Understanding and manipulating micelle morphology are key to exploiting surfactants in various applications. Recent studies have shown surfactant self-assembly in a variety of Deep Eutectic Solvents (DESs) where both the nature of surfactants and the interaction of the surfactant molecule with the solvent components influence the size, shape, and morphology of the micelles formed. So far, micelle formation has only been reported in type III DESs, consisting solely of organic species. In this work, we have explored the self-assembly of cationic surfactant dodecyl trimethylammonium nitrate/bromide (C12TANO3/C12TAB), anionic surfactant sodium dodecyl sulfate (SDS), and non-ionic surfactants hexaethylene glycol monododecyl ether (C12EO6) and octaethylene glycol monohexadecyl ether (C16EO8) in a type IV DES comprising metal salt, cerium (III) nitrate hexahydrate, and a hydrogen bond donor, urea, in the molar ratio 1:3.5. C12TANO3, C12TAB, C12EO6, and C16EO8 form spherical micelles in the DES with the micelle size dependent on both the surfactant alkyl chain length and the head group, whereas SDS forms cylindrical micelles. We hypothesize that the difference in the micelle shape can be explained by counterion stabilization of the SDS headgroup by polycations in the DES compared to the nitrate/bromide anion interaction in the case of cationic surfactants or molecular interaction of the urea and the salting out effect of (CeNO3)3 in the DES on the alkyl chains/polyethoxy headgroup for non-ionic surfactants. These studies deepen our understanding of amphiphile self-assembly in this novel, ionic, and hydrogen-bonding solvent, raising the opportunity to use these structures as liquid crystalline templates to generate porosity in metal oxides (ceria) that can be synthesized using these DESs.

Ionic liquids and deep eutectic solvents (DESs) have gained interest as alternatives to traditional solvents in many applications from material synthesis1–6 and electrochemistry7–13 to pharmaceuticals14–16 due to their low vapor pressure, low flammability, wide liquid range, and tunable nature. In recent times, there has been an upsurge in the field of studying self-assembly of ionic17–23 and non-ionic23–27 surfactants in ionic liquids and DESs, ever since the first study reporting self-assembly of surfactants in ethylammonium nitrate in 1982.28 Studies of self-assembly in these novel solvents give an insight into both fundamental interactions in these alternative media and offer unique potential for uses in the fields of emulsification and templating applications, e.g., synthesis of polymers and polymer membranes29–32 or porous materials33–35 and use in transdermal delivery systems.36 

In addition to their shared characteristics, DESs offer some unique advantages in comparison to ionic liquids. They often comprise cost-effective, widely available, and environmentally benign components making them biocompatible and have therefore attracted great interest as ionic liquid analogs. DESs are generally made by complexation of a hydrogen bond donor with a salt capable of sustaining a hydrogen bond network, which results in a significant dip in their melting point, which is lowest at a particular molar ratio called the eutectic point.37 Their hydrophobicity, physicochemical properties, and nanostructure can be altered by changing the constituents or their ratios, making them highly tunable for specific applications. Studies have shown that both DES nanostructure and composition play a crucial role in self-assembly of various amphiphiles.38,39

Surfactant micelle templated synthesis of mesoporous materials, such as silica40,41 and metal oxides,42 has widely been exploited in traditional solvents. Recently, Hu et al. reported surfactant template-assisted synthesis of hierarchical ZIF-8 particles using sodium dodecyl sulfate (SDS) in choline chloride/urea (ChCl:U) DESs.35 They reported that the porosity in the ZIF-8 particles can be tuned by varying the amount of SDS and water in the reaction mixture. Hammond et al. also reported solvothermal synthesis of iron oxide5 and cerium oxide6 nanoparticles from ChCl:U DESs where the morphology and porosity are controlled by the water content of the mixture. Furthermore, it is understood that when synthesizing these oxides, the urea component of the DES is crucial to the solvothermal reaction mechanism. Therefore, self-assembly of surfactants in urea containing DESs offers an interesting potential for templating metal oxides in solvothermal synthesis.

Until now, the most commonly studied DESs are ones that comprise a quaternary ammonium halide salt with a hydrogen bond donor, e.g., choline chloride (ChCl) with urea (U), glycerol (Gly), ethylene glycol (EG), or some carboxylic acid. Most of these fall into the category of type III DESs.43 However, there are other types of DESs comprising metal salts or their hydrates with quaternary ammonium salts, type I (ZnCl2:ChCl) and type II (ZnCl2·3H2O:ChCl), respectively, or metal salts with a hydrogen bond donor, type IV (ZnCl2:urea, AlCl3:urea, and CrCl3·6H2O:urea). Recently, a halide-free type IV DES was reported by Hammond et al. comprising hydrated salts of lanthanide metals (cerium, praseodymium, and neodymium) with urea.44 These are low melting point and low viscosity DESs with very high density and surface tension. Their Gordon parameter is similar to that of water, suggesting potential for surfactant aggregation/micellization due to solvophobic interactions.45 Lanthanide oxides can be prepared from these solvents by controlled combustion, showing potential use of these solvents in efficient synthesis applications. Surfactant self-assembly in DESs besides type III DESs is a relatively unexplored field and offers the unique opportunity to understand intermolecular interactions in a solvent comprising metal salts along with organic molecular components as well as the potential for templated synthesis applications.

In this work, we have explored the self-assembly of a range of surfactants: cationic (C12TAB and C12TANO3), anionic (SDS), and non-ionic (C12EO6 and C16EO8); in a type IV DES comprising cerium (III) nitrate hexahydrate and urea in the molar ratio 1:3.5, dodecyl chain (C12) surfactants were the main focus in this study as they have a low Krafft temperature facilitating measurements at room temperature and a wide range of data are available for the self-assembly of these surfactants in various solvents for comparison. The bromide and nitrate variants of the C12 alkyl chain cationic surfactant allow us to discern any differences arising in the DES interaction with the solvent due to complexation of different anions with the cerium salt. In addition to the dodecyl chain lengths, we have also studied C16EO8 to understand the effect of change in both alkyl chain length and polyethoxy head group length on the surfactant self-assembled structures. Surfactant self-assembly is studied using interfacial tension to establish the surface activity of these surfactants in the DES along with small angle neutron scattering (SANS) measurements to demonstrate the existence of micelles and determine their structures.

Cerium nitrate hexahydrate [Ce(NO3)3·6H2O, 99.5%] was purchased from Acros Organics, and urea (h-U, ≥99.5%) was purchased from Sigma-Aldrich. Deuterated urea-d4 (d-U, 99% atom, 98% D) was purchased from Cambridge Isotope Laboratories.

Dodecyl trimethylammonium bromide (C12TAB, ≥99%), sodium dodecyl sulfate (SDS, ≥99%), hexaethylene glycol monododecyl ether (C12EO6, ≥98%), and octaethylene glycol monohexadecyl ether (C16EO8, ≥98%) were purchased from Sigma-Aldrich. Dodecyl trimethylammonium nitrate (C12TANO3) was prepared by using an ion exchange procedure46 using Amberlite® IRN78 hydroxide (Acros Organics) from C12TAB. The ion exchange resin was stirred with 1M NaOH for 2 h and then rinsed with water until a pH of 7 was obtained. Thereafter, 10 g of C12TAB was added to a flask containing 100 g of the ion exchange resin and 150 ml of deionized water. The solution was stirred for a few hours until the entire C12TAB dissolved. The solution was filtered into a new batch of 100 g of ion exchange resin and 150 ml of water and stirred for a further 2 h and repeated once more. The strongly basic solution is separated from the resin by filtration and rinsing with copious amount of water. This results in C12TAOH, which is then titrated with dilute nitric acid until a pH ≈6 was obtained to get C12TANO3. The solution was tested for any remaining C12TAB using silver nitrate as any bromide will result in silver bromide, which will precipitate out of the solution. Once checked, the solutions were freeze dried to obtain dry C12TANO3 powder.

Isotopically labeled C12TAB (d34, d-C12TAB) was supplied by the ISIS Deuteration Facility. Deuterated C12TANO3 (d34, d-C12TANO3) was prepared from d-C12TAB using the ion exchange method described above and using D2O instead of deionized H2O.

The Ce(NO3)3·6H2O:urea (Ce:U) DES was prepared by combining the components in molar ratios of 1:3.5. The mixtures were stirred at room temperature until a clear, homogeneous liquid was obtained, which was subsequently sealed and equilibrated overnight. Once formed, the mixtures were stable and were used within 2–3 weeks from preparation. For the DES containing deuterated urea, Ce:d-U, Ce(NO3)3·6H2O, and d4-urea were mixed in the molar ratio of 1:3.5.

The surfactant in DES solutions containing C12TAB, C12TANO3, SDS, C12EO6, and C16EO8 was mixed and equilibrated at room temperature until homogeneous mixtures were obtained. Concentration series of various surfactants were obtained by diluting high concentration (10 wt. %) solutions.

The density of the DES was determined from a triplicate average of measurements on an Anton Paar DMA 4500M at 25 °C. Differential scanning calorimetry (DSC) measurements to determine the melting/transition temperature of the neat DES were carried out on a TA Instruments DSC-Q20 differential scanning calorimeter. The sample was first equilibrated at 25 °C and held for 5 min and then cooled to −75 °C at a ramp rate of 5 °C min−1 and held for 10 min before heating to 30 °C at the same ramp rate. The viscosity of the DES was measured using a TA Instruments HR-3 Discovery Hybrid Rheometer operating in a flat plate geometry with Peltier temperature control. Room temperature viscosity data for the DES were obtained for the applied shear rate ranging from 0.1 to 100 s−1 at 25 °C and a Newtonian behavior was observed. Therefore, the temperature sweep data were measured at a shear rate of 1 s−1 for the temperature ranging from −10 to 40 °C.

Surface tension measurements were made using the drop-shape-analysis method47 with a FTA1000 drop shape analyzer. A concentration series of the surfactants in DESs ranging from 0.002 to 1 wt. % (0.1–60 mM, depending on the surfactant) were prepared and measured at room temperature. Each concentration was measured five times using a fresh drop, with each measurement comprising 100 data points taken within 1 min. The data reported are the average value of the five measurements with the standard error.

Small angle neutron scattering (SANS) measurements were carried out on the SANS2D48 instrument at the ISIS Pulsed Neutron and Muon Source, UK (RB192067649). Using a sample-to-detector distance of 4 m and neutrons with wavelengths of 2–16.5 Å, a q-range of 0.0042–0.82 Å−1 was obtained. The samples were loaded into 1 mm path length rectangular quartz cuvettes (Hellma GmbH) and placed on the automatic sample changer on the beamline. The measurements were performed at 25 °C. Data reduction was performed according to the standard procedures using the routines within Mantid,50 resulting in output converted to absolute units of the scattering intensity [I(q), cm−1] vs the momentum transfer (q, Å−1). Subtraction of the scattering from the pure solvents was performed afterward using the NIST NCNR macros in Igor Pro51 to account for the incoherent contribution to each sample.

