ACCEPTED MANUSCRIPT

Values as low as 10−2 − 10−3 mN m−1 are reported for C12 E3 CH2 COOH[190]. Even lower values are obtained in mixtures with nonionic surfactants[191]. The reason for the use of AECs in oil recovery, is their ability to form stable

T

emulsions also in the presence of large amounts of monovalent and divalent cations and high temperature[44, 186,

RI P

190], conditions at which many efficient sulfated and nonionic surfactants fail[185]. Improved performance can be obtained using branched alkyl chains, for instance, based on Guerbet alcohols[44, 192].

This multifunctional class of surfactants finds application beyond oil recovery and home and body care products.

SC

For instance, they are used as additives during textile processing[193, 194], as lubricants[195–197], or as dispersant for dyes, inks, and pesticides[198–201]. AECs with low ethylene oxide content were shown to be promising scale

NU

inhibitor agents[202]. As for the previously mentioned applications, their use is justified by their good wettability, emulsification, and solubilization capacity also in conditions where many other surfactants fail. For a more detailed description of the large-scale use of AECs we address the reader to the work of Meijer et al.[34].

MA

In addition to their widespread on an industrial scale, academic research on the use of AECs for a variety of purposes is being carried out. Very promising results have been obtained regarding their use for pollutant recovery[62, 66, 144, 203, 204]. C18:1 E9 CH2 COOH has been shown to be particularly effective for the removal of heavy metal

ED

cations in micellar enhanced ultrafiltration (MEUF)[66, 204]. MEUF is a membrane-based separation technique, in which the wastewater, enriched with surfactants, is filtered through a membrane which enables the separation of the surfactant micelles and associated components from the solution[205, 206]. A high efficiency is obtained

PT

when surfactants with a low cmc and a high affinity towards the pollutant are used. A widely used surfactant for the removal of heavy metal ions is sodium dodecyl sulfate (SDS). However, to decrease the amount of surfactant

AC CE

required it is a common procedure to add a nonionic surfactant to the mixture, with the aim of reducing the cmc[207– 209]. Polyoxyethylene alkyl ether carboxylic acids have three major advantages, when compared to other commonly employed systems: (i) surfactants with a long alkyl chain can be used, exhibiting very low cmc values; (ii) the affinity between surfactant and metal ion is pH dependent and can be exploited for a selective separation of complex metal ion mixtures[204]; and (iii) surfactant – metal pairs can be easily separated and the surfactant recovered for the next cycle by simply decreasing the pH of the solution and increasing the temperature above the cloud point of the surfactant[66]. Metal ions can also be extracted from solution by ion foam flotation[210]. The procedure consists of bubbling a gas, usually air, into a surfactant solution and separating the resultant foam, enriched in surfactant and its associated counterions. Studies on C18:1 E9 CH2 COOH have demonstrated its high efficiency for ion foam flotation applications[62, 203].

5. Conclusions Polyoxyethylene alkyl ether carboxylic acids are highly versatile, multiresponsive surfactants, which combine in a single molecule, peculiarities of pH-responsive fatty acids, and temperature responsive nonionic polyoxyethylene alkyl ethers. This gives rise to a complex, multiresponsive behavior in aqueous solutions. The surfactants properties 23

ACCEPTED MANUSCRIPT

change strongly with pH, as the magnitude of the charge of its headgroup is varied. At low pH, the behavior of AECs is similar to nonionic alkyl ethers, exhibiting similar cmc values and temperature-responsive behavior. With increasing

T

pH, the surfactant headgroup is increasingly charged, thus affecting the properties of the surfactant solutions. The

RI P

extent of the structural changes on pH variation depends on the number of ethylene oxide units in the headgroup. The transition from lamellar aggregates to globular micelles via well-defined vesicles was shown for C12 E4.5 CH2 COOH, while only minor changes in shape and size were observed for C12 E10 CH2 COOH. Soft, globular particles, whose

SC

degree of charge can be systematically varied without affecting their shape and size form a rare and valuable model system for studying charge-driven assembly processes[63]. The surfactants exhibit pKa values around 4, which allows

NU

for the fine-tuning of the surfactants properties in a pH range particularly useful for cosmetic application[64, 65]. The high reactivity of the carboxylic acid termination enables the preparation of surfactants tailored to specific purposes, e.g., the thermoresponsive and chelating amino-acid modified AECs for the recovery of uranyl ions[150,

MA

151]. However, the possibilities offered by the chemical conjugation of AECs, e.g., (i) with single amino acids or peptides, for the preparation of temperature-responsive functional surfactants; (ii) with solid surfaces, to prepare tethered lipid membranes with a variable hydrophilic cushion; (iii) with polymers, for the preparation of hydrophilic

ED

and temperature-responsive drug carriers, have scarcely been explored. AECs, in addition to being very interesting components for colloidal scientists, are widely used in large-scale, industrial applications. Due to the simultaneous presence of the ethylene oxide units and the carboxylic acid termina-

PT

tion, clear solutions of surfactants with long alkyl chains can be obtained in alkaline, as well as acidic solutions which are remarkably stable in hard water and high salinity conditions. Further advantages are given by their mildness, low

AC CE

toxicity, stability to hydrolysis and oxidizing agents and high temperatures. For these reasons, in addition to their low production cost, they find application in different fields, most notably as detergents and in home and body-care products, for enhanced oil recovery and as additives in the textile and metal processing industry. In summary, polyoxyethylene alkyl ether carboxylic acids are highly interesting components in colloid and surfactant science. The complex and highly tunable morphologies of the self-assembled micelles can be exploited to formulate supramolecular aggregates with a degree of diversity rarely found in colloid science. From the perspective of large-scale applications, their mildness and high solubility have opened the door to numerous consumer products and industrial uses. This work provides a comprehensive overview of the physico-chemical properties of this class of surfactants, of their use as building blocks in colloidal chemistry, and of their industrial application, with the aim of facilitating their spread in current research activities.

Acknowledgments Sylvain Prévost, Ralf Schweins, and Pierre Bauduin are gratefully acknowledged for the useful discussions and suggestions. The partnership for soft condensed matter (PSCM) is acknowledged for providing the light scattering apparatus and the tensiometer, the Institut Laue-Langevin (ILL) for the allocation of SANS beamtime. LC acknowledges 24

ACCEPTED MANUSCRIPT

the TU-Berlin and the PSCM at the ILL for postdoctoral funding through a three-year cooperation agreement.

RI P

T

Supporting Information

Additional details and characterization of the AECs used in this work; detailed description of the expressions used for the analysis of the SANS patterns and additional SANS curves; details on the calculations of the surface

SC

electrostatic potential of the surfactant micelles.

References

NU

[1] M. J. Rosen, Surfactants and interfacial phenomena, 3rd ed., John Wiley & Sons, Inc., Hoboken, New Jersey., 2004. URL: http://www.amazon.com/Surfactants-Interfacial-Phenomena-Milton-Rosen/dp/0470541946. doi:10.1002/0471670561. [2] P. Brown, C. P. Butts, J. Eastoe, Stimuli-responsive surfactants, Soft Matter 9 (2013) 2365–2374.

MA

[3] A. Czajka, G. Hazell, J. Eastoe, Surfactants at the Design Limit, Langmuir 31 (2015) 8205–8217. [4] B. Lindman, B. Medronho, G. Karlström, Clouding of nonionic surfactants, Curr. Opin. Colloid Interface Sci. 22 (2016) 23–29. [5] F. Fu, Q. Wang, Removal of heavy metal ions from wastewaters: A review, J. Environ. Manage. 92 (2011) 407–418. [6] C. Stubenrauch, Sugar surfactants - Aggregation, interfacial, and adsorption phenomena, Curr. Opin. Colloid Interface Sci. 6 (2001)

ED

160–170.

