Amphiphilic Self-Assembly of Alkanols in Protic Ionic LiquidsClick to copy article linkArticle link copied!
Abstract
Strong cohesive forces in protic ionic liquids (PILs) can induce a liquid nanostructure consisting of segregated polar and apolar domains. Small-angle X-ray scattering has shown that these forces can also induce medium chain length n-alkanols to self-assemble into micelle- and microemulsion-like structures in ethylammonium (EA+) and propylammonium (PA+) PILs, in contrast to their immiscibility with both water and ethanolammonium (EtA+) PILs. These binary mixtures are structured on two distinct length scales: one associated with the self-assembled n-alkanol aggregates and the other with the underlying liquid nanostructure. This suggests that EA+ and PA+ enable n-alkanol aggregation by acting as cosurfactants, which EtA+ cannot do because its terminating hydroxyl renders the cation nonamphiphilic. The primary determining factor for miscibility and self-assembly is the ratio of alkyl chain lengths of the alkanol and PIL cation, modulated by the anion type. These results show how ILs can support the self-assembly of nontraditional amphiphiles and enable the creation of new forms of soft matter.
Introduction
Experimental Section
Materials and Methods
Small- and Wide-Angle X-ray Scattering
Miscibility Measurements
Results and Discussion
EAN | PAN | EtAN | EAF | PAF | EtAF | |
---|---|---|---|---|---|---|
mp = | 13 °C (22) | 6 °C (20) | 51 °C (20) | –15 °C (20) | 48 °C (46) | –82 °C (47) |
ethanol | miscible | miscible | miscible | miscible | – | miscible |
butanol | miscible | miscible | immiscible | miscible | – | immiscible |
hexanol | miscible | miscible | immiscible | miscible | <10 wt %c | immiscible |
octanol | partialb | miscible | immiscible | miscible | <6.4 wt %c | immiscible |
decanol | immiscible | miscible | – | miscible | – | – |
dodecanol | immiscible | miscible | – | partial | – | – |
Data are for liquid mixtures at 25 °C, except for PAF which melts at 48 °C. Note that EtAN is a metastable liquid at this temperature. “–” denotes mixtures not examined in this study.
Partly miscible at room temperature. Octanol becomes fully miscible with EAN above its UCST. (41)
Solubility of solid PAF in alcohol at 25 °C. PAF is fully miscible with both alcohols above 48 °C.
Conclusions
Acknowledgment
This work was supported by the Australian Research Council and the University of Sydney.
References
This article references 76 other publications.
- 1Rogers, R. D.; Seddon, K. R. Ionic Liquids - Solvents of the Future Science 2003, 302, 792– 793Google ScholarThere is no corresponding record for this reference.
- 2MacFarlane, D. R.; Seddon, K. R. Ionic Liquids—Progress on the Fundamental Issues Aust. J. Chem. 2007, 60, 3– 5Google ScholarThere is no corresponding record for this reference.
- 3Earle, M. J.; Seddon, K. R. Pure Appl. Chem. 2000, 72, 1391– 1398Google ScholarThere is no corresponding record for this reference.
- 4Welton, T. Room-Temperature Ionic Liquids. Solvents for Synthesis and Catalysis Chem. Rev. 1999, 99, 2071– 2083Google Scholar4https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK1MXkt1artrw%253D&md5=e17d4c2a7f45438755b34161e86f24e6Room-Temperature Ionic Liquids. Solvents for Synthesis and CatalysisWelton, ThomasChemical Reviews (Washington, D. C.) (1999), 99 (8), 2071-2083CODEN: CHREAY; ISSN:0009-2665. (American Chemical Society)A review with 124 refs. covering org. reactions in alkylhalo- and haloaluminate ionic liqs.
- 5Swatloski, R. P.; Visser, A. E.; Reichert, W. M.; Broker, G. A.; Farina, L. M.; Holbrey, J. D.; Rogers, R. D. Solvation of 1-Butyl-3-methylimidazolium Hexafluorophosphate in Aqueous Ethanol - a Green Solution for Dissolving ’Hydrophobic’ Ionic Liquids Chem. Commun. (Cambridge, U. K.) 2001, 2070– 2071Google Scholar5https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3MXnsFWqu70%253D&md5=b1b8650a731fb7c3f970caf1839e6960Solvation of 1-butyl-3-methylimidazolium hexafluorophosphate in aqueous ethanol-a green solution for dissolving 'hydrophobic' ionic liquidsSwatloski, Richard P.; Visser, Ann E.; Reichert, W. Matthew; Broker, Grant A.; Farina, Lindsy M.; Holbrey, John D.; Rogers, Robin D.Chemical Communications (Cambridge, United Kingdom) (2001), (20), 2070-2071CODEN: CHCOFS; ISSN:1359-7345. (Royal Society of Chemistry)The relatively hydrophobic ionic liq. 1-butyl-3-methylimidazolium hexafluorophosphate has been found to be totally miscible with aq. ethanol between 0.5 and 0.9 mol fraction ethanol, whereas the ionic liq. is only partially miscible with either pure water or abs. ethanol. The ability to dissolve 1-butyl-3-methylimidazolium hexafluorophosphate in a 'green' aq. solvent system has important implications for cleaning, purifn., and sepns. using ionic liqs.