Surfactant mixtures in DESs were prepared at concentrations higher than the CMC, as determined by the surface tension, at 1 wt. % in different isotopic mixtures of the DES to aid in resolving the micelle structure. Protonated surfactants (C12TANO3, C12TAB3, SDS, C12EO6, and C16EO8) were measured in Ce:d-U and Ce:U DES to provide information on the size of the micelle core as the scattering is dominated by the micelle core–solvent scattering length density contrast plus some information about the shell structure. Furthermore, to get the details of the micelle structure and solvent surfactant interactions, d-C12TANO3 and d-C12TAB were also measured in Ce:U and Ce:d-U DES. In addition to the 1 wt. % concentrations, for C12TANO3 and C16EO8, a series of measurements were performed with concentration ranging from 0.5 to 10 wt. % to study the concentration dependence of the surfactant micelle structure. The neutron scattering length densities (SLDs) for the constituent components of the surfactants and DES and the SLDs of the DES at the two contrasts are given in the supplementary material (Tables S1 and S2).

The SANS data were analyzed by co-fitting the various isotopic contrasts using standard spherical and cylindrical form factors and their core–shell variants in SasView.52 In the case of high concentration data, where inter-particle interactions were observed, the Percus–Yevick hard sphere structure factor53 was used to account for the structure factor contribution to the scattering as employed in previous work on similar systems.21,38 The details for the fitting models and procedure can be found in the supplementary material (Sec. S3).

Cerium nitrate hexahydrate and urea form a homogeneous room-temperature liquid at a molar ratio of 1:3.5 as previously reported by Hammond et al.44 The glass transition temperature as determined using extrapolation of the viscosity/temperature relationship with the Vogel–Fulcher–Tammann (VFT) equation54 and differential scanning calorimetry (DSC), shown in the supplementary material, was found to be ∼−60 °C, consistent with the studies by Hammond et al.Table I summarizes the main physical properties of the Ce:U DES along with the corresponding values for Ce:U DESs reported by Hammond et al., a typical type III DES ChCl:U, water and formamide, and a polar organic solvent, for comparison. Molar volumes for the DES were calculated using their measured densities and average molar masses, and the Gordon parameter was calculated using the surface tension and molar volume.45 This parameter is a description of the “solvophobicity” of a solvent and provides a measure of the cohesiveness of solvent molecules. It may therefore be used to predict the capability of a solvent to promote the self-assembly of amphiphilic molecules.19,55,56

TABLE I.

Measured physical properties at room temperature (298 K) for Ce:U DES and the computed molar volume and Gordon parameter, alongside literature data for Ce:U DES, “type III” ChCl:urea DES, and H2O for comparison.

Average molar massDensityMolar volumeaViscositySurface tensionGordon parameter (G)b
DES(g mol−1)(g cm−3)( cm3 mol−1) (mPa s)(mN m−1)(J m−3)
Ce:U (this work) 47.73 1.786 26.73 179 ± 3 85.65 ± 0.25 2.86 
Ce:Uc 47.73 1.789 26.67 168 83.33 ± 0.32 2.79 
ChCl:Ud 86.6 1.195 72.47 1570 66 ± 1 1.58 
H2Oe 18.02 0.997 18.02 0.889 72 2.74–2.75 
Formamidef 45.04 1.133 39.75 3.23 58.2 1.50–1.70 
Average molar massDensityMolar volumeaViscositySurface tensionGordon parameter (G)b
DES(g mol−1)(g cm−3)( cm3 mol−1) (mPa s)(mN m−1)(J m−3)
Ce:U (this work) 47.73 1.786 26.73 179 ± 3 85.65 ± 0.25 2.86 
Ce:Uc 47.73 1.789 26.67 168 83.33 ± 0.32 2.79 
ChCl:Ud 86.6 1.195 72.47 1570 66 ± 1 1.58 
H2Oe 18.02 0.997 18.02 0.889 72 2.74–2.75 
Formamidef 45.04 1.133 39.75 3.23 58.2 1.50–1.70 
a

Calculated by dividing the average molar mass of the DES by its density.

b

Gordon parameter is calculated using G=γ/Vm3, where γ is the air–solvent interfacial tension and Vm is the molar volume.

c

The values as reported by Hammond et al.44 

d

From the work of Arnold et al.20 and Xie et al.57 

e

From the work of Vargaftik et al.,58 Adamenko et al.,59 and Greaves et al.55 

f

From the work of Rico and Lattes60 and Wärnheim and Jönsson.61 

The Ce:U DES shows physical properties, which are significantly different from typical type III DESs, i.e., ChCl:U. While the differences in density, average molar mass, and volume can be attributed to the presence of relatively heavy cerium ions, the more important and interesting differences are the ones observed in viscosity and surface tension. Based on structural investigations, Hammond et al.44 suggested that the low viscosity and high surface tension for the Ce:U DES result from the presence of strongly bound transient oligomers comprising [–Ce–NO3–], which form a molten salt-like network, in flux with an interpenetrating H-bonded network of urea, water, and some nitrate ions. The Ce:U DESs have significantly high surface tensions (∼85 mN m−1) when compared to typical type III DESs (40–70 mN m−1),62 ionic liquids (50–60 mN m−1),63 and even higher than water (72 mN m−1). The high surface tension and the small molar volume result in a Gordon parameter, which is higher than in most DES and ionic liquids19,20 but comparable to that of water, suggesting that their cohesive energy density is comparable to or even slightly higher than water. In addition to the comparatively high Gordon parameter, the polarizability of the solvent is also expected to be high due to the ionic nature of the constituents, which should also increase the tendency of amphiphiles to form micelles.64 A combined effect would suggest that the Ce:U DES could promote self-assembly in a range of surfactants.

The surface tension vs molar concentration curves for C12TAB, C12TANO3, SDS, C12EO6, and C16EO8 in Ce:U DESs are shown in Fig. 1. Since our conversion to the molar concentration is dependent on the measured density of the DES assuming that the volume of the DES is not influenced by the presence of surfactants, we have also plotted the data against the percentage weight concentration and mole fraction, and they are shown in the supplementary material (Fig. S6). The surface tension shows a clear variation with concentration, indicating that all the surfactants measured here remain surface-active in the DES solutions. For each surfactant, the shapes of the surface isotherm exhibit the classic behavior of a surfactant in aqueous solution65 and other DESs.20–22 Initially, the surface tension of the solution gradually decreases with the addition of surfactant until it reaches the critical micellar concentration (CMC) beyond which the surface tension essentially remained the same and a plateau is observed. We estimated the CMC from the intersection of the two straight lines: one obtained by a linear approximation to the first three data points in the region of the graph just below the break point concentration and the second being the surface tension in the plateau region above the break point concentration. The black lines (solid and dashed) in Fig. 1 represent the trends that help determine the CMC. Additionally, the surface excess (Γ) was calculated from the surface tension (γ) vs ln concentration ([conc]) graph using the following relation:

(1)

where dγdln[conc] is the tangent to the surface tension vs ln concentration graph just before the CMC, R is the gas constant, T is the absolute temperature, and n takes the value of 1 for non-ionic surfactants and 2 for ionic surfactants. The surface excess can then be used to calculate the area per molecule (Apm) in Å2 using the following relation:

(2)

where NA is Avogadro’s number. Table II shows the CMC values in mM for the various surfactants (CMC values in wt. % and mole fraction are available in the supplementary material, Table S7), the surface tension at CMC (γCMC), surface excess (Γ) and the area per molecule (Apm) in the Ce:U DES along with the literature CMC, γCMC, surface excess and area per molecule values for the surfactants in water, and the CMC values in an organic polar solvent, formamide, for comparison.

FIG. 1.

Surface tension vs molar concentration (mM) of various surfactants in Ce:U DESs: (a) cationic surfactants C12TAB and C12TANO3, (b) anionic surfactant SDS, and (c) non-ionic surfactants C12EO6 and C16EO8. The black lines (solid and dashed) are guides to the eye, representing the trends that help us to determine the CMC.

FIG. 1.

Surface tension vs molar concentration (mM) of various surfactants in Ce:U DESs: (a) cationic surfactants C12TAB and C12TANO3, (b) anionic surfactant SDS, and (c) non-ionic surfactants C12EO6 and C16EO8. The black lines (solid and dashed) are guides to the eye, representing the trends that help us to determine the CMC.

Close modal
TABLE II.

CMC values calculated from CE:U/air surface tension measurements for the surfactants in this study along with literature CMC values of the surfactants in H2O and formamide.

Ce:UH2OFormamide
CMCγCMCaΓ × 10−6ApmCMCγCMCaΓ × 10−6ApmCMC
Surfactant(mM)(mN m−1)(mol m−2)−2)(mM)(mN m−1)(mol m−2)−2)(mM)
C12TAB 5.4 ± 0.9 35.7 2.36 70 15.0b 36.4b 2.6b 64b Solublec 
C12TANO3 5.4 ± 0.9 36.5 2.20 75 12d … … … … 
SDS 1.9 ± 0.2 28.6 3.10 53 8.2e 38–40e 3.9e 42e 220f 
C12EO6 1.2 ± 0.2 38.1 2.56 65 0.087g 31.9g 4.3g 38g 31h 
C16EO8 0.7 ± 0.06 39.6 2.45 68 0.0016i 43i 4.4i 38i 11h 
Ce:UH2OFormamide
CMCγCMCaΓ × 10−6ApmCMCγCMCaΓ × 10−6ApmCMC
Surfactant(mM)(mN m−1)(mol m−2)−2)(mM)(mN m−1)(mol m−2)−2)(mM)
C12TAB 5.4 ± 0.9 35.7 2.36 70 15.0b 36.4b 2.6b 64b Solublec 
C12TANO3 5.4 ± 0.9 36.5 2.20 75 12d … … … … 
SDS 1.9 ± 0.2 28.6 3.10 53 8.2e 38–40e 3.9e 42e 220f 
C12EO6 1.2 ± 0.2 38.1 2.56 65 0.087g 31.9g 4.3g 38g 31h 
C16EO8 0.7 ± 0.06 39.6 2.45 68 0.0016i 43i 4.4i 38i 11h 
a

γCMC is the limiting surface tension of the surfactant at the CMC.

b

From the work of Mata et al.66 and Hooshyar and Sadeghi.67 

c

Couper et al. reported no micellization of C12TAB in formamide up to the limit of solubility.74 CMC of C14TAB is reported as 270 mM75 and C16TAB as 90 mM.60 

d

From the work of Dolan et al.18 

e

From the work of Xu et al.68 and Ananda et al.69 

f

From the work of Rico and Lattes.60 

g

From the work of Corkill et al.70 and Wijaya et al.,71γCMC reported here is measured at 15 °C.

h

From the work of Jonstroemer et al.76 

i

From the work of Kabin et al.72 and Tamura et al.73 

The CMC of cationic surfactants, C12TAB and C12TANO3, in Ce:U DESs is less than that in water and even lower than that observed for ionic liquids19,23,28 and also ChCl-based DESs.21,22,77 This is possibly due to the salting out effect in the DES owing to its ionic nature in addition to a high Gordon parameter. This is similar to the observations in aqueous solutions where the CMC of CTAB surfactants decreases as the salt concentration of the solution increases.78–80 In Ce:U DESs, the entire solvent comprises ions and a combination of this along with the high Gordon parameter drives micellization at a lower concentration. There is no measurable effect of counterions on the CMC of dodecyltrimethylammonium (C12TA+) surfactants in Ce:U, an effect also observed in ionic liquids, e.g., ethanolammonium nitrate;18 however, this is contrary to measurements of the CMC for cationic surfactants in aqueous solutions81–83 and some other ionic liquids.18 This could potentially be explained by counterion exchange with the nitrate ions in the Ce:U DES resulting in little effect of counterions having a similar size at low surfactant concentrations. This is also observed for hexadecyltrimethylammonium (C16TA+) in ChCl:Gly where the CMCs of C16TAB (15.2 mM) and C16TACl (15.4 mM) are very similar; however, that of C16TATS, which has a significantly larger counterion, is higher (20.8 mM).84 In the case of the anionic surfactant SDS in Ce:U, the CMC is lower than that in water and comparable to the CMC observed in ChCl:U,20 which is attributed to the high polarizability parameter for these eutectic solvents and the presence of counterions, i.e., [Choline]+ in ChCl:U or [–Ce–NO3–]+ polycations in the Ce:U DES.