[7] R. Sharma, A. Kamal, M. Abdinejad, R. K. Mahajan, H.-B. Kraatz, Advances in the synthesis, molecular architectures and potential applications of gemini surfactants, Adv. Colloid Interface Sci. (2017) doi:10.1016/j.cis.2017.07.032.

PT

[8] K. Holmberg, Natural surfactants, Curr. Opin. Colloid Interface Sci. 6 (2001) 148–159. [9] L. Pérez, A. Pinazo, R. Pons, M. Infante, Gemini surfactants from natural amino acids, Adv. Colloid Interface Sci. 205 (2014) 134–155. 178–188.

AC CE

[10] I. Johansson, M. Svensson, Surfactants based on fatty acids and other natural hydrophobes, Curr. Opin. Colloid Interface Sci. 6 (2001) [11] P. Foley, A. K. Pour, E. S. Beach, J. B. Zimmerman, Derivation and synthesis of renewable surfactants, Chem. Soc. Rev. 41 (2012) 1499–1518.

[12] S. De, S. Malik, A. Ghosh, R. Saha, B. Saha, A review on natural surfactants, RSC Adv. 5 (2015) 65757–65767. [13] H. Fukuda, A. Goto, H. Yoshioka, R. Goto, K. Morigaki, P. Walde, Electron spin resonance study of the pH-induced transformation of micelles to vesicles in an aqueous oleic acid/oleate system, Langmuir 17 (2001) 4223–4231. [14] K. Morigaki, P. Walde, Fatty acid vesicles, Curr. Opin. Colloid Interface Sci. 12 (2007) 75–80. [15] R. Klein, D. Touraud, W. Kunz, Choline carboxylate surfactants: biocompatible and highly soluble in water, Green Chem. 10 (2008) 433. [16] R. Zana, Partial phase behavior and micellar properties of tetrabutylammonium salts of fatty acids: Unusual solubility in water and formation of unexpectedly small micelles, Langmuir 20 (2004) 5666–5668. [17] R. Klein, E. Müller, B. Kraus, G. Brunner, B. Estrine, D. Touraud, J. Heilmann, M. Kellermeier, W. Kunz, Biodegradability and cytotoxicity of choline soaps on human cell lines: effects of chain length and the cation, RSC Adv. 3 (2013) 23347. [18] S. Wolfrum, J. Marcus, D. Touraud, W. Kunz, A renaissance of soaps? - How to make clear and stable solutions at neutral pH and room temperature, Adv. Colloid Interface Sci. 236 (2016) 28–42. [19] T. Yamaguchi, M. Kaneda, K. Kakinuma, Effect of saturated and unsaturated fatty acids on the oxidative metabolism of human neutrophils. The role of calcium ion in the extracellular medium, Biochim. Biophys. Acta - Biomembr. 861 (1986) 440–446. [20] H. Schott, Hydrophile-Lipophile Balance and Cloud Points of Nonionic Surfactants, J. Pharm. Sci. 58 (1969) 1443–1449. [21] L. Chiappisi, S. Prévost, I. Grillo, M. Gradzielski, From Crab Shells to Smart Systems: Chitosan–Alkylethoxy Carboxylate Complexes, Langmuir 30 (2014) 10608–10616.

25

ACCEPTED MANUSCRIPT

[22] M. Hato, K. Shinoda, Krafft points of calcium and sodium dodecylpoly(oxyethylene) sulfates and their mixtures, J. Phys. Chem. 77 (1973) 378–381.

T

[23] M. J. Schwuger, Interfacial and performance properties of sulfated polyoxyethylenated alcohols, ACS Symp. Ser. 253 (1984) 3–26. [24] F. Tokiwa, Solubilization behavior of sodium dodecylpolyoxyethylene sulfates in relation to their polyoxyethylene chain lengths, J. Phys.

RI P

Chem. 72 (1968) 1214–1217.

[25] R. Dong, J. Hao, Complex fluids of poly(oxyethylene) monoalkyl ether nonionic surfactants, Chem. Rev. 110 (2010) 4978–5022. [26] D. J. Mitchell, G. J. T. Tiddy, L. Waring, T. Bostock, M. P. McDonald, Phase behaviour of polyoxyethylene surfactants with water.

SC

Mesophase structures and partial miscibility (cloud points), J. Chem. Soc. Faraday Trans. 1 79 (1983) 975. [27] P. Mukherjee, S. K. Padhan, S. Dash, S. Patel, B. K. Mishra, Clouding behaviour in surfactant systems, Adv. Colloid Interface Sci. 162 (2011) 59–79.

[28] S. Zarate-Muñoz, A. B. Troncoso, E. J. Acosta, The Cloud Point of Alkyl Ethoxylates and Its Prediction with the Hydrophilic-Lipophilic

NU

Difference (HLD) Framework, Langmuir 31 (2015) 12000–12008.

[29] K. Weckström, A. C. Papageorgiou, Lower consolute boundaries of the nonionic surfactant C8E5 in aqueous alkali halide solutions: An approach to reproduce the effects of alkali halides on the cloud-point temperature, J. Colloid Interface Sci. 310 (2007) 151–162.

MA

[30] P. Parekh, U. R. Yerramilli, P. Bahadur, Cloud point and thermodynamic parameters of a non-ionic surfactant heptaoxyethylene dodecyl ether (C12E7) in presence of various organic and inorganic additives, Indian J. Chem. - Sect. A Inorganic, Phys. Theor. Anal. Chem. 52 (2013) 1284–1290.

[31] W. L. Hinze, E. Pramauro, A Critical Review of Surfactant-Mediated Phase Separations (Cloud-Point Extractions): Theory and Applications,

ED

Crit. Rev. Anal. Chem. 24 (1993) 133–177.

[32] E. K. Paleologos, D. L. Giokas, M. I. Karayannis, Micelle-mediated separation and cloud-point extraction, TrAC Trends Anal. Chem. 24 (2005) 426–436.

PT

[33] R. Carabias-Martinez, E. Rodriguez-Gonzalo, B. Moreno-Cordero, J. Perez-Pavon, C. Garcia-Pinto, E. F. Laespada, Surfactant cloud point extraction and preconcentration of organic compounds prior to chromatography and capillary electrophoresis, J. Chromatogr. A 902 (2000) 251–265.

AC CE

[34] H. Meijer, J. K. Smid, Polyether Carboxylates, in: H. Stache (Ed.), Anionic surfactants Org. Chem., Marcel Dekker, Basel, 1996, pp. 314– 361. URL: https://books.google.fr/books?id=B-27_lCJTx8C&pg=PA337&lpg=PA337&dq=Alkyl+ether+carboxylates&source =bl&ots=7P6V_vh-Qh&sig=_FHv-KrEnGLNHdwHmu7zJLRmJOo&hl=fr&sa=X&ved=0ahUKEwiOg9i11tPUAhXFnBoKHcE4Cvs4ChDoAQhiMAs #v=onepage&q&f=false.

[35] R. J. Farn (Ed.), Chemistry and Technology of Surfactants, Blackwell Publishing Ltd, Oxford, 2006. doi:10.1002/9780470988596. arXiv:arXiv:1011.1669v3.