- 6Huddleston, J. G.; Visser, A. E.; Reichert, W. M.; Willauer, H. D.; Broker, G. A.; Rogers, R. D. Characterization and Comparison of Hydrophilic and Hydrophobic Room Temperature Ionic Liquids Incorporating the Imidazolium Cation Green Chem. 2001, 3, 156– 164Google Scholar6https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3MXmtVWhsr8%253D&md5=db13c3edffe333bc489b6ceeccf3bbbcCharacterization and comparison of hydrophilic and hydrophobic room temperature ionic liquids incorporating the imidazolium cationHuddleston, Jonathan G.; Visser, Ann E.; Reichert, W. Matthew; Willauer, Heather D.; Broker, Grant A.; Rogers, Robin D.Green Chemistry (2001), 3 (4), 156-164CODEN: GRCHFJ; ISSN:1463-9262. (Royal Society of Chemistry)A series of hydrophilic and hydrophobic 1-alkyl-3-methylimidazolium room temp. ionic liqs. (RTILs) have been prepd. and characterized to det. how water content, d., viscosity, surface tension, m.p., and thermal stability are affected by changes in alkyl chain length and anion. In the series of RTILs studied here, the choice of anion dets. water miscibility and has the most dramatic effect on the properties. Hydrophilic anions (e.g., chloride and iodide) produce ionic liqs. that are miscible in any proportion with water but, upon the removal of some water from the soln., illustrate how sensitive the phys. properties are to a change in water content. In comparison, for ionic liqs. contg. more hydrophobic anions (e.g., PF6- and N(SO2CF3)2-), the removal of water has a smaller affect on the resulting properties. For a series of 1-alkyl-3-methylimidazolium cations, increasing the alkyl chain length from Bu to hexyl to octyl increases the hydrophobicity and the viscosities of the ionic liqs. increase, whereas densities and surface tension values decrease. Thermal analyses indicate high temps. are attainable prior to decompn. and DSC studies reveal a glass transition for several samples. ILs incorporating PF6- have been used in liq./liq. partitioning of org. mols. from water and the results for two of these are also discussed here. On a cautionary note, the chem. of the individual cations and anions of the ILs should not be overlooked as, in the case of certain conditions for PF6- ILs, contact with an aq. phase may result in slow hydrolysis of the PF6- with the concomitant release of HF and other species.
- 7Anderson, J. L.; Pino, V.; Hagberg, E. C.; Sheares, V. V.; Armstrong, D. W. Surfactant Solvation Effects and Micelle Formation in Ionic Liquids Chem. Commun. (Cambridge, U. K.) 2003, 2444– 2445Google Scholar7https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3sXntlWhsLo%253D&md5=d268c6f6fde4796153de37a3472bafa4Surfactant solvation effects and micelle formation in ionic liquidsAnderson, Jared L.; Pino, Veronica; Hagberg, Erik C.; Sheares, Valerie V.; Armstrong, Daniel W.Chemical Communications (Cambridge, United Kingdom) (2003), (19), 2444-2445CODEN: CHCOFS; ISSN:1359-7345. (Royal Society of Chemistry)The formation of micelles in 1-butyl-3-Me imidazolium chloride (BMIM-Cl) and hexafluorophosphate (BMIM-PF6) were explored using different surfactants and the solvation behavior of the new micellar-ionic liq. solns. examd. using inverse gas chromatog.
- 8Li, N.; Zhang, S.; Ma, H.; Zheng, L. Role of Solubilized Water in Micelles Formed by Triton X-100 in 1-Butyl-3-methylimidazolium Ionic Liquids Langmuir 2010, 26, 9315– 9320Google ScholarThere is no corresponding record for this reference.
- 9He, Y.; Li, Z.; Simone, P.; Lodge, T. P. Self-Assembly of Block Copolymer Micelles in an Ionic Liquid J. Am. Chem. Soc. 2006, 128, 2745– 2750Google Scholar9https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD28XpsVKisA%253D%253D&md5=aa9a5a20e2e14c37ce5ddb8de6bfcfb0Self-Assembly of Block Copolymer Micelles in an Ionic LiquidHe, Yiyong; Li, Zhibo; Simone, Peter; Lodge, Timothy P.Journal of the American Chemical Society (2006), 128 (8), 2745-2750CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)Four amphiphilic poly((1,2-butadiene)-block-ethylene oxide) (PB-PEO) diblock copolymers were shown to aggregate strongly and form micelles in an ionic liq., 1-butyl-3-methylimidazolium hexafluorophosphate ([BMIM][PF6]). The universal micellar structures (spherical micelle, wormlike micelle, and bilayered vesicle) were all accessed by varying the length of the corona block while holding the core block const. The nanostructures of the PB-PEO micelles formed in an ionic liq. were directly visualized by cryogenic transmission electron microscopy (cryo-TEM). Detailed micelle structural information was extd. from both cryo-TEM and dynamic light scattering measurements, with excellent agreement between the two techniques. Compared to aq. solns. of the same copolymers, [BMIM][PF6] solns. exhibit some distinct features, such as temp.-independent micellar morphologies between 25° and 100°. As in aq. solns., significant nonergodicity effects were also obsd. This work demonstrates the flexibility of amphiphilic block copolymers for controlling nanostructure in an ionic liq., with potential applications in many arenas.