On the other hand, in the case of non-ionic surfactants in Ce:U, a significantly higher CMC is observed as compared to water. This is similar, although smaller in magnitude, to the effect observed in ionic liquids,19,25,26 where CMCs that are 3–5 orders of magnitude higher than water are observed. In ionic liquids, this is explained due to the weaker solvophobic interaction in the solvent compared to the hydrophobic effect in water owing to the presence of alkyl chains. A similar effect on CMC is also observed for non-ionic surfactants in formamide76 and ethylene glycol.19,25 These solvents, though polar in nature, exhibit some organic character, which increases the solubility of the surfactant alkyl chains and reduces the solvophobic interactions with the non-ionic surfactants leading to a higher CMC, an effect also observed upon addition of urea to aqueous solutions of non-ionic surfactants.85,86 For C12EO6 and C16EO8 in Ce:U, the presence of urea has the potential to increase the solubility of the surfactant alkyl chain, while Ce(NO3)3 could result in a salting in/out effect on both the hydrophilic and hydrophobic moieties in the surfactant.87 The increased solubility of the alkyl chains and the salting in effect on the polyethyoxy head group result in an increase in CMC, whereas the salting out effect on the hydrophobic moieties decreases the CMC.88 A combination of the two reduced the solvophobic interactions of the surfactant with Ce:U when compared to water, though not to the same extent as that in polar organic solvents or ionic liquids. Therefore, for non-ionic surfactants in Ce:U, a CMC in between that in water and polar organic/ionic liquids is observed. The Gibbs free energy of micellization (ΔG = RT ln CMC) for C12EO6 was calculated to be −16.7 kJ/mol in Ce:U DES, compared to −23.3 kJ/mol in water and −8.7 kJ/mol in formamide. In addition, we see a decrease in CMC with an increase in the alkyl chain length/polyethoxy group length. This can be expected due to an increase in the solvophobic interaction as the alkyl or polyethoxy chain length increases and is consistent with observations in water;70,72 polar organic liquids, such as formamide and ethylene glycol;25,76 ionic liquids;19,26 and aqueous solutions in the presence of urea85,86 or salts.87 As expected, with a decrease in CMC, the Gibbs free energy of micellization for C16EO8 also decreases and was calculated to be −18.2 kJ/mol in Ce:U DES, compared to −33.3 kJ/mol in water and −11.2 kJ/mol in formamide. The magnitude of the decrease in CMC and ΔG in Ce:U are less than that in water and comparable to the effect observed in formamide and ionic liquids;19,25,26(ΔGC16EO8ΔGC12EO6) was calculated as −1.5 kJ/mol for Ce:U DES, −10 kJ/mol for water, and −2.5 kJ/mol for formamide. The trends in the change in CMC and ΔG are another indication that the interactions of the non-ionic surfactants with the DES constituents are more akin to polar organic solvents or ionic liquids and both the organic (urea) and ionic [Ce(NO3)3] constituents of the DES impact the CMC.

The surface excess for all surfactants in the Ce:U DES is positive, indicating that the concentration of the surfactants is higher at the air/DES interface compared to the bulk, representative of a surface-active monomer with a hydrophobic chain that prefers to be at the interface compared to in the bulk solution. As expected, the surface excess and the area per molecule of the two cationic surfactants and the two non-ionic surfactants are in good agreement with one another. The area per molecule of the surfactants in Ce:U DES is greater than that observed in water and similar to some polar organic solvent and ionic liquids.71 This suggests that the solvophobic interactions, potentially governed by the hydrophobic chain, in the Ce:U DES are weaker than those in water and similar to organic polar solvents or ionic liquids, possibly explained by the presence of urea in the DES.

1. C12TAB and C12TANO3 in Ce:U

SANS was measured from 1 wt. % solutions of C12TAB and C12TANO3 in Ce:U for h-C12TA+ in hh-DES and hd-DES and d-C12TA+ hh-DES and hd-DES contrasts to resolve the structure of the C12TA+ micelles in the DES. The four SANS contrasts were co-fitted to a spherical form factor and its core–shell variant (see the supplementary material, Fig. S3 and Table S3) with the dodecyl chains forming the core and the trimethylammonium cationic headgroup forming the shell. It is not possible to fit all the contrasts to a spherical model with a single radius for either of the two cationic surfactants, as can be seen in Fig. S3. However, using the core–shell model, if we allow the shell SLD to vary to account for the changes in the SLDs arising due to the solvation of the micelle head group by the ions in the DES, all four contrasts can be fit to a single core-radius and shell thickness with reasonable confidence. This is also reflected in the reduced χ2 parameter, which drops from 10.2 to 2.5 for C12TAB and 6.05 to 1.67 for C12TANO3 when the data are fitted to the core–shell spherical form factor instead of a homogenous spherical model (see the supplementary material, Table S3). The SANS data are shown along with the best fit core–shell spherical model for the various contrasts in Fig. 2, and the parameters are summarized in Table III. In Table III, the core-radius and shell thickness obtained from the combined fit using the h-C12TA+ in Ce:U, h-C12TA+ in Ce:d-U, d-C12TA+ in Ce:U, and d-C12TA+ in Ce:d-U contrasts are reported along with the SLD of the shell at the four contrasts and the calculated solvent volume fraction of the DES in the shell region (ϕ).

FIG. 2.

SANS patterns for 1 wt. % (a) C12TAB and (b) C12TANO3 in Ce:U at various contrasts: red circles represent h-C12TA+ in Ce:U (bottom), blue triangles represent h-C12TA+ in Ce:d-U (second bottom), green diamonds represent d-C12TA+ in Ce:U (second top), and orange squares represent d-C12TA+ in Ce:d-U (top). The various traces are offset along the y axis for clarity. The black dashed lines represent the best fits obtained to the core–shell spherical form factor through co-refinement of all the contrasts.

FIG. 2.

SANS patterns for 1 wt. % (a) C12TAB and (b) C12TANO3 in Ce:U at various contrasts: red circles represent h-C12TA+ in Ce:U (bottom), blue triangles represent h-C12TA+ in Ce:d-U (second bottom), green diamonds represent d-C12TA+ in Ce:U (second top), and orange squares represent d-C12TA+ in Ce:d-U (top). The various traces are offset along the y axis for clarity. The black dashed lines represent the best fits obtained to the core–shell spherical form factor through co-refinement of all the contrasts.

Close modal
TABLE III.

Core–shell spherical form factor fit parameters for the SANS data from 1 wt. % C12TAB and C12TANO3 in Ce:U shown in Fig. 2 at the different contrasts.

SurfactantCore-radius (Å)Shell thickness (Å)ContrastSLD shell (Å−2 × 10−6)ϕaNaggb
C12TAB 15.0 ± 1.1 8.3 ± 0.7 h-C12TAB in Ce:U 1.5 ± 0.25 0.72 ± 0.12 37 ± 8 
   h-C12TAB in Ce:d-U 2.9 ± 0.18 0.65 ± 0.04  
   d-C12TAB in Ce:U 4.4 ± 0.23 0.54 ± 0.03  
   d-C12TAB in Ce:d-U 5.6 ± 0.11 0.58 ± 0.01  
C12TANO3 16.2 ± 0.9 5.8 ± 0.6 h-C12TANO3 in Ce:U 1.5 ± 0.21 0.72 ± 0.10 45 ± 7 
   h-C12TANO3 in Ce:d-U 2.2 ± 0.58 0.51 ± 0.13  
   d-C12TANO3 in Ce:U 3.3 ± 0.34 0.77 ± 0.08  
   d-C12TANO3 in Ce:d-U 4.8 ± 0.24 0.92 ± 0.05  
SurfactantCore-radius (Å)Shell thickness (Å)ContrastSLD shell (Å−2 × 10−6)ϕaNaggb
C12TAB 15.0 ± 1.1 8.3 ± 0.7 h-C12TAB in Ce:U 1.5 ± 0.25 0.72 ± 0.12 37 ± 8 
   h-C12TAB in Ce:d-U 2.9 ± 0.18 0.65 ± 0.04  
   d-C12TAB in Ce:U 4.4 ± 0.23 0.54 ± 0.03  
   d-C12TAB in Ce:d-U 5.6 ± 0.11 0.58 ± 0.01  
C12TANO3 16.2 ± 0.9 5.8 ± 0.6 h-C12TANO3 in Ce:U 1.5 ± 0.21 0.72 ± 0.10 45 ± 7 
   h-C12TANO3 in Ce:d-U 2.2 ± 0.58 0.51 ± 0.13  
   d-C12TANO3 in Ce:U 3.3 ± 0.34 0.77 ± 0.08  
   d-C12TANO3 in Ce:d-U 4.8 ± 0.24 0.92 ± 0.05  
a

ϕ is the solvent volume fraction in the shell. This is calculated from the SLD of the shell (TA+ head group) and solvent at that contrast.

b

Nagg is the aggregation number calculated by dividing the micelle core volume by the tail volume calculated from the Tanford equation.89 

The radii of the micelles are 15.0 ± 1.1 Å for C12TAB and 16.2 ± 0.9 Å for C12TANO3, which are consistent with extended C12 chain lengths,89 and the shell thicknesses are 8.3 ± 0.7 Å for C12TAB and 5.8 ± 0.6 Å for C12TANO3. Both the radius and shell thickness are in good agreement with values for alkyltrimethylammonium surfactants in other deep eutectic solvents,21,38 ionic liquid ethanolammonium nitrate,18 and in water.83,90 A more recent study by Hargreaves et al. of C10TAB micelles in water using wide q-range neutron diffraction and atomistic modeling gives a hydrated shell thickness of 7.5 Å.91 Our values of solvated shell thickness, obtained using co-refinement of multiple contrasts, in SANS are in good agreement with this value and provide an average solvent volume fraction of 0.62 ± 0.07 for C12TAB and 0.73 ± 0.15 for C12TANO3 in the shell region, possibly as a result of interaction of the trimethylammonium head group with the nitrate counterions in the DES. The cross section of the micelles is similar to that reported in type III DES-like ChCl:Gly21 and ChCl:U:Gly;38 however, unlike type III DESs where elliptical micelles are reported for C12TAB, spherical micelles are observed for C12TA+ in the Ce:U DES. It must be noted here that for the elliptical micelles of C12TAB in type III DESs, the axial ratio <2,21,38 and arguably, they can be fitted to spherical form factors as well and therefore are generally treated as spheroidal. The aggregation number for the micelles was calculated to be 37 ± 8 for C12TAB and 45 ± 7 for C12TANO3, giving an area per molecule of 79 ± 15 and 74 ± 12 Å2, respectively. These are in good agreement with the values calculated from the surface tension data. Both the structure and size of the micelles show no particular dependence on the counterion. This indicates that the ions in this DES behave similarly to a swamping amount of electrolytes in a solution, resulting in all counterions being exchanged with the solvent, and therefore, no effect of counterion is observed on the shape and size of the micelles, an effect also observed for these cationic micelles in ionic liquid ethanolammonium nitrate.18 