[36] W. W. Schmidt, D. R. Durante, R. Gingell, J. W. Harbell,

Alcohol Ethoxycarboxylates — Mild ,

High-Foaming Surfactants for Personal-Care Products, J. Am. Oil Chem. Soc. 74 (1997) 25–31. [37] H. Fried, Preparation of alkoxyalkanoic acids, 1992. URL: https://www.google.com/patents/US5162579, uS Patent 5,162,579. [38] T. Cripe, Process for making alkyl ethoxy carboxylates, 1993. URL: https://www.google.com/patents/US5233087, uS Patent 5,233,087. [39] T. Nakai, T. Fujioka, Y. Amishige, Method for producing polyoxyalkylene alkyl ether carboxylic acid or salt thereof, 2016. URL: https://www.google.com/patents/US9233904, uS Patent 9,233,904. [40] K. Heidkamp, N. Decker, K. Martens, U. Prüe, K. D. Vorlop, O. Franke, A. Stankowiak, tensidic alcohol to its corresponding carboxylic acid via Au catalysts,

Oxidation of a

Eur. J. Lipid Sci. Technol. 112

(2010) 51–57. [41] D. A. McKenzie, Nonionic surfactants, J. Am. Oil Chem. Soc. 55 (1978) 93–97. [42] N. M. van Os, Nonionic surfactants :

organic chemistry, Marcel Dekker, New York, 1998. URL:

26

ACCEPTED MANUSCRIPT

https://books.google.fr/books?id=YoZ6CjYNLoQC&printsec=frontcover&dq=Nonionic+surfactants+synthesis&hl=en&sa =X&ved=0ahUKEwj_2PqG7PbUAhXJtBQKHaSzDSUQ6AEIJDAA#v=onepage&q=Nonionic surfactants synthesis&f=false.

T

[43] A. J. O’Lenick, Guerbet chemistry, J. Surfactants Deterg. 4 (2001) 311–315. [44] J. Lu, P. J. Liyanage, S. Solairaj, S. Adkins, G. P. Arachchilage, D. H. Kim, C. Britton, New surfactant developments for chemical enhanced oil recovery,

RI P

U. Weerasooriya, G. A. Pope, Sci. Eng. 120 (2014) 94–101.

[45] Q. Zhang, Y. Li, Y. Song, H. Fu, J. Li, Z. Wang,

J. Pet.

Properties of vesicles formation of single-chain

SC

branched carboxylate anionic surfactant in aqueous solutions, J. Mol. Liq. 243 (2017) 431–438. [46] D. Leinweber, U. Dahlmann, R. Kupfer, Process for the solvent-free preparation of ethercarboxylic acids having a low residual salt content, 2007. URL: https://www.google.com/patents/US7208118, uS Patent 7,208,118.

NU

[47] P. Klug, R. Kupfer, I. Wimmer, R. Winter, Process for the preparation of ether carboxylic acids with low residual alcohol content, 2001. URL: https://www.google.com/patents/US6326514, uS Patent 6,326,514. [48] A. Matsunaga, K. Ohmae, Process for production of ether carboxylates, 2010. URL: https://encrypted.google.

[49] T. Mallat, A. Baiker,

MA

com/patents/US20100081716, uS Patent App. 12/597,019.

Oxidation of alcohols with molecular oxygen on platinum metal catalysts in aqueous

solutions, Catal. Today 19 (1994) 247–283. [50] S. Zlatanos, K. Laskaridis, A. Sagredos,

Oxidation of ethoxylated fatty alcohols to alkylpolyglycol

ED

carboxylic acids using noble metals as catalysts, Grasas y Aceites 60 (2009) 360–366. [51] J. Sha, E.-J. J. Zheng, W.-J. J. Zhou, A. Liebens, M. Pera-Titus,

Selective oxidation of fatty alcohol

ethoxylates with H2O2 over Au catalysts for the synthesis of alkyl ether carboxylic acids in alkaline

PT

solution, J. Catal. 337 (2016) 199–207.

[52] L. Tran-Vinh, G. Göbel, F. Carduck, S. Herbst, Process for the preparation of ethercarboxylic acids, 1989. URL: http://www.google.dj/patents/EP0304763A1?cl=en, eP Patent App. EP19,880,113,245.

AC CE

[53] H. Fiege, K. Wedemeyer, Process for preparing alkoxyacetic acids, 1983. URL: http://www.google.com.na/ patents/DE3135946A1?cl=en, dE Patent App. DE19,813,135,946. [54] L. Slaugh, C. Willis, A process for the preparation of alkoxyalkanoic acids, 1983. URL: https://www.google.com/patents/EP0039111B1?cl=en, eP Patent 0,039,111. [55] A. Stankowiak, O. Franke, U. Pruesse, N. Decker, K. Vorlop, Method for producing alkyl polyglycol carboxylic acids and polyglycol dicarboxylic acids, 2010. URL: http://www.google.com.na/patents/US20100056735, uS Patent App. 12/595,199.

[56] K. Heidkamp, N. Decker, U. Prüße,

Catalyst development for the selective oxidation of ethoxylates to

their corresponding ether carboxylic acids, Catal. Commun. 40 (2013) 88–92. [57] K. Heidkamp, M. Aytemir, K.-D. Vorlop, U. Prüße,

Ceria supported gold–platinum catalysts for the

selective oxidation of alkyl ethoxylates, Catal. Sci. Technol. 4 (2013) 2984–2992. [58] Ether Carboxylates Market:

Global Industry Analysis and Forecast to 2015 to 2021,

http://www.persistencemarketresearch.com/market-research/ether-carboxylates-market.asp, Accessed:

July,

10, 2017. [59] K. Ogino, N. Tsubaki, M. Abe,

Solution properties of mixed surfactant system (II): Electric properties

of anionic-nonionic surfactants in aqueous solutions, J. Colloid Interface Sci. 98 (1984) 78–83. [60] W. Binana-Limbelé, R. Zana, E. Platone,

Micellar properties of ethoxylated sodium alkylcarboxylates,

J.

Colloid Interface Sci. 124 (1988) 647–651. [61] T. Sakai, R. Ikoshi, N. Toshida, M. Kagaya,

Thermodynamically stable vesicle formation and

27

ACCEPTED MANUSCRIPT

vesicle-to-micelle transition of single-tailed anionic surfactant in water,

J. Phys. Chem. B 117 (2013)

5081–5089. Specific salt and pH effects on foam film of a pH sensitive

T

[62] C. Micheau, P. Bauduin, O. Diat, S. Faure, surfactant, Langmuir 29 (2013) 8472–8481.

Chitosan/alkylethoxy carboxylates:

RI P

[63] L. Chiappisi, S. Prévost, I. Grillo, M. Gradzielski, variety of structures, Langmuir 30 (2014) 1778–1787.

[64] N. Vlachy, C. Merle, D. Touraud, J. Schmidt, Y. Talmon, J. Heilmann, W. Kunz,

cytotoxicity studies, Langmuir 24 (2008) 9983–9988.

Spontaneous formation

Effect of pH and aging and

SC

of bilayers and vesicles in mixtures of single-chain alkyl carboxylates:

A surprising

[65] A. Renoncourt, P. Bauduin, E. Nicholl, D. Touraud, J. M. Verbavatz, M. Dubois, M. Drechsler, W. Kunz,

Spontaneous vesicle formation of an industrial single-chain surfactant at acidic pH and at

NU

room-temperature, ChemPhysChem 7 (2006) 1892–1896.

[66] M. Schwarze, L. Chiappisi, S. Prévost, M. Gradzielski,

Oleylethoxycarboxylate – An efficient surfactant

Interface Sci. 421 (2014) 184–190.