- 10Sharma, S. C.; Atkin, R.; Warr, G. G. The Effect of Ionic Liquid Hydrophobicity and Solvent Miscibility on Pluronic Amphiphile Self-Assembly J. Phys. Chem. B 2013, 117, 14568– 14575Google ScholarThere is no corresponding record for this reference.
- 11Yue, X.; Chen, X.; Wang, X.; Li, Z. Lyotropic Liquid Crystalline Phases Formed by Phyosterol Ethoxylates in Room-Temperature Ionic Liquids Colloids Surf., A 2011, 392, 225– 232Google Scholar11https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3MXhsFKrsrjO&md5=58b3dd8eedd05cbfcbba4069fe5184bdLyotropic liquid crystalline phases formed by phyosterol ethoxylates in room-temperature ionic liquidsYue, Xiu; Chen, Xiao; Wang, Xudong; Li, ZhihongColloids and Surfaces, A: Physicochemical and Engineering Aspects (2011), 392 (1), 225-232CODEN: CPEAEH; ISSN:0927-7757. (Elsevier B.V.)The phase behaviors of four phytosterol ethoxylates surfactants (BPS-n, n = 5, 10, 20, and 30) with different oxyethylene units in room temp. ionic liq., 1-butyl-3-methylimidazolium tetrafluoroborate ([Bmim]BF4), have been studied. The polarized optical microscopy and small-angle X-ray scattering techniques are used to characterize the phase structures of these binary systems at 25 °C. The structure and ordering of the liq. cryst. (LC) phases in such BPS-n/[Bmim]BF4 systems are found to be influenced by BPS-n concn. and the temp. Due to the bulky and rigid cholesterol group, the phytosterol ethoxylates surfactants exhibit different properties and interaction mechanism from the conventional CnEOm type nonionic surfactant systems. The rheol. measurements indicate a highly viscoelastic nature of these lyotropic LC phases and disclose a lamellar phase characteristic with a rather strong rigidity at high surfactant concns. The control expt. with Brij 97(polyoxyethylene (10) oleyl ether)/[Bmim]BF4 system and the FTIR measurements help to recognize that the solvophobic interaction combining with the hydrogen bonding are the main driving forces for the LC phases formation.
- 12Sakai, H.; Saitoh, T.; Misono, T.; Tsuchiya, K.; Sakai, K.; Abe, M. Phase Behavior of Phytosterol Ethoxylates in an Imidazolium-Type Room-Temperature Ionic Liquid J. Oleo Sci. 2012, 61, 135– 141Google Scholar12https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38XjsV2ltrk%253D&md5=e7cfefb1579bbb42f28d8cc597e8cabePhase behavior of phytosterol ethoxylates in an imidazolium-type room-temperature ionic liquidSakai, Hideki; Saitoh, Takanori; Misono, Takeshi; Tsuchiya, Koji; Sakai, Kenichi; Abe, MasahikoJournal of Oleo Science (2012), 61 (3), 135-141CODEN: JOSOAP; ISSN:1345-8957. (Japan Oil ChemistsÏ Society)The temp.-concn. phase behavior of nonionic surfactants in an aprotic imidazolium-type room-temp. ionic liq. (RT-IL) was evaluated on the basis of a combination of visual appearance, polarized optical microscopy, and small angle X-ray scattering data. Phytosterol ethoxylates (BPS-n, where n denotes oxyethylene chain lengths of 5, 10, 20, and 30) were used as surfactants in the RT-IL, 1-butyl-3-methylimidazolium hexafluorophosphate (BmimPF6). The two component mixts. yielded various phases such as discontinuous cubic, hexagonal, and lamellar phases. An increased tendency toward formation of lesser-curved mol. assemblies was obsd. at higher BPS-n concns., at lower temps., and for shorter oxyethylene chain surfactants. These trends are similar to those obsd. in aq. BPS-n systems; however, notable differences in the phase states of the aq. system vs. the BmimPF6 system were evident. Comparison with the water system showed that the BmimPF6 system yielded fewer phases and generally required higher BPS-n concns. to induce phase transitions. Evaluation of the effects of addn. of a third component (e.g., 1-dodecanol and dodecane) to the binary system on the phase behavior showed that at a given compn. ratio of BPS-20 to BmimPF6, the addn. of 1-dodecanol generally results in the phase transition to lesser-curved assemblies whereas dodecane generated no significant effects. The obsd. phase change is satisfactorily rationalized by localized solubilization of the third component into the binary surfactant assemblies.
- 13Wang, L. Y.; Chen, X.; Chai, Y. C.; Hao, J. C.; Sui, Z. M.; Zhuang, W. C.; Sun, Z. W. Lyotropic Liquid Crystalline Phases Formed in an Ionic Liquid Chem. Commun. (Cambridge, U. K.) 2004, 24, 2840– 2841Google ScholarThere is no corresponding record for this reference.
- 14Anjum, N.; Guedeau-Boudeville, M.-A.; Stubenrauch, C.; Mourchid, A. Phase Behavior and Microstructure of Microemulsions Containing the Hydrophobic Ionic Liquid 1-Butyl-3-methylimidazolium Hexafluorophosphate J. Phys. Chem. B 2009, 113, 239– 244Google ScholarThere is no corresponding record for this reference.