2. SDS in Ce:U

SANS data were collected for 1 wt. % SDS in Ce:U in both hh-DES and hd-DES contrast and are shown in Fig. 3. The data show the scattering characteristic of elongated structures where I(q) ∝ q−1. The two SANS contrasts were co-fitted to a cylindrical form factor and its core–shell variant (see the supplementary material, Fig. S4 and Table S4) with the dodecyl chains forming the core and the sulfate headgroup forming the shell. The SLD of the shell region was allowed to fit in order to take into account solvation of the sulfate head group by the DES. The core–shell cylindrical form factor gives a marginally better fit in this case; reduced χ2 = 3.19 as compared to 3.44 for cylindrical model (see the supplementary material, Table S4). It also allows for a better comparison with the data obtained for other surfactants and is therefore presented in Fig. 3, and the fit parameters are given in Table IV. In Table IV, the core-radius, shell thickness, and length obtained for SDS micelles from the combined fit using the hh-DES and hd-DES contrasts are reported along with the SLD of the shell at the two contrasts and the calculated solvent volume fraction of the DES in the shell region (ϕ).

FIG. 3.

SANS patterns for 1 wt. % SDS in Ce:U at two contrasts: red circles represent h-SDS in Ce:U (bottom) and blue triangles represent h-SDS in Ce:d-U (top). The various traces are offset along the y axis for clarity. The black dashed lines represent the best fits obtained to the core–shell cylindrical form factor through co-refinement of the two contrasts.

FIG. 3.

SANS patterns for 1 wt. % SDS in Ce:U at two contrasts: red circles represent h-SDS in Ce:U (bottom) and blue triangles represent h-SDS in Ce:d-U (top). The various traces are offset along the y axis for clarity. The black dashed lines represent the best fits obtained to the core–shell cylindrical form factor through co-refinement of the two contrasts.

Close modal
TABLE IV.

Core–shell cylindrical form factor fit parameters for the SANS data from 1 wt. % SDS in Ce:U shown in Fig. 3 at the different contrasts.

SurfactantCore-radius (Å)Shell thickness (Å)Length (Å)ContrastSLD shell (Å × 10−6)ϕaNaggb
SDS 15.1 ± 0.25 4.5 ± 0.22 4069 ± 33c h-SDS in Ce:U 1.75 ± 0.08 0.74 ± 0.03 7315 ± 35 
    h-SDS in Ce:d-U 2.92 ± 0.18 0.60 ± 0.04  
SurfactantCore-radius (Å)Shell thickness (Å)Length (Å)ContrastSLD shell (Å × 10−6)ϕaNaggb
SDS 15.1 ± 0.25 4.5 ± 0.22 4069 ± 33c h-SDS in Ce:U 1.75 ± 0.08 0.74 ± 0.03 7315 ± 35 
    h-SDS in Ce:d-U 2.92 ± 0.18 0.60 ± 0.04  
a

ϕ is the solvent volume fraction in the shell. This is calculated from the SLD of the shell (SO−4 head group) and solvent at that contrast.

b

Nagg is the aggregation number calculated by dividing the micelle core volume by the tail volume calculated from the Tanford equation.89 

c

The length obtained for the cylindrical micelles from the model fits is beyond the detection limit of the instrument at a detector length of 4 m (∼1500 Å). From the data, we can estimate that the length is ≫1500 Å, but the exact value cannot be calculated from these measurements.

SDS forms extremely elongated micelles in Ce:U DESs. The radius of the micelles is 15.1 ± 0.25 Å, shell thickness is 4.5 ± 0.22 Å, and length is ≫1500 Å. The aggregation number for the SDS micelles was calculated to be 7315 ± 350, giving an area per molecule of 53 ± 3 Å2, consistent with the surface tension measurements. The solvent volume fraction in the shell region, calculated from the SLD of the shell, is 60%–75%, indicating highly solvated shells possibly due to interactions between the cations in the DES with the sulfate head group. The cross section of the micelles is similar to that reported in type III DES-like ChCl:U,20 ChCl:Gly,92 and ChCl:U:Gly38 and is consistent with extended C12 chain lengths;89 however, the elongation of the micelles is much greater in this case. For ionic liquids and DES, the surfactant–solvent counterion interaction is used to explain the morphology of micellar structures, such as the elongation of SDS micelles in aqueous solutions and other solvents. For SDS, spherical micelles are observed in water;90 however, in salt + water, extremely long micelles are observed,93 and in ChCl:U, cylindrical micelles with length ∼100 Å are observed. The elongation of the micelles is attributed to Coulombic interactions between the salt ions or DES [cholinium]+ with the surfactant SO4− headgroups, which provides a charge screening effect, allowing for the headgroups to pack more closely and causing micelle elongation.94 In our case for Ce:U DES, the [Ce–NO3–]+ polycations that form the structural network in the DES interact with the SO4− headgroup of the SDS and provide the charge screening effect. The ionic nature of the DES components offers greater charge screening than in the case of type III DES and leads to micelles that are longer in size. The salting out effects and strong interactions of the [Ce–NO3–]+ polycations with the SO4− head group also cause SDS to form a gel when dissolved in the DES at concentrations >1 wt. %. Formation of a gel phase was also reported by Matthews et al. in ChCl:Gly DES at room temperature.95 Microscopy, SANS, and rheology studies by the group indicate self-assembled SDS fractal dendrites with dimensions of up to ∼mm in ChCl:Gly, which lead to the formation of a gel phase at SDS concentration above 1.9 wt. %. The formation of these aggregate morphologies is attributed to the condensation of choline ions around the SDS aggregates. In our case, it could be the condensation of the [Ce–NO3–]+ polycations around the SDS aggregates that results in the formation of a gel.

3. C12EO6 and C16EO8 in Ce:U

SANS was measured from 1 wt. % solutions of C12EO6 and C16EO8 in Ce:U for hh-DES and hd-DES contrasts and is shown in Fig. 4. The data show the scattering characteristic of spherical micelles; therefore, the two SANS contrasts for each surfactant were fitted to a spherical form factor and its core–shell variant (see the supplementary material, Fig. S5 and Table S5) with the alkyl chains forming the core-radius and the polyethoxy headgroup forming the shell of the structure. The shell SLD was once again allowed to vary to account for solvation of the polyethoxy head group by the DES components. The core–shell spherical model gives a significantly improved fit in this case; reduced χ2 = 5.36 for C12EO6 and χ2 = 3.6 for C16EO8 as compared to 9.21 and 7.1, respectively, for the homogenous spherical model (see the supplementary material, Table S5). Therefore, this is considered a better representation of the structure of these non-ionic surfactants in Ce:U DES and is presented in Fig. 4, and the fit parameters are given in Table V. In Table V, the core-radius and shell thickness for the C12EO6 and C16EO8 micelles obtained from the combined fits using both contrasts are reported along with the SLD of the shell at the two contrasts and the calculated solvent volume fraction of the DES in the shell region (ϕ).

FIG. 4.

SANS patterns for 1 wt. % (a) C12EO6 and (b) C16EO8 in Ce:U at two contrasts: orange circles represent h-surfactant in Ce:U (bottom) and green triangles represent h-surfactant in Ce:d-U (top). The various traces are offset along the y axis for clarity. The black dashed lines represent the best fits obtained to the core–shell spherical form factor through co-refinement of all the contrasts.

FIG. 4.

SANS patterns for 1 wt. % (a) C12EO6 and (b) C16EO8 in Ce:U at two contrasts: orange circles represent h-surfactant in Ce:U (bottom) and green triangles represent h-surfactant in Ce:d-U (top). The various traces are offset along the y axis for clarity. The black dashed lines represent the best fits obtained to the core–shell spherical form factor through co-refinement of all the contrasts.

Close modal
TABLE V.

Core–shell spherical form factor fit parameters for the SANS data from 1 wt. % C12EO6 and C16EO8 in Ce:U shown in Fig. 4 at the different contrasts.

SurfactantCore-radius (Å)Shell thickness (Å)ContrastsSLD shell (Å × 10−6)ϕaNaggb
C12EO6 18.7 ± 1.8 7.2 ± 1.0 h-C12EO6 in Ce:U 1.50 ± 0.27 0.55 ± 0.10 69 ± 20 
   h-C12EO6 in Ce:d-U 2.20 ± 0.66 0.39 ± 0.12  
C16EO8 23.5 ± 1.5 7.9 ± 0.9 h-C16EO8 in Ce:U 1.42 ± 0.24 0.50 ± 0.08 107 ± 21 
   h-C18EO8 in Ce:d-U 2.19 ± 0.55 0.39 ± 0.10  
SurfactantCore-radius (Å)Shell thickness (Å)ContrastsSLD shell (Å × 10−6)ϕaNaggb
C12EO6 18.7 ± 1.8 7.2 ± 1.0 h-C12EO6 in Ce:U 1.50 ± 0.27 0.55 ± 0.10 69 ± 20 
   h-C12EO6 in Ce:d-U 2.20 ± 0.66 0.39 ± 0.12  
C16EO8 23.5 ± 1.5 7.9 ± 0.9 h-C16EO8 in Ce:U 1.42 ± 0.24 0.50 ± 0.08 107 ± 21 
   h-C18EO8 in Ce:d-U 2.19 ± 0.55 0.39 ± 0.10  
a

ϕ is the solvent volume fraction in the shell calculated from the SLD of the shell (polyethoxy head group) and solvent at that contrast.

b

Nagg is the aggregation number calculated by dividing the micelle core volume by the tail volume calculated from the Tanford equation.89 

The radius of the micelles was found to be 18.7 ± 1.8 Å for C12EO6 and 23.5 ± 1.5 Å for C16EO8, and the shell thickness was found to be 7.2 ± 1.0 Å for C12EO6 and 7.8 ± 0.9 Å for C16EO8. The aggregation number was calculated to be 69 ± 20 for C12EO6 and 107 ± 21 for C16EO8, giving an area per molecule of 64 ± 18 and 65 ± 12, respectively, in good agreement with the values obtained from the surface tension measurements. The fitted core-radius for both C12EO6 and C16EO8 micelles while slightly on the higher side is consistent with fully extended C12 and C16 chain lengths,89 respectively, within the fit error. Micelles of core-radius of up to 18.1 Å have been reported previously for C12EOn micelles in ionic liquid ethylammonium nitrate.24 The larger size could imply that some of the “headgroup” is in the core due to poor solvation in these liquids, also manifested in the fact that the headgroup region (shell thickness) is smaller than the size of the polyethoxy head group for both C12EO6 and C16EO8. The solvent volume fraction in the shell region, calculated from the SLD of the shell, is 40%–55%, lower than the solvation of ionic micelles. The smaller shell thickness may be due to loss of scattering contrast beyond a certain distance resulting from the solvation of the shell region by the DES or coiling of the polyethoxy head group in the DES96 compared to water. In this case, the coiling of the polyethoxy head group could be due to the presence of a large number of ions in the DES97 or due to the fact that the interactions of the polyethoxy head group with the DES are less solvophilic as compared to that in water because of the presence of the organic species, an effect also observed for C12EOx micelles in ethylammonium nitrate.24 The overall size of the C12EO6 micelle (∼2.6 nm) is smaller than that observed for water (∼3.2 nm)98 but larger than that observed in ionic liquids (1.5–2.0 nm).24–27 A similar observation can be made for C16EO8 where the micelle size (∼3.2 nm) is smaller than water (∼3.7 nm)98 but comparable to some ionic liquids (∼3.2 nm).26 The solvophobic interactions influence not only the CMC, as previously discussed, but also the size of the micelles. The urea in the DES along with the ionic components results in solvophobic interactions for non-ionic surfactants in Ce:U that are intermediate compared to the ones observed for water and polar organic solvents/ionic liquids, resulting in both sizes of micelles and CMCs, which lie in between the two.