MA

for copper extraction and surfactant recycling via micellar enhanced ultrafiltration,

[67] G. Czichocki, H.-R. Holzbauer, C. Martens,

Eigenschaften, Anwendungsgebiete und physiko-chemische

Untersuchungen von Polyethercarbonsäuren und ihren Salzen, 66–73.

ED

[68] L. Chiappisi, S. D. Leach, M. Gradzielski,

J. Colloid

Fett Wiss. Technol. Sci. Technol. 94 (1992)

Precipitating polyelectrolyte–surfactant systems by admixing

a nonionic surfactant – a case of cononsurfactancy, Soft Matter 13 (2017) 4988–4996. [69] X. Yue, X. Fan, Q. Li, X. Chen, C. Wang,

Aggregation behaviors of alkyl ether carboxylate surfactants in

[70] H. C. Kim, J. D. Kim,

PT

water, J. Mol. Liq. 227 (2016) 161–167.

The polydispersity effect of distributed oxyethylene chains on the cloud points of

nonionic surfactants, J. Colloid Interface Sci. 352 (2010) 444–448.

AC CE

[71] E. H. Crook, D. B. Fordyce, G. F. Trebbi,

Molecular weight distribution of nonionic surfactants. I.

Surface and interfacial tension of normal distribution and homogeneous p, t-octylphenoxyethoxyethanols (OPE’S), J. Phys. Chem. 67 (1963) 1987–1994. [72] D. Hui, M. Nawaz, D. P. Morris, M. Edwards, B. R. Saunders,

Study of pH-triggered heteroaggregation and

gel formation within mixed dispersions, J. Colloid Interface Sci. 324 (2008) 110–117. [73] L. Sun, R. M. Crooks,

Interactions between dendrimers and charged probe molecules. 1. Theoretical

methods for simulating proton and metal ion binding to symmetric polydentate ligands,

J. Phys. Chem.

B 106 (2002) 5864–5872.

[74] Q. R. Huang, P. L. Dubin, C. N. Moorefield, G. R. Newkome, Spheres:

Counterion Binding on Charged

Effect of pH and Ionic Strength on the Mobility of Carboxyl-Terminated Dendrimers,

J. Phys.

Chem. B 104 (2000) 898–904. [75] A. Katchalsky, J. Gillis,

Theory of the potentiometric titration of polymeric acids,

Recl. des Trav.

Chim. des Pays-Bas 68 (1949) 879–897. [76] R. Arnold, J. T. G. Overbeek, The dissociation and specific viscosity of polymethacrylic acid, Recl. des Trav. Chim. des Pays-Bas 69 (2010) 192–206. [77] F. Uhlík, P. Košovan, Z. Limpouchová, K. Procházka, O. V. Borisov, F. Leermakers,

Modeling of Ionization

and Conformations of Starlike Weak Polyelectrolytes, Macromolecules 47 (2014) 4004–4016. [78] J. Lützenkirchen,

The Constant Capacitance Model and Variable Ionic Strength:

Applications and Applicability, J. Colloid Interface Sci. 217 (1999) 8–18.

28

An Evaluation of Possible

ACCEPTED MANUSCRIPT

[79] J. Lützenkirchen, J. van Male, F. Leermakers, S. Sjberg,

Comparison of Various Models to Describe the

Charge−pH Dependence of Poly(acrylic acid), J. Chem. Eng. Data 56 (2011) 1602–1612.

T

[80] P. Linse, PB, Version 2.4.1, 2015. URL: http://www.polyelectrolytes2002.fkem1.lu.se/PB/PB.htm. [81] S. Lin, Y. Lin, E. Chen, C. Hsu, C. Kwan,

A Study of the Equilibrium Surface Tension and the Critical

RI P

Micelle Concentration of Mixed Surfactant Solutions, Langmuir 15 (1999) 4370–4376. [82] I. Svanedal, G. Persson, M. Norgren, H. Edlund,

Anomalies in solution behavior of an alkyl

aminopolycarboxylic chelating surfactant, Langmuir 29 (2013) 13708–13716.

Adsorption of single and mixed ionic surfactants at fluid

SC

[83] V. B. Fainerman, E. H. Lucassen-Reynders,

interfaces, Adv. Colloid Interface Sci. 96 (2002) 295–323.

[84] J. Spink, Ionization of monolayers of fatty aicd from C14 to C18, J. Colloid Sci. 18 (1963) 512–525. [85] M. Tomoaia-Cotisel, J. Zsako, A. Mocanu, M. Lupea, E. Chifu,

Insoluble mixed monolayers. III. The J. Colloid Interface Sci.

NU

ionization characteristics of some fatty acids at the air/water interface, 117 (1987) 464–476.

[86] S. B. Sulthana, P. V. C. Rao, S. G. T. Bhat, T. Y. Nakano, G. Sugihara, A. K. Rakshit,

MEGA-10, Langmuir 16 (2000) 980–987. [87] P. H. Elworthy, C. B. Macfarlane,

Solution

polyoxyethylene (10) alkyl ether [CnE10] and

Surface activity of a series of synthetic non-ionic detergents,

Pharm. Pharmacol. 14 (1962) 100T–102T.

J.

The determination of critical micelle concentrations of nonionic surfactants

ED

[88] K. Deguchi, K. Meguro,

MA

properties of nonionic surfactants and their mixtures:

by charge-transfer solubilization of 7, 7, 8, 8 tetracyanoquinodimethane, (1972) 596–600.

J. Colloid Interface Sci. 38

The Formation of Surface

PT

[89] H. Xu, J. Penfold, R. K. Thomas, J. T. Petkov, I. Tucker, J. P. R. Webster,

Multilayers at the Air–Water Interface from Sodium Polyethylene Glycol Monoalkyl Ether Sulfate/AlCl 3 Solutions:

Critical Micellization Concentration of Surfactants in Aqueous Solution and Free Energy of

AC CE

[90] R. Zana,

The Role of the Size of the Polyethylene Oxide Group, Langmuir 29 (2013) 11656–11666.

Micellization, Langmuir 12 (1996) 1208–1211. [91] S. Prévost, M. Gradzielski,

SANS investigation of the microstructures in catanionic mixtures of SDS/DTAC

and the effect of various added salts., J. Colloid Interface Sci. 337 (2009) 472–84. [92] M. Lopez-Lopez, F. Montilla, M. Olivares, J. A. Lebron, M. L. Moya, P. Lopez-Cornejo,

Interaction

between monomers of two surfactants derived from the [Ru(2,2’-bpy)(3)](2+) complex and α, β and γ-cyclodextrins:

formation of [2]- and [3]-pseudorotaxanes., Dalton Trans. 42 (2013) 6171–81.

[93] S. Bae, K. Haage, K. Wantke, H. Motschmann,

On the Factor in Gibbs Equation for Ionic Surfactants,

J.

Phys. Chem. B 103 (1999) 1045–1050. [94] J. V. Joshi, V. K. Aswal, P. S. Goyal,

Combined SANS and SAXS studies on alkali metal dodecyl sulphate

micelles, J. Phys. Condens. Matter 19 (2007) 196219. [95] R. Zana,

Ionization of cationic micelles:

Effect of the detergent structure,

J. Colloid Interface Sci.

78 (1980) 330–337. [96] L. F. W. Vleugels, J. Pollet, R. Tuinier, complexes:

Polycation-sodium lauryl ether sulfate-type surfactant

influence of ethylene oxide length, J. Phys. Chem. B 119 (2015) 6338–6347.