- 15Gao, H. X.; Li, J. C.; Han, B. X.; Chen, W. N.; Zhang, J. L.; Zhang, R.; Yan, D. D. Microemulsions with Ionic Liquid Polar Domains Phys. Chem. Chem. Phys. 2004, 6, 2914– 2916Google Scholar15https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2cXkt1arsbs%253D&md5=6c6c5027c64ba1dc07fc5977b3a74a43Microemulsions with ionic liquid polar domainsGao, Haixiang; Li, Junchun; Han, Buxing; Chen, Wenna; Zhang, Jianling; Zhang, Rui; Yan, DadongPhysical Chemistry Chemical Physics (2004), 6 (11), 2914-2916CODEN: PPCPFQ; ISSN:1463-9076. (Royal Society of Chemistry)In this work 1-butyl-3-methylimidazolium tetrafluoroborate ([bmim][BF4], an ionic liq.)/Triton X-100 (TX-100)/cyclohexane microemulsions have been prepd. and characterized by phase behavior, cond. measurement, dynamic light scattering, freeze-fracturing electron microscopy, and UV-vis techniques, and our attention is concd. on the microemulsions with the ionic liq. as the nano-sized polar domains.
- 16Gao, Y. N.; Han, S. B.; Han, B. X.; Li, G. Z.; Shen, D.; Li, Z. H.; Du, J. M.; Hou, W. G.; Zhang, G. Y. TX-100/water/1-butyl-3-methylimidazolium Hexafluorophosphate Microemulsions Langmuir 2005, 21, 5681– 5684Google ScholarThere is no corresponding record for this reference.
- 17Walden, P. Über die Molekulargröβe und elektrische Leitfähigkeit Einiger Geschmolzener Salze Bull. Acad. Imp. Sci. St.-Petersbourg 1914, 8, 405– 422Google ScholarThere is no corresponding record for this reference.
- 18Belieres, J. P.; Angell, C. A. Protic Ionic Liquids: Preparation, Characterization, and Proton Free Energy Level Representation J. Phys. Chem. B 2007, 111, 4926– 4937Google Scholar18https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2sXjvFCitb8%253D&md5=465b63065f29b7b0fcd88d76a1c83dd6Protic ionic liquids: preparation, characterization, and proton free energy level representationBelieres, Jean-Philippe; Angell, C. AustenJournal of Physical Chemistry B (2007), 111 (18), 4926-4937CODEN: JPCBFK; ISSN:1520-6106. (American Chemical Society)The authors give a perspective on the relations between inorg. and org. cation ionic liqs. (ILs), including members with m.ps. that overlap around the borderline 100 °C. The paper presents the synthesis and properties (melting, boiling, glass temps., etc.) of a large no. of an intermediate group of liqs. that cover the ground between equimolar mol. mixts. and ILs, depending on the energetics of transfer of a proton from one member of the pair to the other. These proton-transfer ILs have interesting properties, including the ability to serve as electrolytes in solvent-free fuel cell systems. This work provides a basis for assessing their relation to aprotic ILs by means of a Gurney-type proton-transfer free energy level diagram, with approx. values of the energy levels based on free energy of formation and pKa data. The energy level scheme allows verifying the relation between solvent-free acidic and basic electrolytes, and the familiar aq. variety, and to identify neutral protic electrolytes that are unavailable in the case of aq. systems.
- 19Yoshizawa, M.; Xu, W.; Angell, C. A. Ionic Liquids by Proton Transfer: Vapor Pressure, Conductivity, and the Relevance of Delta pKa from Aqueous Solutions J. Am. Chem. Soc. 2003, 125, 15411– 15419Google ScholarThere is no corresponding record for this reference.
- 20Greaves, T. L.; Drummond, C. J. Protic Ionic Liquids: Properties and Applications Chem. Rev. 2008, 108, 206– 237Google Scholar20https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2sXhsVCrsb%252FL&md5=16a3f06eef103a54dc1e38f9e469506bProtic Ionic Liquids: Properties and ApplicationsGreaves, Tamar L.; Drummond, Calum J.Chemical Reviews (Washington, DC, United States) (2008), 108 (1), 206-237CODEN: CHREAY; ISSN:0009-2665. (American Chemical Society)This review covers the ionicity, physicochem. and thermal properties of protic ionic liqs., along with a description of their applications, and where they have been used, including org. synthesis, chromatog., biol. applications, fuel cells, explosives, and, very recently, industrial lubricants.
- 21Angell, C. A.; Byrne, N.; Belieres, J.-P. Parallel Developments in Aprotic and Protic Ionic Liquids: Physical Chemistry and Applications Acc. Chem. Res. 2007, 40, 1228– 1236Google Scholar21https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2sXht1GmtrnJ&md5=8e0c1cb0803e1730718477540b13174aParallel Developments in Aprotic and Protic Ionic Liquids: Physical Chemistry and ApplicationsAngell, C. Austen; Byrne, Nolene; Belieres, Jean-PhilippeAccounts of Chemical Research (2007), 40 (11), 1228-1236CODEN: ACHRE4; ISSN:0001-4842. (American Chemical Society)A review. This account covers research dating from the early 1960s in the field of low-melting molten salts and hydrates, which has recently become popular under the rubric of "ionic liqs.". It covers understanding gained in the principal author's labs. (initially in Australia, but mostly in the U.S.A.) from spectroscopic, dynamic, and thermodn. studies and includes recent applications of this understanding in the fields of energy conversion and biopreservation. Both protic and aprotic varieties of ionic liqs. are included, but recent studies have focused on the protic class because of the special applications made possible by the highly variable proton activities available in these liqs.