In addition to scattering data from low surfactant concentration solutions, SANS was also measured from h-C12TANO3 and h-C16EO8 for concentrations ranging from 0.5 to 10 wt. % in hh-DES and hd-DES to investigate the effect of concentration on the surfactant micellar structure. The data are shown in Fig. 5. It is not possible to do a similar concentration study for SDS due to the formation of a gel at concentrations above 1 wt. %.

FIG. 5.

SANS patterns and best fits for 0.5–10 wt. % C12TANO3 and C16EO8 in Ce:U: (a) h-C12TANO3 in hh-DES, (b) h-C12TANO3 in hd-DES, (c) h-C16EO8 in hh-DES, and (d) h-C16EO8 in hd-DES. Red circles represent 10 wt. %, orange upward triangles represent 7.5 wt. %, mustard downward triangles represent 5 wt. %, green leftward triangles represent 3 wt. %, cyan rightward triangles represent 1 wt. %, and blue squares represent 0.5 wt. %. The various traces are offset along the y axis for clarity.

FIG. 5.

SANS patterns and best fits for 0.5–10 wt. % C12TANO3 and C16EO8 in Ce:U: (a) h-C12TANO3 in hh-DES, (b) h-C12TANO3 in hd-DES, (c) h-C16EO8 in hh-DES, and (d) h-C16EO8 in hd-DES. Red circles represent 10 wt. %, orange upward triangles represent 7.5 wt. %, mustard downward triangles represent 5 wt. %, green leftward triangles represent 3 wt. %, cyan rightward triangles represent 1 wt. %, and blue squares represent 0.5 wt. %. The various traces are offset along the y axis for clarity.

Close modal

As can be seen from the appearance of a shoulder-like feature at q ∼ 0.1 Å−1, in the SANS data for solutions above 3 wt. %, intermicellar interactions are observed in the scattering pattern. Therefore, these data were fitted using a core–shell form factor multiplied with a hard sphere structure factor with a radius independent of the micelle size. This is a methodology applied to micelles in DES so as to take into account interactions that may be mediated via structured solvent.21,38 The core-radius, shell thickness, and SLD of the shell for the micelle were fixed at the values obtained from the fit at 1 wt. %. The effective radius was co-fitted for the two contrasts and various concentrations (>3 wt. %) with the volume fraction varied to account for the change in concentration. The fits obtained are shown as the dashed lines in Fig. 5 with the calculated and fitted structure factor volume fractions summarized in the supplementary material, Table S6. For both surfactant concentration series, the above model provides reasonable fits with the fit volume fraction showing good agreement with the calculated ones. This suggests that the surfactant micellar structure (shape and size) is not affected by concentration and a structure factor with a radius independent of the concentration can explain the intermicellar interaction. In order to verify this, the concentrations were each fitted to individual radii. De-constraining the fits resulted in radii, which were within 2% of the value obtained for the combined fit and did not result in any significant improvement to the fit. Therefore, a single radius and varying volume fractions between different concentrations are considered the best suited model to describe the intermicellar interaction contribution to these scattering patterns.

The value of the interaction radius obtained from the structure factor fits at high concentration (>3 wt. %) for C12TANO3 was found to be 22.4 ± 0.1 Å. Even though no fitting bounds were applied to the intermicellar interaction radius, this value is consistent with the size of the C12TANO3 micelles, which was found to be 22.0 ± 1.1 Å from the fits for 1 wt. % data at multiple contrasts. This is analogous to the approach used for CnTAB micelles in ChCl:Gly21 and ChCl:U:Gly38 DES where the effective interaction radius with the same second virial coefficient as the scatterer is used in the case of ellipsoidal micelles. Here, we have additionally demonstrated the equivalence of the two radii by obtaining a consistent value for the two despite, allowing the interaction radius to vary.

A similar approach as the one above applied to C12TANO3 gives an interaction radius of 36.6 ± 0.1 Å for the high concentration (>3 wt. %) C16EO8 micelles. This is slightly larger than the micelle size, 31.4 ± 1.8 Å, obtained from 1 wt. % C16EO8 in Ce:U. The difference can be explained by the fact that the size of the micelle is potentially an underestimate due to the solvation of the shell region by the DES resulting in loss of scattering contrast beyond a certain distance, manifested also in the shell thickness for the C16EO8 micelles being smaller than the size of the polyethoxy (EO8) head group expected in pure water. Assuming that the shell thickness obtained from the fitting of multiple contrast at 1 wt. % is an underestimate, the actual shell thickness would be 13.1 ± 1.2 Å as opposed to 7.9 ± 0.9 Å. Another reason for the higher interaction radius could be that the interaction is mediated by structured solvent between the micelles, a more likely explanation given we suspect polyethoxy chains to be collapsed in the core due to less favorable interactions with the solvent.

The type IV DES studied in this work, Ce(NO3)3·6H2O:3.5*Urea (Ce:U), has unique characteristics and physicochemical properties, such as high density, small molar volume, low viscosity, and high surface tension when compared to typical type III DESs. This results in the solvent having a Gordon parameter comparable to, even slightly higher than, that of water. The high Gordon parameter along with the highly ionic nature of the solvent results in self-assembly of a range of surfactants in this DES.

The cationic (C12TAB and C12TANO3), anionic (SDS), and non-ionic (C12EO6 and C16EO8) surfactants studied in this work remain surface-active in the Ce:U DES as shown by the surface tension data. They form micelles at concentrations above the CMCs, as determined by changes in the DES/air surface tension upon increasing surfactant concentration, as summarized in Table II. The CMCs of cationic and anionic surfactants, C12TAB, C12TANO3, and SDS, in Ce:U DESs are lower than that observed in water18,68,99 and ionic liquids.19,23,28 This is attributed to the presence of highly charged ions in the DES along with the high Gordon parameter. For C12TA+ surfactants, this results in a CMC, which is even lower than that observed in type III DESs.21,22,77 In the case of non-ionic surfactants, the CMC lies in between that of water70,72 and ionic liquids.19,25,26 We attribute this to a combined effect of the organic and highly ionic character of the solvent, which offers some reduction in the solvophobic interactions between the surfactant and the DES but not to the same extent as that in ionic liquids or polar organic solvents, such as formamide.

The structure of the surfactant micelles at concentrations above the CMC was investigated using Small Angle Neutron Scattering (SANS). The scattering data show the presence of micelles with a core–shell structure that comprise the alkyl tail core and the trimethylammonium/sulfate/polyethoxy head group shell. The size and shape of the micelles are summarized in Tables IIIV. The scattering curves for dodecyl surfactants and a visualization of the micellar structures are shown in Fig. 6. The core-radius in each case is consistent with the corresponding C12/C16 chain lengths and is in good agreement with values obtained for these surfactants in other solvents. The shell region, on the other hand, comprises highly solvated headgroups, and co-fitting the various contrasts gives an indication of the dominant interaction mechanism for the particular surfactant. For cationic and anionic surfactants, the interaction is primarily electrostatic modulated by counterion binding, as is the case for ionic surfactants in water. However, in the case of non-ionic surfactants, in addition to the ions, urea in the DES offers polar organic interactions resulting in a different pathway to surfactant morphology modulation. The micelles are spherical in the case of cationic (C12TAB and C12TANO3) and non-ionic (C12EO6 and C16EO8) surfactants and cylindrical in the case of the anionic surfactant SDS.

FIG. 6.

(a) SANS patterns and best fits for 1 wt. % C12TANO3 (red circles), SDS (green downward triangles), and C12EO6 (blue upward triangles) in Ce:d-U. (b) Visualization of the micellar cross sections: spherical for C12TANO3 and C12EO6 and cylindrical for SDS. The dodecyl chains are shown in gray, the headgroups are shown in red for C12TANO3, blue for C12EO6, and green for SDS, and the solvent molecules are shown in orange.

FIG. 6.

(a) SANS patterns and best fits for 1 wt. % C12TANO3 (red circles), SDS (green downward triangles), and C12EO6 (blue upward triangles) in Ce:d-U. (b) Visualization of the micellar cross sections: spherical for C12TANO3 and C12EO6 and cylindrical for SDS. The dodecyl chains are shown in gray, the headgroups are shown in red for C12TANO3, blue for C12EO6, and green for SDS, and the solvent molecules are shown in orange.

Close modal

For surfactant/ionic liquid solutions, it has been shown that a surfactant–solvent counterion exchange can occur in the system, where surfactant counterions are incorporated into the bulk solvent, while ions in the solvent interact with the micelle surface.18 As has been observed for aqueous systems,100 the morphology of ionic surfactant micelles in type III DES is strongly dependent on the interaction of the surfactant headgroup and counterion with the solvent.20–22,38,94 These types of interactions can also explain the morphology of ionic surfactant micelles in the type IV Ce:U DES. For SDS, Coulombic interactions between the DES [–Ce–NO3–]+ polycations [Ce3(NO3)72+]44 with the surfactant SO4− headgroups provide a charge screening effect, which allows for the headgroups to pack more closely, causing elongation of the micelles, whereas in the case of C12TA+, spherical/spheroidal rather than elongated micelles are observed due to the presence of nitrate/bromide ions within the DES, which interact with the positively charged headgroups but do not cause any significant disruption to the micelle shape for C12TA+ micelles. This effect is also observed for C12TA+ surfactants in water where spherical micelles are observed until extremely high concentrations are reached.101 Little difference in micelles size and shape is observed between C12TAB and C12TANO3 in the Ce:U DES; Br and NO3 are next to each other in the Hofmeister series and give approximately equivalent micelle properties when binding to micelles in water.102 Similar observations of micelle morphology have also been made for SDS and CnTAB micelles in ChCl-based DES, where the interacting ions are [cholinium]+ and chloride/bromide.20,21,38,94

SANS studies show no effect of surfactant concentration on C12TANO3 and C16EO8 micelle structure (shape or size) in Ce:U, apart from an intermicellar interaction contribution to the scattering for concentrations >3 wt. %. The intermicellar interaction radius was found to be in good agreement with the micelle size for both C12TANO3 and C16EO8, despite being allowed to fit independently. This is in contrast to studies on cationic surfactants in ChCl:Gly21 and ChCl:U:Gly38 where small differences in micelle structure with surfactant concentration are observed. The authors attribute this to small fluctuations in the fit arising from the fitting procedure. In this work, each individual concentration was not fitted to a different core-radius and shell thickness, as was the case for studies in ChCl:Gly or ChCl:U:Gly, but instead, the values were fixed at those obtained from 1 wt. % surfactant co-fitted using multiple contrast and only the interaction radius was co-fitted for the different concentrations.