[97] G. Gunnarsson, B. Joensson, H. Wennerstroem, J. J. Lagowski, A. Press, G. Manantov, P. Press, W. J. Green, D. E. Mather, P. G. Zambonin, R. Grimes, M. Watson, E. Drain, S. Frank, M. W. Evans, G. Gunnarsson, B. Jonsson, H. Wennerstroem, B. Joensson, electrostatic approach, J. Phys. Chem. 84 (1980) 3114–3121.

29

Surfactant association into micelles. An

ACCEPTED MANUSCRIPT

[98] E. Vikingstad,

The mean activity and the activities of the separate ions of sodium decanoate above

and below the CMC determined by a surfactant selective silver/silver decanoate electrode,

J. Colloid

T

Interface Sci. 72 (1979) 68–74. Modelling of small-angle scattering data from colloids and polymer systems,

P. Lindner, T. Zemb (Eds.), Neutrons , X-rays and Light :

in:

Scattering Methods Applied to Soft Condensed

RI P

[99] J. S. Pedersen,

Matter, Elsevier Science Bv, 2002, pp. 391–420.

[100] L. Baba-Ahmed, M. Benmouna, M. J. Grimson, Elastic Scattering from Charged Colloidal Dispersions, Phys.

SC

Chem. Liq. 16 (1987) 235–238. [101] T. Sato, Y. Saito, I. Anazawa,

Micellar structure and micellar inner polarity of octaethylene glycol

n-alkyl ethers, J. Chem. Soc. Faraday Trans. 1 Phys. Chem. Condens. Phases 84 (1988) 275. [102] J. R. Bartlett, R. P. Cooney,

Raman Spectroscopic Studies of Poly (oxyethylene) Chain Conformations in

NU

Non-ionic Surfactants, J. Chem. Soc. Faraday Trans. 82 (1986) 597–605. [103] C. Tanford, Y. Nozaki, M. F. Rohde,

Size and shape of globular micelles formed in aqueous solution by

n-alkyl polyoxyethylene ethers, J. Phys. Chem. 81 (1977) 1555–1560.

studies of nonionic surfactant: [105] N. Zoeller, D. Blankschtein,

MA

[104] K. S. Sharma, J. V. Joshi, V. K. Aswal, P. S. Goyal, a. K. Rakshit,

Small-angle neutron scattering

Effect of sugars, Pramana 63 (2004) 297–302. Experimental Determination of Micelle Shape and Size in Aqueous Solutions

of Dodecyl Ethoxy Sulfates, Langmuir 14 (1998) 7155–7165. Phase Behavior of Aqueous Mixtures of Some Polyethylene Glycol Decyl Ethers Revealed

ED

[106] Y. Nibu, T. Inoue,

by DSC and FT-IR Measurements., J. Colloid Interface Sci. 205 (1998) 305–315. [107] H. Kunieda, K. Shigeta, K. Ozawa, M. Suzuki,

Self-organizing structures in poly(oxyethylene) oleyl

PT

ether-water system, J. Phys. Chem. B 101 (1997) 7952–7957. [108] K.-L. Huang, K. Sigeta, H. Kunieda,

Phase behavior of polyoxyethylene dodecyl ether-water systems,

Prog. Colloid Polym. Sci. 110 (1998) 171–174.

AC CE

[109] M. W. Sulek, T. Wasilewski, K. J. Kurzydłowski,

The effect of concentration on lubricating properties

of aqueous solutions of sodium lauryl sulfate and ethoxylated sodium lauryl sulfate,

Tribol. Lett. 40

(2010) 337–345.

[110] P. Kékicheff, C. Grabielle-Madelmont, M. Ollivon,

Phase diagram of sodium dodecyl sulfate-water system.

1. A calorimetric study, J. Colloid Interface Sci. 131 (1989) 112–132. [111] E. Jahns, H. Finkelmann,

Lyotropic liquid crystalline phase behavior of a polymeric amphiphile

polymenzed via their hydrophilic ends, Colloids and Surfaces 265 (1987) 304–311. [112] J. Conroy, C. Hall, C. Leng, K. Rendall, G. Tiddy, J. Walsh, G. Lindblom,

Nonionic surfactant phase

behavior. The effect of CH3 capping of the terminal OH. Accurate measurements of cloud curves,

Prog.

Colloid Polym. Sci. 82 (1990) 253–262. [113] M. Fujiwara, M. Miyake, I. Hama,

Phase behavior of methoxypolyoxyethylene dodecanoate as compared to

polyoxyethylene dodecyl ether and polyoxyethylene methyl dodecyl ether,

Colloid Polym. Sci. 272 (1994)

797–802. [114] T. Sato, T. Akahane, K. Amano, R. Hyodo, K. Yanase, T. Ogura,

Scattering and Spectroscopic Study on

the Hydration and Phase Behavior of Aqueous Alcohol Ethoxylate and Methyl Ester Ethoxylate:

Effects of

Terminal Groups in Hydrophilic Chains, J. Phys. Chem. B 120 (2016) 5444–5454. [115] K. S. Sharma, S. R. Patil, A. K. Rakshit,

Study of the cloud point of C12En nonionic surfactants:

Effect of additives, Colloids Surfaces A Physicochem. Eng. Asp. 219 (2003) 67–74. [116] A. E. Fernandez, Z. S. Ferrera, J. J. S. Rodriguez,

30

Determination of polychlorinated biphenyls by

ACCEPTED MANUSCRIPT

liquid chromatography following cloud-point extraction, Anal. Chim. Acta 358 (1998) 145–155. [117] Z. S. Ferrera, C. P. Sanz, C. M. Santana, J. J. S. Rodríguez,

The use of micellar systems in the

Chem. 23 (2004) 469–479. I. Experimental Facts:

Water-Nonionic Surfactant Systems, and the Effect of Additives,

RI P

[118] R. Strey,

TrAC - Trends Anal.

T

extraction and pre-concentration of organic pollutants in environmental samples,

Berichte der Bunsengesellschaft für Phys. Chemie 100 (1996) 182–189. [119] A. Bernheim-Groswasser, E. Wachtel, Y. Talmon,

Micellar Growth, Network Formation, and Criticality in

SC

Aqueous Solutions of the Nonionic Surfactant C12E5, Langmuir 16 (2000) 4131–4140. [120] B. Lindman, H. Wennerstroem,

Nonionic micelles grow with increasing temperature,

(1991) 6053–6054. [121] P. G. Nilsson, H. Wennerstroem, B. Lindman,

J. Phys. Chem. 95

Structure of micellar solutions of nonionic surfactants.

ethers, J. Phys. Chem. 87 (1983) 1377–1385. [122] D. Danino, Y. Talmon, R. Zana,

NU

Nuclear magnetic resonance self-diffusion and proton relaxation studies of poly(ethylene oxide) alkyl

Aggregation and Microstructure in Aqueous Solutions of the Nonionic

[123] Z. Lin, L. E. Scriven, H. T. Davis,

MA

Surfactant C12E8, J. Colloid Interface Sci. 186 (1997) 170–179. Cryogenic electron microscopy of rodlike or wormlike micelles in

aqueous solutions of nonionic surfactant hexaethylene glycol monohexadecyl ether, 2200–2205.

Molecular packing parameter and surfactant self-assembly:

The neglected role of the

ED

[124] R. Nagarajan,

Langmuir 8 (1992)

a predictive molecular thermodynamic

[126] V. K. Aswal, P. S. Goyal, P. Thiyagarajan,

Small-angle neutron-scattering and viscosity studies of

surfactant tail, Langmuir 18 (2002) 31–38. [125] R. Nagarajan, E. Ruckenstein,

Theory of surfactant self-assembly:

PT

approach, Langmuir 7 (1991) 2934–2969.