- 22Evans, D. F.; Yamauchi, A.; Roman, R.; Casassa, E. Z. Micelle Formation in Ethylammonium Nitrate, a Low-Melting Fused Salt J. Colloid Interface Sci. 1982, 88, 89– 96Google Scholar22https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaL38XktFahtbs%253D&md5=25a5726d4d7af7f22422bfcc4eb17442Micelle formation in ethylammonium nitrate, a low-melting fused saltEvans, D. Fennell; Yamauchi, Akira; Roman, Ronald; Casassa, Ethel Z.Journal of Colloid and Interface Science (1982), 88 (1), 89-96CODEN: JCISA5; ISSN:0021-9797.Crit. micelle concns. (CMC) were detd. from surface tension measurements for alkyltrimethylammonium bromides and alkylpyridinium bromides at 50° and for Triton X-100 at 20 and 50° in ethylammonium nitrate, a low-melting anhyd. fused salt. The CMC's are ∼5 to 10 times larger than those obsd. in water. From the change of CMC with surfactant chain length, the free energy of transfer of a methylene group from the fused salt to the micelle interior is calcd. to be -370 cal/mol compared to -680 cal/mol for a similar transfer from water to the micelle. With respect to solvophobic behavior, ethylammonium nitrate and water show a no. of similarities.
- 23Evans, D. F.; Yamauchi, A.; Wei, G. J.; Bloomfield, V. A. Micelle Size in Ethylammonium Nitrate as Determined by Classical and Quasi-Elastic Light Scattering J. Phys. Chem. 1983, 87, 3537– 3541Google Scholar23https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaL3sXkvFCit74%253D&md5=a3df15122ddfe5942e7c9fb033a43879Micelle size in ethylammonium nitrate as determined by classical and quasi-elastic light scatteringEvans, D. Fennell; Yamauchi, A.; Wei, G. Jason; Bloomfield, Victor A.Journal of Physical Chemistry (1983), 87 (18), 3537-41CODEN: JPCHAX; ISSN:0022-3654.The aggregation of surfactants to form micelles in ethylammonium nitrate, a low melting fused salt, was investigated by classical and quasi-elastic light scattering. For tetradecylpyridinium bromide and hexadecylpyridinium bromide, the following data are obtained: crit. micelle concns., 8.0 × 10-2 and 2.0 × 10-2 mol kg-1; micellar aggregation nos., 17 and 26; 2nd virial coeffs., 1.64 × 10-3 and 1.30 × 10-3 mol3 g-2; and hydrodynamic radii, 14 and 22 Å, resp. The results are consistent with either a small classical spherical micelle contg. only surfactant or a spherical mixed micelle contg. surfactant and ethylammonium ions as a cosurfactant. The 2nd virial coeffs. are almost equal to those calcd. for hard spheres and reflect highly screened electrostatic interactions in the totally ionized solvent.
- 24Araos, M. U.; Warr, G. G. Self-Assembly of Nonionic Surfactants into Lyotropic Liquid Crystals in Ethylammonium Nitrate, a Room-Temperature Ionic Liquid J. Phys. Chem. B 2005, 109, 14275– 14277Google Scholar24https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2MXmtVWisL4%253D&md5=9d6f0f5f1cfff23f7947cec06cd3d309Self-Assembly of Nonionic Surfactants into Lyotropic Liquid Crystals in Ethylammonium Nitrate, a Room-Temperature Ionic LiquidAraos, Miguel U.; Warr, Gregory G.Journal of Physical Chemistry B (2005), 109 (30), 14275-14277CODEN: JPCBFK; ISSN:1520-6106. (American Chemical Society)The stability of a variety of lyotropic liq. crystals formed by a no. of polyoxyethylene nonionic surfactants in the room-temp. ionic liq. ethylammonium nitrate (EAN) is surveyed and reported. The pattern of self-assembly behavior and mesophase formation is strikingly similar to that obsd. in water, even including the existence of a lower consolute boundary or cloud point. The only quant. difference from water is that longer alkyl chains are necessary to drive the formation of liq. cryst. mesophases in EAN, suggesting that a rich pattern of "solvophobic" self-assembly should exist in this solvent.
- 25Araos, M. U.; Warr, G. G. Structure of Nonionic Surfactant Micelles in the Ionic Liquid Ethylammonium Nitrate Langmuir 2008, 24, 9354– 9360Google Scholar25https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1cXpt1yntr4%253D&md5=98b23022b0fcfdd828838b4926412e1cStructure of Nonionic Surfactant Micelles in the Ionic Liquid Ethylammonium NitrateAraos, Miguel U.; Warr, Gregory G.Langmuir (2008), 24 (17), 9354-9360CODEN: LANGD5; ISSN:0743-7463. (American Chemical Society)The structure of micelles formed by nonionic polyoxyethylene alkyl ether nonionic surfactants, CnEm, in room-temp. ionic liq., ethylammonium nitrate (EAN), was detd. by small-angle neutron scattering (SANS) as a function of alkyl and ethoxy chain length, concn., and temp. Micelles form for all surfactant alkyl chain lengths from dodecyl through to octadecyl. The dodecyl-chain surfactants have high crit. micelle concn., around 1%, and form weakly structured micelles. Surfactants with longer alkyl chains readily form micelles in EAN. The obsd. micelle structure changes systematically with alkyl and ethoxy chain length, in parallel with observations in aq. solns. Decreasing ethoxy chain length at const. alkyl chain length leads to a sphere to rod transition. These micelles also grow into rods with increasing temp. as their cloud point is approached in EAN.