Self-assembly of cationic, C12TAB and C12TANO3, anionic, SDS, and non-ionic, C12EO6 and C16EO8, surfactants in the type IV DES comprising metal salt, cerium(III) nitrate hexahydrate, and hydrogen bond donor, urea, in the molar ratio of 1:3.5 is investigated using interfacial tension and SANS measurements. The Gordon parameter for the DES is comparable to that of water, which drives micellization for a range of surfactants. Surface tension measurements have shown that the CMC of ionic surfactants in this DES is less than that of water, whereas the CMC of non-ionic surfactants is higher than that of water. These studies show that both the ionic character of the DES, which drives interactions in the case of ionic surfactants, and the presence of organic moieties in the DES, which along with the ions affect interactions with non-ionic surfactants, play a role in micellization. The structure and morphology of surfactant micelles were investigated using contrast variation SANS. The results show the formation of core–shell micelles with a spherical (C12TAB, C12TANO3, C12EO6, and C16EO8) or cylindrical geometry (SDS) depending on the nature and strength of interaction of the surfactant headgroups with the solvent components. Counterion binding plays an important role in the morphology modulation in the case of ionic surfactants; [–Ce–NO3–]+ polycations bind to the SO4− headgroups in SDS leading to charge screening and close packing of the headgroup resulting in cylindrical micelles, whereas the nitrate/bromide ions that bind to the cationic surfactant headgroup do not cause any significant disruption in headgroup packing or micelle shape, resulting in spherical micelles. On the other hand, for non-ionic surfactants, both the organic (urea) and ionic [Ce(NO3)3] components play a role in micellization leading to a CMC and structure that lies between that in aqueous solutions and organic polar solvent/ionic liquids. To the best of our knowledge, this is the first reported work on micellization of non-ionic surfactants in a deep eutectic solvent. The results of this study have provided an insight into the mechanism of micelle formation in these novel highly ionic but hydrogen-bonded solvents, leading to an important understanding about the interaction of DES components with amphiphiles and how this influences their self-assembly. This raises the opportunity to use these soft structures as liquid crystalline templates for generating porosity in synthesis applications.

See the supplementary material for characterization data for the Ce:U DES (DSC data, viscosity measurements with temperature, and glass transition temperatures calculated from viscosity); scattering length densities for the Ce:U DES, its components, and surfactant components used to fit SANS data; model fitting trials and details of parameters from the fitting of the SANS data; surface tension measurements and the calculated critical micellar concentrations in different surfactant concentration units (mM, wt% and mole fraction).

I.M. acknowledges funding from EPSRC (Grant No. EP/S020772/1). M.R.A. acknowledges EPSRC for funding (Grant No. EP/S021019/1). R.S.A. acknowledges the EPSRC Centre for Doctoral Training in Sustainable Chemical Technologies for funding (Grant No. EP/L016354/1), and J.H. acknowledges the University of Bath for Ph.D. funding. The authors acknowledge the ISIS Neutron and Muon Source for neutron beamtime (Experiment No. RB1920676). This work benefited from the use of the SasView application, originally developed under NSF Award No. DMR-0520547. SasView contains the code developed with funding from the European Union’s Horizon 2020 research and innovation program under the SINE2020 project, Grant No. 654000.

The authors declare no conflicts of interest.

The data that support the findings of this study are openly available in the Bath research data archive system at https://doi.org/10.15125/BATH-01049.