CTAB/NaSal viscoelastic micellar solutions, J. Phys. Chem. B 102 (1998) 2469–2473. Microstructural changes in SDS micelles induced by

AC CE

[127] P. A. Hassan, S. R. Raghavan, E. W. Kaler, hydrotropic salt, Langmuir 18 (2002) 2543–2548. [128] V. K. Aswal, P. S. Goyal,

Counterions in the growth of ionic micelles in aqueous electrolyte solutions:

A small-angle neutron scattering study, Phys. Rev. E 61 (2000) 2947–2953. [129] H. Hoffmann, H. Rehage, G. Platz, W. Schorr, H. Thurn, W. Ulbricht,

Investigations on a detergent

system with rodlike micelles, Colloid Polym. Sci. 260 (1982) 1042–1056. [130] I. Hoffmann, P. Heunemann, S. Prévost, R. Schweins, N. J. Wagner, M. Gradzielski,

Self-aggregation

of mixtures of oppositely charged polyelectrolytes and surfactants studied by rheology, dynamic light scattering and small-angle neutron scattering, Langmuir 27 (2011) 4386–4396. [131] K. Nakamura, T. Shikata, N. Takahashi, T. Kanaya,

Highly extended conformation of polyelectrolytes

incorporated into hybrid threadlike micelles studied by small angle neutron scattering,

J. Am. Chem.

Soc. 127 (2005) 4570–4571. [132] J. Merta, V. M. Garamus, R. Willumeit, P. Stenius,

Structure of complexes formed by PDADMAC and sodium

palmitate, Langmuir 18 (2002) 7272–7278. [133] I. Hoffmann, B. Farago, R. Schweins, P. Falus, M. Sharp, S. Prévost, M. Gradzielski, origins of high viscosities in some polyelectrolyte-surfactant mixtures,

On the mesoscopic

J. Chem. Phys. 143 (2015)

74902. [134] I. Hoffmann, B. Farago, R. Schweins, P. Falus, M. Sharp, M. Gradzielski,

Structure and dynamics of

polyelectrolytes in viscous polyelectrolyte-surfactant complexes at the mesoscale,

31

Europhys. Lett. 104

ACCEPTED MANUSCRIPT

(2013) 28001. [135] I. S. Chronakis, P. Alexandridis,

Rheological properties of oppositely charged Effect of polymer molecular weight and surfactant architecture,

T

polyelectrolyte-surfactant mixtures: Macromolecules 34 (2001) 5005–5018.

RI P

[136] N. Beheshti, G. T. M. Nguyen, A. L. Kjøniksen, K. D. Knudsen, B. Nyström,

Structure and dynamics of

aqueous mixtures of an anionic cellulose derivative and anionic or cationic surfactants, Surfaces A Physicochem. Eng. Asp. 279 (2006) 40–49.

Form factor of cylindrical superstructures composed of

SC

[137] L. Chiappisi, S. Prévost, M. Gradzielski,

Colloids

globular particles, J. Appl. Crystallogr. 47 (2014) 827–834. [138] J.-F. Berret, J. Oberdisse,

Electrostatic self-assembly in polyelectrolyte-neutral block copolymers and

oppositely charged surfactant solutions, Phys. B Condens. Matter 350 (2004) 204–206.

Block Copolymers and Surfactants:

NU

[139] J.-F. Berret, P. Hervé, O. Aguerre-Chariol, J. Oberdisse,

Colloidal Complexes Obtained from Charged

A Comparison between Small-Angle Neutron Scattering, Cryo-TEM, and

Simulations, J. Phys. Chem. B 107 (2003) 8111–8118.

MA

[140] M. Uchman, M. Štěpánek, S. Prévost, B. Angelov, J. Bednár, M.-S. Appavou, M. Gradzielski, K. Procházka, Coassembly of Poly(ethylene oxide)- block -poly(methacrylic acid) and N -Dodecylpyridinium Chloride in Aqueous Solutions Leading to Ordered Micellar Assemblies within Copolymer Aggregates, (2012) 6471–6480.

Macromolecules 45

C. Mantzaridis, S. Pispas,

ED

[141] M. Štěpánek, J. Škvarla, M. Uchman, K. Procházka, B. Angelov, L. Kováčik, V. M. Garamus, Wormlike core-shell nanoparticles formed by co-assembly of double hydrophilic

block polyelectrolyte with oppositely charged fluorosurfactant, Soft Matter 8 (2012) 9412.

PT

[142] S. Trabelsi, S. Guillot, E. Raspaud, M. Delsanti, D. Langevin, F. Boué,

New Nano- and Microparticles

with a Liquid-Crystal-Like Interior, Adv. Mater. 18 (2006) 2403–2406. [143] I. Schmidt, F. Cousin, C. Huchon, F. Boué, M. a. V. Axelos,

Spatial structure and composition of

AC CE

polysaccharide-protein complexes from small angle neutron scattering,

Biomacromolecules 10 (2009)

1346–1357.

[144] L. Chiappisi, M. Gradzielski,

Co-assembly in chitosan-surfactant mixtures:

Thermodynamics, structures,

interfacial properties and applications, Adv. Colloid Interface Sci. 220 (2015) 92–107. [145] S. Peretz, M. Florea-Spiroiu, D.-F. Anghel, C. Munteanu, D. Angelescu, C. Stoian, G. Zgherea, Chitosan-sodium lauryl ether sulfate particles and their use for adsorption of Cu(II) ions,

J. Appl.

Polym. Sci. 131 (2014) 40059.

[146] M. Olteanu, I. Mandru, M. Dudau, S. Peretz, O. Cinteza,

The aqueous liquid/liquid interphases formed by

chitosan-anionic surfactant complexes, Prog. Colloid Polym. Sci. 122 (2003) 87–94. [147] E. Konrad, G. Lang, Acid tenside-anion solution of chitosan and its use in cosmetic preparations, 1986. URL: https://www.google.com/patents/US4603048, uS Patent 4,603,048. [148] S. Onitsuka, H. Mohring, Hair tinting shampoo which contains a direct dye and an anionic surfactant, 1997. URL: https://www.google.com/patents/US5635461, uS Patent 5,635,461. [149] G. Lang, T. Clausen, Hair fixing composition in the form of an aqueous solution, foam or gel, 1997. URL: https://encrypted.google.com/patents/US5690921, uS Patent 5,690,921. [150] H. Coulombeau, F. Testard, T. Zemb, C. Larpent,

Effect of recognized and unrecognized salt on the

self-assembly of new thermosensitive metal-chelating surfactants., Langmuir 20 (2004) 4840–50. [151] C. Larpent, S. Prévost, L. Berthon, T. Zemb, F. Testard,

Nonionic metal-chelating surfactants mediated

solvent-free thermo-induced separation of uranyl, New J. Chem. 31 (2007) 1424–1428.

32

ACCEPTED MANUSCRIPT

[152] C. Madic, L. Berthon, N. Zorz, H. Coulombeau, F. Testard, T. Zemb, S. Gibanel, C. Larpent, K. Baczko, Heat-sensitive compounds having a cloud point which can be used as extractants for the separation of

T

metals in aqueous solution, 2010. URL: https://encrypted.google.com/patents/EP1425263B1?cl=en, eP Patent 1,425,263.

RI P

[153] M. C. Morán, A. Pinazo, L. Pérez, P. Clapés, M. Angelet, M. T. García, M. P. Vinardell, M. R. Infante, “Green” amino acid-based surfactants, Green Chem. 6 (2004) 233–240.