- 26Atkin, R.; Warr, G. G. Phase Behavior and Microstructure of Microemulsions with a Room-Temperature Ionic Liquid as the Polar Phase J. Phys. Chem. B 2007, 111, 9309– 9316Google Scholar26https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2sXns12ltb4%253D&md5=df6c503f69753f9b50aa817e4555795aPhase Behavior and Microstructure of Microemulsions with a Room-Temperature Ionic Liquid as the Polar PhaseAtkin, Rob; Warr, Gregory G.Journal of Physical Chemistry B (2007), 111 (31), 9309-9316CODEN: JPCBFK; ISSN:1520-6106. (American Chemical Society)Microemulsions of nonionic alkyl oligoethyleneoxide (CiEj) surfactants, alkanes, and ethylammonium nitrate (EAN), a room-temp. ionic liq., have been prepd. and characterized. Studies of phase behavior reveal that EAN microemulsions have many features in common with corresponding aq. systems, the primary difference being that higher surfactant concns. and longer surfactant tailgroups are required to offset the decreased solvophobicity the surfactant mols. in EAN compared with water. The response of the EAN microemulsions to variation in the length of the alkane, surfactant headgroup, and surfactant tailgroup has been found to parallel that obsd. in aq. systems in most instances. EAN microemulsions exhibit a single broad small-angle X-ray scattering peak, like aq. systems. These are well described by the Teubner-Strey model. A lamellar phase was also obsd. for surfactants with longer tails at lower temps. The scattering peaks of both microemulsion and lamellar phases move to lower wave vector on increasing temp. This is ascribed to a decrease in the interfacial area of the surfactant layer. Phase behavior, small-angle X-ray scattering, and cond. expts. have allowed the weakly to strongly structured transition to be identified for EAN systems.
- 27Atkin, R.; De Fina, L.-M.; Kiederling, U.; Warr, G. G. Structure and Self Assembly of Pluronic Amphiphiles in Ethylammonium Nitrate and at the Silica Surface J. Phys. Chem. B 2009, 113, 12201– 12213Google ScholarThere is no corresponding record for this reference.
- 28Velasco, S. B.; Turmine, M.; Di Caprio, D.; Letellier, P. Micelle Formation in Ethyl-ammonium Nitrate (an Ionic Liquid) Colloids Surf., A 2006, 275, 50– 54Google ScholarThere is no corresponding record for this reference.
- 29Evans, D. F.; Kaler, E. W.; Benton, W. J. Liquid Crystals in a Fused Salt: β,γ-Distearoylphosphatidylcholine in N-Ethylammonium Nitrate J. Phys. Chem. 1983, 87, 533– 535Google ScholarThere is no corresponding record for this reference.
- 30Greaves, T. L.; Weerawardena, A.; Fong, C.; Drummond, C. J. Many Protic Ionic Liquids Mediate Hydrocarbon-Solvent Interactions and Promote Amphiphile Self-Assembly Langmuir 2007, 23, 402– 404Google Scholar30https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD28Xht12qt77M&md5=b664062a092ce3c7dd5ce1c6f13b2ef7Many Protic Ionic Liquids Mediate Hydrocarbon-Solvent Interactions and Promote Amphiphile Self-AssemblyGreaves, Tamar L.; Weerawardena, Asoka; Fong, Celesta; Drummond, Calum J.Langmuir (2007), 23 (2), 402-404CODEN: LANGD5; ISSN:0743-7463. (American Chemical Society)A large no. of protic ionic liqs. (PILs) have been found to mediate solvent-hydrocarbon interactions and promote amphiphile self-assembly. Hexagonal, cubic, and lamellar lyotropic liq. cryst. phases were obsd. in PIL-hexadecyltrimethylammonium bromide systems. The driving force for the formation of the self-assembled aggregate structures has been attributed to an entropic contribution to the free energy of assocn., analogous to the hydrophobic effect in water. The specific aggregate structures formed depend upon the cationic and anionic components of the PIL and their interactions with the amphiphiles.
- 31Evans, D. F.; Chen, S.-H.; Schriver, G. W.; Arnett, E. M. Thermodynamics of Solution of Nonpolar Gases in a Fused Salt. Hydrophobic Bonding Behavior in a Nonaqueous System J. Am. Chem. Soc. 1981, 103, 481– 482Google ScholarThere is no corresponding record for this reference.