1.
M.
Antonietti
,
D.
Kuang
,
B.
Smarsly
, and
Y.
Zhou
, “
Ionic liquids for the convenient synthesis of functional nanoparticles and other inorganic nanostructures
,”
Angew. Chem., Int. Ed.
43
,
4988
4992
(
2004
).
2.
C.
Janiak
, “
Metal nanoparticle synthesis in ionic liquids
,” in
Ionic Liquids (ILs) in Organometallic Catalysis
, edited by
J.
Dupont
and
L.
Kollár
(
Springer Berlin Heidelberg
,
Berlin, Heidelberg
,
2015
), pp.
17
53
.
3.
K.
Richter
,
P. S.
Campbell
,
T.
Baecker
,
A.
Schimitzek
,
D.
Yaprak
, and
A.-V.
Mudring
, “
Ionic liquids for the synthesis of metal nanoparticles
,”
Phys. Status Solidi B
250
,
1152
1164
(
2013
).
4.
H.-G.
Liao
,
Y.-X.
Jiang
,
Z.-Y.
Zhou
,
S.-P.
Chen
, and
S.-G.
Sun
, “
Shape-controlled synthesis of gold nanoparticles in deep eutectic solvents for studies of structure–functionality relationships in electrocatalysis
,”
Angew. Chem., Int. Ed.
47
,
9100
9103
(
2008
).
5.
O. S.
Hammond
,
S.
Eslava
,
A. J.
Smith
,
J.
Zhang
, and
K. J.
Edler
, “
Microwave-assisted deep eutectic-solvothermal preparation of iron oxide nanoparticles for photoelectrochemical solar water splitting
,”
J. Mater. Chem. A
5
,
16189
16199
(
2017
).
6.
O. S.
Hammond
,
K. J.
Edler
,
D. T.
Bowron
, and
L.
Torrente-Murciano
, “
Deep eutectic-solvothermal synthesis of nanostructured ceria
,”
Nat. Commun.
8
,
14150
(
2017
).
7.
A. P.
Abbott
,
J. C.
Barron
,
G.
Frisch
,
K. S.
Ryder
, and
A. F.
Silva
, “
The effect of additives on zinc electrodeposition from deep eutectic solvents
,”
Electrochim. Acta
56
,
5272
5279
(
2011
).
8.
A. P.
Abbott
,
K. E.
Ttaib
,
G.
Frisch
,
K. S.
Ryder
, and
D.
Weston
, “
The electrodeposition of silver composites using deep eutectic solvents
,”
Phys. Chem. Chem. Phys.
14
,
2443
2449
(
2012
).
9.
P.
Sebastian
,
M. I.
Giannotti
,
E.
Gómez
, and
J. M.
Feliu
, “
Surface sensitive nickel electrodeposition in deep eutectic solvent
,”
ACS Appl. Energy Mater.
1
,
1016
1028
(
2018
).
10.
A. P.
Abbott
,
G.
Frisch
, and
K. S.
Ryder
, “
Electroplating using ionic liquids
,”
Annu. Rev. Mater. Res.
43
,
335
358
(
2013
).
11.
S.
Caporali
,
A.
Fossati
,
A.
Lavacchi
,
I.
Perissi
,
A.
Tolstogouzov
, and
U.
Bardi
, “
Aluminium electroplated from ionic liquids as protective coating against steel corrosion
,”
Corros. Sci.
50
,
534
539
(
2008
).
12.
L.
Sun
and
J. F.
Brennecke
, “
Characterization of imidazolium chloride ionic liquids plus trivalent chromium chloride for chromium electroplating
,”
Ind. Eng. Chem. Res.
54
,
4879
4890
(
2015
).
13.
S.
Schaltin
,
N. R.
Brooks
,
K.
Binnemans
, and
J.
Fransaer
, “
Electrodeposition from cationic cuprous organic complexes: Ionic liquids for high current density electroplating
,”
J. Electrochem. Soc.
158
,
D21
(
2011
).
14.
K. D.
Weaver
,
H. J.
Kim
,
J.
Sun
,
D. R.
MacFarlane
, and
G. D.
Elliott
, “
Cyto-toxicity and biocompatibility of a family of choline phosphate ionic liquids designed for pharmaceutical applications
,”
Green Chem.
12
,
507
513
(
2010
).
15.
S. N.
Pedro
,
M. G.
Freire
,
C. S. R.
Freire
, and
A. J. D.
Silvestre
, “
Deep eutectic solvents comprising active pharmaceutical ingredients in the development of drug delivery systems
,”
Expert Opin. Drug Delivery
16
,
497
506
(
2019
).
16.
M.
Wang
,
S.
Fang
, and
X.
Liang
, “
Natural deep eutectic solvents as eco-friendly and sustainable dilution medium for the determination of residual organic solvents in pharmaceuticals with static headspace-gas chromatography
,”
J. Pharm. Biomed. Anal.
158
,
262
268
(
2018
).
17.
C. R.
López-Barrón
and
N. J.
Wagner
, “
Structural transitions of CTAB micelles in a protic ionic liquid
,”
Langmuir
28
,
12722
12730
(
2012
).
18.
A.
Dolan
,
R.
Atkin
, and
G. G.
Warr
, “
The origin of surfactant amphiphilicity and self-assembly in protic ionic liquids
,”
Chem. Sci.
6
,
6189
6198
(
2015
).
19.
T. L.
Greaves
and
C. J.
Drummond
, “
Solvent nanostructure, the solvophobic effect and amphiphile self-assembly in ionic liquids
,”
Chem. Soc. Rev.
42
,
1096
1120
(
2013
).
20.
T.
Arnold
,
A. J.
Jackson
,
A.
Sanchez-Fernandez
,
D.
Magnone
,
A. E.
Terry
, and
K. J.
Edler
, “
Surfactant behavior of sodium dodecylsulfate in deep eutectic solvent choline chloride/urea
,”
Langmuir
31
,
12894
12902
(
2015
).
21.
A.
Sanchez-Fernandez
,
T.
Arnold
,
A. J.
Jackson
,
S. L.
Fussell
,
R. K.
Heenan
,
R. A.
Campbell
, and
K. J.
Edler
, “
Micellization of alkyltrimethylammonium bromide surfactants in choline chloride:glycerol deep eutectic solvent
,”
Phys. Chem. Chem. Phys.
18
,
33240
33249
(
2016
).
22.
A.
Sanchez-Fernandez
,
O. S.
Hammond
,
A. J.
Jackson
,
T.
Arnold
,
J.
Doutch
, and
K. J.
Edler
, “
Surfactant–solvent interaction effects on the micellization of cationic surfactants in a carboxylic acid-based deep eutectic solvent
,”
Langmuir
33
,
14304
14314
(
2017
).
23.
J. L.
Anderson
,
V.
Pino
,
E. C.
Hagberg
,
V. V.
Sheares
, and
D. W.
Armstrong
, “
Surfactant solvation effects and micelle formation in ionic liquids
,”
Chem. Commun.
2003
,
2444
2445
.
24.
M. U.
Araos
and
G. G.
Warr
, “
Structure of nonionic surfactant micelles in the ionic liquid ethylammonium nitrate
,”
Langmuir
24
,
9354
9360
(
2008
).
25.
T. L.
Greaves
,
S. T.
Mudie
, and
C. J.
Drummond
, “
Effect of protic ionic liquids (PILs) on the formation of non-ionic dodecyl poly(ethylene oxide) surfactant self-assembly structures and the effect of these surfactants on the nanostructure of PILs
,”
Phys. Chem. Chem. Phys.
13
,
20441
20452
(
2011
).
26.
C.
Patrascu
,
F.
Gauffre
,
F.
Nallet
,
R.
Bordes
,
J.
Oberdisse
,
N.
de Lauth-Viguerie
, and
C.
Mingotaud
, “
Micelles in ionic liquids: Aggregation behavior of alkyl poly(ethyleneglycol)-ethers in 1-butyl-3-methyl-imidazolium type ionic liquids
,”
ChemPhysChem
7
,
99
101
(
2006
).
27.
T.
Misono
,
H.
Sakai
,
K.
Sakai
,
M.
Abe
, and
T.
Inoue
, “
Surface adsorption and aggregate formation of nonionic surfactants in a room temperature ionic liquid, 1-butyl-3-methylimidazolium hexafluorophosphate (bmimPF6)
,”
J. Colloid Interface Sci.
358
,
527
533
(
2011
).
28.
D. F.
Evans
,
A.
Yamauchi
,
R.
Roman
, and
E. Z.
Casassa
, “
Micelle formation in ethylammonium nitrate, a low-melting fused salt
,”
J. Colloid Interface Sci.
88
,
89
96
(
1982
).
29.
J.
Lu
,
F.
Yan
, and
J.
Texter
, “
Advanced applications of ionic liquids in polymer science
,”
Prog. Polym. Sci.
34
,
431
448
(
2009
).
30.
D. F.
Mohshim
,
H.
Mukhtar
, and
Z.
Man
, “
Ionic liquid polymeric membrane: Synthesis, characterization and performance evaluation
,” in
Advanced Materials Engineering and Technology II
, Key Engineering Materials Vol. 594 (
Trans Tech Publications Ltd.
,
2014
), pp.
18
23
.
31.
E.
Rynkowska
,
K.
Fatyeyeva
, and
W.
Kujawski
, “
Application of polymer-based membranes containing ionic liquids in membrane separation processes: A critical review
,”
Rev. Chem. Eng.
34
,
341
363
(
2018
).
32.
X.
Li
,
J.
Choi
,
W.-S.
Ahn
, and
K. H.
Row
, “
Preparation and application of porous materials based on deep eutectic solvents
,”
Crit. Rev. Anal. Chem.
48
,
73
85
(
2018
).
33.
E. D.
Mohamed Isa
,
M. B.
Abdul Rahman
, and
H.
Ahmad
, “
Monodispersed mesoporous silica nanospheres based on pyridinium ionic liquids
,”
J. Porous Mater.
25
,
1439
(
2018
).
34.
N.
Rameli
,
K.
Jumbri
,
R. A.
Wahab
,
A.
Ramli
, and
F.
Huyop
, “
Synthesis and characterization of mesoporous silica nanoparticles using ionic liquids as a template
,”
J. Phys.: Conf. Ser.
1123
,
012068
(
2018
).
35.
L.
Hu
,
Z.
Yan
,
J.
Zhang
,
X.
Peng
,
X.
Mo
,
A.
Wang
, and
L.
Chen
, “
Surfactant aggregates within deep eutectic solvent-assisted synthesis of hierarchical ZIF-8 with tunable porosity and enhanced catalytic activity
,”
J. Mater. Sci.
54
,
11009
11023
(
2019
).
36.
A.
Banerjee
,
K.
Ibsen
,
Y.
Iwao
,
M.
Zakrewsky
, and
S.
Mitragotri
, “
Transdermal protein delivery using choline and geranate (CAGE) deep eutectic solvent
,”
Adv. Healthcare Mater.
6
,
1601411
(
2017
).
37.
A. P.
Abbott
,
G.
Capper
,
D. L.
Davies
,
R. K.
Rasheed
, and
V.
Tambyrajah
, “
Novel solvent properties of choline chloride/urea mixtures
,”
Chem. Commun.
2003
,
70
71
.
38.
R. S.
Atri
,
A.
Sanchez-Fernandez
,
O. S.
Hammond
,
I.
Manasi
,
J.
Doutch
,
J. P.
Tellam
, and
K. J.
Edler
, “
Morphology modulation of ionic surfactant micelles in ternary deep eutectic solvents
,”
J. Phys. Chem. B
124
,
6004
6014
(
2020
).
39.
S. J.
Bryant
,
R.
Atkin
, and
G. G.
Warr
, “
Effect of deep eutectic solvent nanostructure on phospholipid bilayer phases
,”
Langmuir
33
,
6878
6884
(
2017
).
40.
L.
Han
and
S.
Che
, “
Anionic surfactant templated mesoporous silicas (AMSs)
,”
Chem. Soc. Rev.
42
,
3740
3752
(
2013
).
41.
F.
Gai
,
T.
Zhou
,
G.
Chu
,
Y.
Li
,
Y.
Liu
,
Q.
Huo
, and
F.
Akhtar
, “
Mixed anionic surfactant-templated mesoporous silica nanoparticles for fluorescence detection of Fe3+
,”
Dalton Trans.
45
,
508
514
(
2016
).
42.
H.
Luo
,
Q.
Lin
,
S.
Baber
, and
M.
Naalla
, “
Surfactant-templated mesoporous metal oxide nanowires
,”
J. Nanomater.
2010
,
750960
.
43.
E. L.
Smith
,
A. P.
Abbott
, and
K. S.
Ryder
, “
Deep eutectic solvents (DESs) and their applications
,”
Chem. Rev.
114
,
11060
11082
(
2014
).
44.
O. S.
Hammond
,
D. T.
Bowron
, and
K. J.
Edler
, “
Structure and properties of ‘type IV’ lanthanide nitrate hydrate:urea deep eutectic solvents
,”
ACS Sustainable Chem. Eng.
7
,
4932
4940
(
2019
).
45.
D. F.
Evans
, “
Self-organization of amphiphiles
,”
Langmuir
4
,
3
12
(
1988
).
46.
M.
Nyden
and
O.
Soederman
, “
Structures and emulsification failure in the microemulsion phase in the didodecyldimethylammonium sulfate/hydrocarbon/water system. A self-diffusion NMR study
,”
Langmuir
11
,
1537
1545
(
1995
).
47.
J. D.
Berry
,
M. J.
Neeson
,
R. R.
Dagastine
,
D. Y. C.
Chan
, and
R. F.
Tabor
, “
Measurement of surface and interfacial tension using pendant drop tensiometry
,”
J. Colloid Interface Sci.
454
,
226
237
(
2015
).
48.
R. K.
Heenan
,
S. E.
Rogers
,
D.
Turner
,
A. E.
Terry
,
J.
Treadgold
, and
S. M.
King
, “
Small angle neutron scattering using Sans2d
,”
Neutron News
22
,
19
21
(
2011
).
49.
I.
Manasi
,
M. R.
Andalibi
,
R.
Atri
,
J.
Hooton
,
K. J.
Edler
, and
S. M.
King
, (
STFC ISIS Neutron and Muon Source
,
2020
).
50.
O.
Arnold
,
J. C.
Bilheux
,
J. M.
Borreguero
,
A.
Buts
,
S. I.
Campbell
,
L.
Chapon
,
M.
Doucet
,
N.
Draper
,
R.
Ferraz Leal
,
M. A.
Gigg
,
V. E.
Lynch
,
A.
Markvardsen
,
D. J.
Mikkelson
,
R. L.
Mikkelson
,
R.
Miller
,
K.
Palmen
,
P.
Parker
,
G.
Passos
,
T. G.
Perring
,
P. F.
Peterson
,
S.
Ren
,
M. A.
Reuter
,
A. T.
Savici
,
J. W.
Taylor
,
R. J.
Taylor
,
R.
Tolchenov
,
W.
Zhou
, and
J.
Zikovsky
, “