[154] D. W. P. M. Löwik, J. C. M. van Hest, Peptide based amphiphiles., Chem. Soc. Rev. 33 (2004) 234–245. Amino acid-based surfactants - Do they deserve more attention?,

SC

[155] R. Bordes, K. Holmberg,

Interface Sci. 222 (2015) 79–91.

[156] H. Almasi, B. Ghanbarzadeh, J. Dehghannya, A. A. Entezami, A. K. Asl, fatty acid modified cellulose nanofibers/poly(lactic acid):

Novel nanocomposites based on

Morphological and physical properties,

Food

NU

Packag. Shelf Life 5 (2015) 21–31.

Adv. Colloid

[157] H. Namazi, A. Dadkhah,

Convenient method for preparation of hydrophobically modified starch

nanocrystals with using fatty acids, Carbohydr. Polym. 79 (2010) 731–737. Hydrophobically modified starch using long-chain fatty acids for

MA

[158] H. Namazi, F. Fathi, A. Dadkhah,

preparation of nanosized starch particles, Sci. Iran. 18 (2011) 439–445. [159] J. Zhang, X. G. Chen, Y. Y. Li, C. S. Liu,

Self-assembled nanoparticles based on hydrophobically

modified chitosan as carriers for doxorubicin, Nanomedicine Nanotechnology, Biol. Med. 3 (2007) 258–265.

ED

[160] R. Ramos-González, L. García-Cerda, M. Quevedo-López,

Study of the surface modification with oleic acid

of nanosized HfO2 synthesized by the polymerized complex derived sol–gel method, (2012) 6034–6039.

Creation of High-Density and Low-Defect Single-Layer

PT

[161] A. Fujimori, K. Ohmura, N. Honda, K. Kakizaki,

Appl. Surf. Sci. 258

Film of Magnetic Nanoparticles by the Method of Interfacial Molecular Films, 3254–3261.

Surface modification of magnetic nanoparticles capped by oleic acids:

AC CE

[162] S.-Y. Lee, M. T. Harris,

Characterization and colloidal stability in polar solvents, 401–408.

Langmuir 31 (2015)

[163] S. P. Pujari, L. Scheres, A. T. M. Marcelis, H. Zuilhof,

J. Colloid Interface Sci. 293 (2006)

Covalent surface modification of oxide

surfaces, Angew. Chemie - Int. Ed. 53 (2014) 6322–6356. [164] T. Yoshioka, K. Tsuru, S. Hayakawa, A. Osaka,

Preparation of alginic acid layers on stainless-steel

substrates for biomedical applications, Biomaterials 24 (2003) 2889–2894. [165] Z. Zhang, J. Buitenhuis, A. Cukkemane, M. Brocker, M. Bott, J. K. G. Dhont,

Charge Reversal of the

Rodlike Colloidal fd Virus through Surface Chemical Modification, Langmuir 26 (2010) 10593–10599. [166] V. Stanjek, M. Miksch, P. Lueer, U. Matern, W. Boland,

Biosynthesis of Psoralen:

Mechanism of a

Cytochrome P450 Catalyzed Oxidative Bond Cleavage, Angew. Chemie - Int. Ed. 38 (1999) 400–402. [167] C. Rossi, J. Chopineau,

Biomimetic tethered lipid membranes designed for membrane-protein interaction

studies, Eur. Biophys. J. 36 (2007) 955–965. [168] C. Daniel, K. E. Sohn, T. E. Mates, E. J. Kramer, J. O. Rädler, E. Sackmann, B. Nickel, L. Andruzzi, Structural characterization of an elevated lipid bilayer obtained by stepwise functionalization of a self-assembled alkenyl silane film., Biointerphases 2 (2007) 109–118. [169] E. Rascol, J.-M. Devoisselle, j. c. Chopineau,

The relevance of membrane models to understand

nanoparticles-cell membrane interactions, Nanoscale 8 (2016) 4780–4798. [170] I. Köper,

Insulating tethered bilayer lipid membranes to study membrane proteins.,

33

Mol. Biosyst. 3

ACCEPTED MANUSCRIPT

(2007) 651–657. [171] V. Atanasov, N. Knorr, R. S. Duran, S. Ingebrandt, A. Offenhäusser, W. Knoll, I. Köper, a functional tethered lipid bilayer membrane on silicon oxide surfaces.,

Biophys. J. 89 (2005)

T

chip:

Membrane on a

1780–1788.

[173] T. Ozawa, K. Endo, T. Masui, M. Miyaki, K. Matsuo, S. Yamada,

RI P

[172] C. Edser, What does the future hold for anionic surfactants?, Focus Surfactants 2006 (2006) 1–2. Advantage of Sodium Polyoxyethylene

Lauryl Ether Carboxylate as a Mild Cleansing Agent, J. Surfactants Deterg. 19 (2016) 785–794.

SC

[174] T. Cripe, K. Ofosu-Asante, Alkaline light duty dishwashing detergent composition containing an alkyl ethoxy carboxylate surfactant, magnesium ions, chelator and buffer, 1994. URL: https://www.google.com/patents/US5376310, uS Patent 5,376,310.

[175] R. Wise, T. Cripe, Light-duty liquid or gel dishwashing detergent composition containing an alkyl

NU

ethoxy carboxylate surfactant, 1990. URL: https://www.google.com/patents/EP0399752A2?cl=en, eP Patent App. EP19,900,305,469.

[176] D. Balzer, Detergent compositions containing a carboxymethylated ethoxylate with elevated viscosity,

MA

1992. URL: https://www.google.com/patents/US5098596, uS Patent 5,098,596. [177] D. Balzer, Detergent compositions with elevated viscosity, 1999. URL: https://www.google.com/patents/ CA2010787C?cl=en, cA Patent 2,010,787.

[178] H. Denzer, H. Abe, M. Pytlik, R. Jansen, A. Buhmann, Compositions comprising hydrophobic silicone oils

ED

and alkyl ether carboxylates, 2007. URL: https://www.google.com/patents/US7285263, uS Patent 7,285,263. [179] N. Subirats, A. G. Pilar Castan, Pers. Care Today (2010) 33–38.

Alkyl ether carboxylates Optimizing coacervate formation,

An Alkyl Ether Carboxylate and

PT

[180] J. Sampuntharat, P. Boonkoom, T. Y, T. N, H. M, M. M, T. K, M. K, N. T,

Househ.

Alkyl Carboxylate Formulated Cleanser Decreases Facial Sebum and Inflammatory Acne Without Inducing Dry Xerotic Skin in Thai Females, Int. J. Clin. Dermatology Res. 3 (2015) 60–64.

AC CE

[181] Y. Takagi, M. Shimizu, Y. Morokuma, M. Miyaki, A. Kiba, K. Matsuo, K. Isoda, H. Mizutani, for a mild body cleanser:

A new formula

Sodium laureth sulphate supplemented with sodium laureth carboxylate and

lauryl glucoside, Int. J. Cosmet. Sci. 36 (2014) 305–311. [182] S. Fukui, T. Morikawa, M. Hirahara, Y. Terada, M. Shimizu, K. Takeuchi, Y. Takagi,

A mild hand

cleanser, alkyl ether sulphate supplemented with alkyl ether carboxylic acid and alkyl glucoside, improves eczema on the hand and prevents the growth of Staphylococcus aureus on the skin surface,

Int.