- 32Hayes, R.; Imberti, S.; Warr, G. G.; Atkin, R. The Nature of Hydrogen Bonding in Protic Ionic Liquids Angew. Chem., Int. Ed. 2013, 52, 4623– 4627Google Scholar32https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3sXjtlaksLs%253D&md5=726d2360de55ca62fee0d747e702834fThe Nature of Hydrogen Bonding in Protic Ionic LiquidsHayes, Robert; Imberti, Silvia; Warr, Gregory G.; Atkin, RobAngewandte Chemie, International Edition (2013), 52 (17), 4623-4627CODEN: ACIEF5; ISSN:1433-7851. (Wiley-VCH Verlag GmbH & Co. KGaA)The ion arrangements in protic ionic liqs. (PILs) are a consequence of the balance between intermol. forces and the phys. dimensions of the ions. While similar bicontinuous sponge-like nanostructures form in all systems, there is significant variation in the H-bonding. The PIL nanostructure controls the H-bond strength and structure by dictating the relative orientations of the cations and anions to each other. The distribution of H-bond geometries is related to the ability of each PIL to accommodate H-bonds in the bicontinuous arrangement rather than inducing a different bulk structure. When the nanostructures have a relatively high proportion of H-bonds are linear (RX:HN ≈ 1), attractions between ions increase and phys. properties become more solid-like. At higher RX:HN ratios, bi-or trifurcated H-bonding results, leading to weaker, bent- H-bonds, decreasing cation-anion attractions and leading to a more fluid-like material.
- 33Atkin, R.; Warr, G. G. The Smallest Amphiphiles: Nanostructure in Protic Room-Temperature Ionic Liquids with Short Alkyl Groups J. Phys. Chem. B 2008, 112, 4164– 4166Google Scholar33https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1cXjtlOjsb0%253D&md5=0f0a37486d4fb146d73c2e7435cc711dThe Smallest Amphiphiles: Nanostructure in Protic Room-Temperature Ionic Liquids with Short Alkyl GroupsAtkin, Rob; Warr, Gregory G.Journal of Physical Chemistry B (2008), 112 (14), 4164-4166CODEN: JPCBFK; ISSN:1520-6106. (American Chemical Society)Room-temp. ionic liqs. (ILs) are low-melting-point org. salts that, until recently, were thought to have homogeneous microstructure. In this work, we investigate nanoscale segregation of short (<C4) alkyl chain ILs using propylammonium nitrate (PAN) and ethylammonium nitrate (EAN). Structure peaks at q = 0.54 Å-1 for PAN and at q = 0.66 Å-1 for EAN, corresponding to Bragg spacings (D* = 2π/qmax) of 11.6 and 9.7 Å resp., provide the first exptl. evidence of nanoscale heterogeneity for ILs with alkyl chains less than C4. The observation that these ILs are not optically birefringent and the fits obtained suggest a disordered, locally smectic or sponge-like structure. Solvophobic interaction between alkyl groups is the most important factor for the prodn. of nanoscale heterogeneities, but electrostatic and hydrogen bonding attractions between the amine nitrogen and the nitrate anion will also play a role.
- 34Hayes, R.; Imberti, S.; Warr, G. G.; Atkin, R. Amphiphilicity Determines Nanostructure in Protic Ionic Liquids Phys. Chem. Chem. Phys. 2011, 13, 3237– 3247Google Scholar34https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3MXhsFymtL4%253D&md5=aa28e5cc0144742a23aee635ad0f3567Amphiphilicity determines nanostructure in protic ionic liquidsHayes, Robert; Imberti, Silvia; Warr, Gregory G.; Atkin, RobPhysical Chemistry Chemical Physics (2011), 13 (8), 3237-3247CODEN: PPCPFQ; ISSN:1463-9076. (Royal Society of Chemistry)The bulk structure of the two oldest ionic liqs. (ILs), ethylammonium nitrate (EAN) and ethanolammonium nitrate (EtAN), is elucidated using neutron diffraction. The spectra were modelled using empirical potential structure refinement (EPSR). EAN exhibits a long-range structure of solvophobic origin, similar to a bicontinuous microemulsion or disordered L3-sponge phase, but with a domain size of only 1 nm. The alc. (-OH) moiety in EtAN interferes with solvophobic assocn. between cation alkyl chains resulting in small clusters of ions, rather than an extended network.
- 35Hayes, R.; Imberti, S.; Warr, G. G.; Atkin, R. Pronounced Sponge-like Nanostructure in Propylammonium Nitrate Phys. Chem. Chem. Phys. 2011, 13, 13544– 13551Google Scholar35https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3MXptVyrs78%253D&md5=3589b22ad9cfb74c67a2c2123077f587Pronounced sponge-like nanostructure in propylammonium nitrateHayes, Robert; Imberti, Silvia; Warr, Gregory G.; Atkin, RobPhysical Chemistry Chemical Physics (2011), 13 (30), 13544-13551CODEN: PPCPFQ; ISSN:1463-9076. (Royal Society of Chemistry)The bulk structure of the ionic liq. propylammonium nitrate (PAN) was detd. using neutron diffraction. Empirical potential structure refinement (EPSR) fits to the data show that PAN self-assembles into a quasi-periodic bicontinuous nanostructure reminiscent of an amphiphile L3-sponge phase. Atomic detail on the ion arrangements around the propylammonium cation and nitrate anion yields evidence of hydrogen bonding between ammonium and nitrate groups and of strong alkyl chain aggregation and interdigitation. The resultant amphiphilic PAN nanostructure is more pronounced than that previously detd. for ethylammonium nitrate (EAN) or ethanolammonium nitrate (EtAN).