Mantid—Data analysis and visualization package for neutron scattering and μ SR experiments
,”
Nucl. Instrum. Methods Phys. Res., Sect. A
764
,
156
166
(
2014
).
51.
S. R.
Kline
, “
Reduction and analysis of SANS and USANS data using IGOR Pro
,”
J. Appl. Crystallogr.
39
,
895
900
(
2006
).
52.
See http://sasview.org for SasView for Small Angle Scattering Analysis.
53.
J. K.
Percus
and
G. J.
Yevick
, “
Analysis of classical statistical mechanics by means of collective coordinates
,”
Phys. Rev.
110
,
1
13
(
1958
).
54.
C. A.
Angell
, “
Formation of glasses from liquids and biopolymers
,”
Science
267
,
1924
1935
(
1995
).
55.
T. L.
Greaves
,
A.
Weerawardena
, and
C. J.
Drummond
, “
Nanostructure and amphiphile self-assembly in polar molecular solvents: Amides and the ‘solvophobic effect
,’”
Phys. Chem. Chem. Phys.
13
,
9180
9186
(
2011
).
56.
G. G.
Warr
and
R.
Atkin
, “
Solvophobicity and amphiphilic self-assembly in neoteric and nanostructured solvents
,”
Curr. Opin. Colloid Interface Sci.
45
,
83
96
(
2020
).
57.
Y.
Xie
,
H.
Dong
,
S.
Zhang
,
X.
Lu
, and
X.
Ji
, “
Effect of water on the density, viscosity, and CO2 solubility in choline chloride/urea
,”
J. Chem. Eng. Data
59
,
3344
3352
(
2014
).
58.
N. B.
Vargaftik
,
B. N.
Volkov
, and
L. D.
Voljak
, “
International tables of the surface tension of water
,”
J. Phys. Chem. Ref. Data
12
,
817
820
(
1983
).
59.
I. I.
Adamenko
,
L. A.
Bulavin
,
V.
Ilyin
,
S. A.
Zelinsky
, and
K. O.
Moroz
, “
Anomalous behavior of glycerol–water solutions
,”
J. Mol. Liq.
127
,
90
92
(
2006
), part of Special Issue: International Conference on Physics of Liquid Matter: Modern Problems.
60.
I.
Rico
and
A.
Lattes
, “
Formamide, a water substitute. 12. Krafft temperature and micelle formation of ionic surfactants in formamide
,”
J. Phys. Chem.
90
,
5870
5872
(
1986
).
61.
T.
Wärnheim
and
A.
Jönsson
, “
Phase diagrams of alkyltrimethylammonium surfactants in some polar solvents
,”
J. Colloid Interface Sci.
125
,
627
633
(
1988
).
62.
Y.
Chen
,
W.
Chen
,
L.
Fu
,
Y.
Yang
,
Y.
Wang
,
X.
Hu
,
F.
Wang
, and
T.
Mu
, “
Surface tension of 50 deep eutectic solvents: Effect of hydrogen-bonding donors, hydrogen-bonding acceptors, other solvents, and temperature
,”
Ind. Eng. Chem. Res.
58
,
12741
12750
(
2019
).
63.
M.
Tariq
,
M. G.
Freire
,
B.
Saramago
,
J. A. P.
Coutinho
,
J. N. C.
Lopes
, and
L. P. N.
Rebelo
, “
Surface tension of ionic liquids and ionic liquid solutions
,”
Chem. Soc. Rev.
41
,
829
868
(
2012
).
64.
C. C.
Ruiz
,
L.
Díaz‐López
, and
J.
Aguiar
, “
Micellization of sodium dodecyl sulfate in glycerol aqueous mixtures
,”
J. Dispersion Sci. Technol.
29
,
266
273
(
2008
).
65.
P. H.
Elworthy
and
K. J.
Mysels
, “
The surface tension of sodium dodecylsulfate solutions and the phase separation model of micelle formation
,”
J. Colloid Interface Sci.
21
,
331
347
(
1966
).
66.
J.
Mata
,
D.
Varade
, and
P.
Bahadur
, “
Aggregation behavior of quaternary salt based cationic surfactants
,”
Thermochim. Acta
428
,
147
155
(
2005
).
67.
H.
Hooshyar
and
R.
Sadeghi
, “
Influence of sodium salts on the micellization and interfacial behavior of cationic surfactant dodecyltrimethylammonium bromide in aqueous solution
,”
J. Chem. Eng. Data
60
,
983
992
(
2015
).
68.
H.
Xu
,
P. X.
Li
,
K.
Ma
,
R. K.
Thomas
,
J.
Penfold
, and
J. R.
Lu
, “
Limitations in the application of the Gibbs equation to anionic surfactants at the air/water surface: Sodium dodecylsulfate and sodium dodecylmonooxyethylenesulfate above and below the CMC
,”
Langmuir
29
,
9335
9351
(
2013
).
69.
K.
Ananda
,
O. P.
Yadav
, and
P. P.
Singh
, “
Studies on the surface and thermodynamic properties of some surfactants in aqueous and water+1,4-dioxane solutions
,”
Colloids Surf.
55
,
345
358
(
1991
).
70.
J. M.
Corkill
,
J. F.
Goodman
, and
S. P.
Harrold
, “
Thermodynamics of micellization of non-ionic detergents
,”
Trans. Faraday Soc.
60
,
202
207
(
1964
).
71.
E. C.
Wijaya
,
F.
Separovic
,
C. J.
Drummond
, and
T. L.
Greaves
, “
Micelle formation of a non-ionic surfactant in non-aqueous molecular solvents and protic ionic liquids (PILs)
,”
Phys. Chem. Chem. Phys.
18
,
24377
24386
(
2016
).
72.
J. A.
Kabin
,
S. L.
Tolstedt
,
A. E.
Sáez
,
C. S.
Grant
, and
R. G.
Carbonell
, “
Removal of organic films from rotating disks using aqueous solutions of nonionic surfactants: Effect of surfactant molecular structure
,”
J. Colloid Interface Sci.
206
,
102
111
(
1998
).
73.
T.
Tamura
,
Y.
Takeuchi
, and
Y.
Kaneko
, “
Influence of surfactant structure on the drainage of nonionic surfactant foam films
,”
J. Colloid Interface Sci.
206
,
112
121
(
1998
).
74.
A.
Couper
,
G. P.
Gladden
, and
B. T.
Ingram
, “
Adsorption of surface active agents in a non-aqueous solvent
,”
Faraday Discuss. Chem. Soc.
59
,
63
75
(
1975
).
75.
G.
Andersson
,
E.
Carey
, and
C.
Stubenrauch
, “
Disjoining pressure study of formamide foam films stabilized by surfactants
,”
Langmuir
26
,
7752
7760
(
2010
).
76.
M.
Jonstroemer
,
M.
Sjoeberg
, and
T.
Waernheim
, “
Aggregation and solvent interaction in nonionic surfactant systems with formamide
,”
J. Phys. Chem.
94
,
7549
7555
(
1990
).
77.
I.
Manasi
,
S. J.
Bryant
,
O. S.
Hammond
, and
K. J.
Edler
,
Interactions of Water and Amphiphiles with Deep Eutectic Solvent Nanostructures
(
Academic Press
,
2020
).
78.
G. A.
Cookey
and
C. C.
Obunwo
, “
Effects of sodium bromide salt and temperature on the behaviour of aqueous solution of cetyltrimethylammonium bromide
,”
IOSR J. Appl. Chem.
7
,
34
38
(
2014
).
79.
The effect of temperature, salinity, and alcohol on the critical micelle concentration of surfactants, SPE International Conference on Oilfield Chemistry, Vol. All Days (1991) sPE-21032-MS, https://onepetro.org/SPEOCC/proceedings-pdf/91OCS/All-91OCS/SPE-21032-MS/1999724/spe-21032-ms.pdf.
80.
S.
Chauhan
,
M.
Kaur
,
K.
Kumar
, and
M. S.
Chauhan
, “
Study of the effect of electrolyte and temperature on the critical micelle concentration of dodecyltrimethylammonium bromide in aqueous medium
,”
J. Chem. Thermodyn.
78
,
175
181
(
2014
).
81.
J. E.
Brady
,
D. F.
Evans
,
G. G.
Warr
,
F.
Grieser
, and
B. W.
Ninham
, “
Counterion specificity as the determinant of surfactant aggregation
,”
J. Phys. Chem.
90
,
1853
1859
(
1986
).
82.
M. F.
Emerson
and
A.
Holtzer
, “
On the ionic strength dependence of micelle number. II
,”
J. Phys. Chem.
71
,
1898
1907
(
1967
).
83.
S. S.
Berr
, “
Solvent isotope effects on alkytrimethylammonium bromide micelles as a function of alkyl chain length
,”
J. Phys. Chem.
91
,
4760
4765
(
1987
).
84.
M.
Pal
,
R. K.
Singh
, and
S.
Pandey
, “
Evidence of self-aggregation of cationic surfactants in a choline chloride+glycerol deep eutectic solvent
,”
ChemPhysChem
16
,
2538
2542
(
2015
).
85.
G.
Briganti
,
S.
Puvvada
, and
D.
Blankschtein
, “
Effect of urea on micellar properties of aqueous solutions of nonionic surfactants
,”
J. Phys. Chem.
95
,
8989
8995
(
1991
).
86.
C. L.
Bianco
,
C. S.
Schneider
,
M.
Santonicola
,
A. M.
Lenhoff
, and
E. W.
Kaler
, “
Effects of urea on the microstructure and phase behavior of aqueous solutions of poly(oxyethylene) surfactants
,”
Ind. Eng. Chem. Res.
50
,
85
96
(
2011
).
87.
N.
Nishikido
and
R.
Matuura
, “
The effect of added inorganic salts on the micelle formation of nonionic surfactants in aqueous solutions
,”
Bull. Chem. Soc. Jpn.
50
,
1690
1694
(
1977
).
88.
W.-H.
Xie
,
W.-Y.
Shiu
, and
D.
Mackay
, “
A review of the effect of salts on the solubility of organic compounds in seawater
,”
Mar. Environ. Res.
44
,
429
444
(
1997
).
89.
C.
Tanford
, “
Micelle shape and size
,”
J. Phys. Chem.
76
,
3020
3024
(
1972
).
90.
J. B.
Hayter
and
J.
Penfold
, “
Determination of micelle structure and charge by neutron small-angle scattering
,”
Colloid Polym. Sci.
261
,
1022
1030
(
1983
).
91.
R.
Hargreaves
,
D. T.
Bowron
, and
K.
Edler
, “
Atomistic structure of a micelle in solution determined by wide Q-range neutron diffraction
,”
J. Am. Chem. Soc.
133
,
16524
16536
(
2011
).
92.
A.
Sanchez-Fernandez
,
O. S.
Hammond
,
K. J.
Edler
,
T.
Arnold
,
J.
Doutch
,
R. M.
Dalgliesh
,
P.
Li
,
K.
Ma
, and
A. J.
Jackson
, “
Counterion binding alters surfactant self-assembly in deep eutectic solvents
,”
Phys. Chem. Chem. Phys.
20
,
13952
13961
(
2018
).
93.
M.
Bergström
and
J.
Skov Pedersen
, “
Structure of pure SDS and DTAB micelles in brine determined by small-angle neutron scattering (SANS)
,”
Phys. Chem. Chem. Phys.
1
,
4437
4446
(
1999
).
94.
A.
Sanchez-Fernandez
,
K. J.
Edler
,
T.
Arnold
,
R. K.
Heenan
,
L.
Porcar
,
N. J.
Terrill
,
A. E.
Terry
, and
A. J.
Jackson
, “
Micelle structure in a deep eutectic solvent: A small-angle scattering study
,”
Phys. Chem. Chem. Phys.
18
,
14063
14073
(
2016
).
95.
L.
Matthews
,
S.
Ruscigno
,
S. E.
Rogers
,
P.
Bartlett
,
A. J.
Johnson
,
R.
Sochon
, and
W. H.
Briscoe
, “
Fracto-eutectogels: SDS fractal dendrites via counterion condensation in a deep eutectic solvent
,”
Phys. Chem. Chem. Phys.
23
,
11672
11683
(
2021
).
96.
R.
Stefanovic
,
G. B.
Webber
, and
A. J.
Page
, “
Polymer solvation in choline chloride deep eutectic solvents modulated by the hydrogen bond donor
,”
J. Mol. Liq.
279
,
584
593
(
2019
).
97.
C. V.
Teixeira
,
H.
Amenitsch
,
P.
Linton
,
M.
Lindén
, and
V.
Alfredsson
, “
The role played by salts in the formation of SBA-15, an in situ small-angle x-ray scattering/diffraction study
,”
Langmuir
27
,
7121
7131
(
2011
).
98.
M.
Zulauf
,
K.
Weckstrom
,
J. B.
Hayter
,
V.
Degiorgio
, and
M.
Corti
, “
Neutron scattering study of micelle structure in isotropic aqueous solutions of poly(oxyethylene) amphiphiles
,”
J. Phys. Chem.
89
,
3411
3417
(
1985
).
99.
V.
Peyre
,
S.
Bouguerra
, and
F.
Testard
, “
Micellization of dodecyltrimethylammonium bromide in water–dimethylsulfoxide mixtures: A multi-length scale approach in a model system
,”
J. Colloid Interface Sci.
389
,
164
174
(
2013
).
100.
E.
Fuguet
,
C.
Ràfols
, and
M.
Rosés
, “
Characterization of the solvation properties of surfactants by solvatochromic indicators
,”
Langmuir
19
,
6685
6692
(
2003
).
101.
S.
Berr
,
R. R. M.
Jones
, and
J. S.
Johnson
, “
Effect of counterion on the size and charge of alkyltrimethylammonium halide micelles as a function of chain length and concentration as determined by small-angle neutron scattering
,”
J. Phys. Chem.
96
,
5611
5614
(
1992
).
102.
W.
Kunz
, “
Specific ion effects in colloidal and biological systems
,”
Curr. Opin. Colloid Interface Sci.
15
,
34
39
(
2010
).
All article content, except where otherwise noted, is licensed under a Creative Commons Attribution (CC BY) license (http://creativecommons.org/licenses/by/4.0/).

Supplementary Material