J. Cosmet. Sci. 38 (2016) 599–606. [183] C. Negin, S. Ali, Q. Xie, Most common surfactants employed in chemical enhanced oil recovery, Petroleum 3 (2017) 197–211. [184] G. A. Jürgenson, C. Bittner, V. Kurkal-Siebert, G. Oetter, J. Tinsley,

Alkyl Ether

Carboxylate Surfactants for Chemically Enhanced Oil Recovery in Harsh Field Conditions,

in:

SPE Asia Pacific Enhanc. Oil Recover. Conf., Society of Petroleum Engineers, 2015. URL: http://www.onepetro.org/doi/10.2118/174589-MS. doi:10.2118/174589-MS. [185] Y.-C. Chiu, H. Hwang,

The use of carboxymethyl ethoxylates in enhanced oil recovery,

Colloids and

Surfaces 28 (1987) 53–65. [186] A. Hussain, P. F. Luckham, T. F. Tadros,

Phase Behaviour of Ph Dependent Microemulsions At High

Temperatures and High Salinities, Rev. L’Institut Français Du Pétrole 52 (1997) 228–231. [187] D. Balzer, Method for recovering largely emulsion-free oil from underground deposits, 1984. URL: https://www.google.com/patents/CA1171778A?cl=en, cA Patent 1,171,778.

34

ACCEPTED MANUSCRIPT

[188] D. Balzer, Process for oil recovery from subterranean reservoir rock formations, 1988. URL: https://www.google.com/patents/US4722396, uS Patent 4,722,396.

T

[189] E. C. Donaldson, G. V. Chilingar, T. F. Yen (Eds.), Enhanced Oil Recovery, II - Processes and

03767361/17/part/PB. [190] Z. Liu, L. Zhang, X. Cao, X. Song, Z. Jin, L. Zhang, S. Zhao,

RI P

Operations, volume 17, Elsevier, Amsterdam, 1989. URL: http://www.sciencedirect.com/science/bookseries/

Effect of electrolytes on interfacial

tensions of alkyl ether carboxylate solutions, Energy and Fuels 27 (2013) 3122–3129.

SC

[191] L. Zhang, Z. Y. Liu, Z. Q. Li, X. W. Song, J. C. Zhang, L. Zhang, S. Zhao,

Dynamic interfacial tensions

of binary nonionic-anionic and nonionic surfactant mixtures at water-alkane interfaces, 91–98.

[192] M. Abe, M. Abe, R. D. Selliah, B. Sheikh, W. H. Wade, R. S. Schechter,

NU

tail ethoxylated carboxylate surfactant/electrolyte/alkane systems, 157–172.

Fuel 135 (2014)

Phase behavior of branched

J. Dispers. Sci. Technol. 8 (1987)

[193] H. Abe, H. Meijer, L. Nogues, Lubrication of textile fibres, 2005. URL: https://www.google.com/patents/

MA

EP1576227A1?cl=en, eP Patent App. EP20,030,795,901.

[194] T. Li, L. Koey, S. Gonzalez, Composition and method for treating textiles, 2007. URL: https://www.google.com/patents/US7279519, uS Patent 7,279,519. [195] R. Ruhr, G. Wichmann, M. Malvey, R. Johnson, D. Halsrud, Conveyor lubricants for use in the food and

ED

beverage industries, 2004. URL: https://www.google.com/patents/US20040167044, uS Patent App. 10/781,385. [196] R. De, H. Meijer, M. Vierbaum, Lubrication of conveyor systems, 2009. URL: https://www.google.com/patents/ EP2042587A1?cl=en, eP Patent App. EP20,070,018,934.

PT

[197] H. Theyssen, S. Grober, H. Kany, M. Wloka, Dry lubrication method employing oil-based lubricants, 2011. URL: https://www.google.com/patents/US20110020558, uS Patent App. 12/934,518. [198] C. Akers, M. Bensing, R. Bogale, X. Cai, J. Li, J. Sun, Q. Yin, Wax emulsion for inkjet ink, 2010. URL:

AC CE

https://www.google.com/patents/US7682434, uS Patent 7,682,434. [199] F. Lang, Liquid detergents containing colour fixing agents, 2009. URL: https://www.google.com/patents/US7541327, uS Patent 7,541,327.

[200] M. Hayashi, T. Suzuki, K. Kurita, M. Kamei, K. Yamaguchi, Fertilizer composition, 2002. URL: https://www.google.com/patents/US6387147, uS Patent 6,387,147. [201] R. Pazhianur, H. Modaressi, F. Smith, Agricultural adjuvant compositions, pesticide compositions, and methods for using such compositions, 2014. URL: https://www.google.com/patents/CA2629862C?cl=en, cA Patent 2,629,862.

[202] L. C. Wang, K. Cui, L. B. Wang, H. X. Li, S. F. Li, Q. ling Zhang, H. B. Liu,

The effect of ethylene

oxide groups in alkyl ethoxy carboxylates on its scale inhibition performance,

Desalination 379 (2016)

75–84. [203] C. Micheau, A. Schneider, L. Girard, P. Bauduin,

Evaluation of ion separation coefficients by foam

flotation using a carboxylate surfactant, Colloids Surfaces A Physicochem. Eng. Asp. 470 (2015) 52–59. [204] M. Schwarze, M. Groß, M. Moritz, G. Buchner, L. Kapitzki, L. Chiappisi, M. Gradzielski, enhanced ultrafiltration (MEUF) of metal cations with oleylethoxycarboxylate,

Micellar

J. Memb. Sci. 478 (2015)

140–147. [205] M. Schwarze,

Micellar-enhanced ultrafiltration (MEUF) – state of the art,

Environ. Sci. Water Res.

Technol. 3 (2017) 598–624. [206] A. A. Mungray, S. V. Kulkarni, A. K. Mungray,

Removal of heavy metals from wastewater using micellar

35

ACCEPTED MANUSCRIPT

enhanced ultrafiltration technique:

a review, Cent. Eur. J. Chem. 10 (2012) 27–46.

[207] B. R. Fillipi, L. W. Brant, J. F. Scamehorn, S. D. Christian,

Use of Micellar-Enhanced Ultrafiltration

T

at Low Surfactant Concentrations and with Anionic–Nonionic Surfactant Mixtures, Sci. 213 (1999) 68–80.

RI P

[208] Y.-Y. Fang, G.-M. Zeng, J.-H. Huang, J.-X. Liu, X.-M. Xu, K. Xu, Y.-H. Qu,

J. Colloid Interface

Micellar-enhanced

ultrafiltration of cadmium ions with anionic–nonionic surfactants, J. Memb. Sci. 320 (2008) 514–519. [209] L. Yurlova, A. Kryvoruchko, B. Kornilovich,

Removal of Ni(II) ions from wastewater by micellar-enhanced

SC

ultrafiltration, Desalination 144 (2002) 255–260.

AC CE

PT

ED

MA

NU

[210] F. Sebba, Concentration by Ion Flotation, Nature 184 (1959) 1062–1063.

36

ACCEPTED MANUSCRIPT

Graphical abstract OH O

O

O

C

T

O

Multiresponsive properties

Industrial Applications

Complex self-assembly

RI P

Colloidal building blocks

SC

O pH-responsive

degree of io ionization on

Temperature-sensitive

AC CE

PT

ED

MA

NU

Headgroup volume fraction

37

ACCEPTED MANUSCRIPT

Highlights

• pH- and Temperature-responsive behavior is reported

RI P

T

• Overview of physico-chemical properties of multiresponsive surfactant

• Extremely versatile building blocks for colloidal supramolecular assemblies

AC CE

PT

ED

MA

NU

SC

• Review of technical and large-scale applications of the surfactants

38