- 36Greaves, T. L.; Kennedy, D. F.; Mudie, S. T.; Drummond, C. J. Diversity Observed in the Nanostructure of Protic Ionic Liquids J. Phys. Chem. B 2010, 114, 10022– 10031Google Scholar36https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3cXptVSmuro%253D&md5=a217468aa9c8707d0dcc9eaa5262b09bDiversity Observed in the Nanostructure of Protic Ionic LiquidsGreaves, Tamar L.; Kennedy, Danielle F.; Mudie, Stephen T.; Drummond, Calum J.Journal of Physical Chemistry B (2010), 114 (31), 10022-10031CODEN: JPCBFK; ISSN:1520-6106. (American Chemical Society)The nanostructure of a series of 20 protic ionic liqs. (PILs) has been investigated using small- and wide-angle X-ray scattering (SAXS and WAXS). The PILs contained alkylammonium, dialkylammonium, trialkylammonium, and cyclic ammonium cations combined with org. or inorg. anions. The presence of hydroxyl and methoxy substituents on the alkyl chains of the cations was also explored. Many of the PILs showed a nanostructure resulting from segregation of the polar and nonpolar components of the ionic liq. It was found that this segregation was enhanced for longer alkyl chains, with a corresponding increase in the length scale, whereas the presence of hydroxyl groups on the alkyl chains led to much less ordered liqs. The broad range of protic ionic liqs. studied allowed several structure-property relationships to be established. The solvophobic effect was shown to be dependent on the nanostructure of the PILs. These PILs support amphiphile self-assembly, and it was shown that the less structured PILs had more "water-like" behavior in the diversity of lyotropic liq.-crystal phases supported, and the thermal stability ranges for these phases.
- 37Hayes, R.; Imberti, S.; Warr, G. G.; Atkin, R. How Water Dissolves in Protic Ionic Liquids Angew. Chem., Int. Ed. 2012, 51, 7468– 7471Google Scholar37https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38Xos1altbo%253D&md5=af19ae3e1fef9d34bd2f38b6e06d9d97How Water Dissolves in Protic Ionic LiquidsHayes, Robert; Imberti, Silvia; Warr, Gregory G.; Atkin, RobAngewandte Chemie, International Edition (2012), 51 (30), 7468-7471, S7468/1-S7468/13CODEN: ACIEF5; ISSN:1433-7851. (Wiley-VCH Verlag GmbH & Co. KGaA)When equal masses of water and the protic ionic liq. ethylammonium nitrate are mixed a bicontinuous nanostructure results. This nanostructure resembles aq. surfactant mesophases but has length scales at least an order of magnitude smaller. The local structure of both the water and the ionic liq. are strikingly similar to that found in the pure liqs.
- 38Greaves, T. L.; Kennedy, D. F.; Weerawardena, A.; Tse, N. M. K.; Kirby, N.; Drummond, C. J. Nanostructured Protic Ionic Liquids Retain Nanoscale Features in Aqueous Solution While Precursor Bronsted Acids and Bases Exhibit Different Behavior J. Phys. Chem. B 2011, 115, 2055– 2066Google ScholarThere is no corresponding record for this reference.
- 39Hadded, M.; Mayaffre, A.; Letellier, P. Surface-Tension Of Ideal Solutions - Application To Binary-Mixtures Of Methanol And Molten Ethylammonium Nitrate At 298-K J. Chim. Phys. Phys.-Chim. Biol. 1989, 86, 525– 537Google Scholar39https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaL1MXlsVWit78%253D&md5=7aba3f1284aef3e535b6c8fd8a77bbcbSurface tension of ideal solutions: application to binary mixtures of methanol and molten ethylammonium nitrate at 298 KHadded, M.; Mayaffre, A.; Letellier, P.Journal de Chimie Physique et de Physico-Chimie Biologique (1989), 86 (3), 525-37CODEN: JCPBAN; ISSN:0021-7689.Surface tension and excess vols. measurements of MeOH-molten EtNH2.HNO3 mixts. were detd. at 298 K. In terms of excess vols., the behavior of these solns. can be described approx. by the laws of ideal solns. For the surface tensions, an unusual formalism of the Gibbs equation, which involves concns. of the soln. components and a distance parameter, Δh, is proposed. The parameter Δh does not depend on the choice of surface location. This formalism is able to define the behavior of the surface for an ideal mixt. Surface tensions of MeOH-EtNH2.HNO3 mixts. are described accurately by this equation over the entire concn. range.
- 40Russina, O.; Sferrazza, A.; Caminiti, R.; Triolo, A. Amphiphile Meets Amphiphile: Beyond the Polar–Apolar Dualism in Ionic Liquid/Alcohol Mixtures J. Phys. Chem. Lett. 2014, 5 (10) 1738– 1742Google Scholar40https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2cXntlKnu74%253D&md5=babb2d72aaacfa008b1c8ef42b2892a1Amphiphile Meets Amphiphile: Beyond the Polar-Apolar Dualism in Ionic Liquid/Alcohol MixturesRussina, Olga; Sferrazza, Alessio; Caminiti, Ruggero; Triolo, AlessandroJournal of Physical Chemistry Letters (2014), 5 (10), 1738-1742CODEN: JPCLCD; ISSN:1948-7185. (American Chemical Society)The mesoscopic morphol. of binary mixts. of ethylammonium nitrate (EAN), the protic ionic liq. par excellence, and methanol is explored using neutron/X-ray diffraction and computational techniques. Both compds. are amphiphilic and characterized by an extended hydrogen bonding network: surprisingly, though macroscopically homogeneous, these mixts. turn out to be mesoscopically highly heterogeneous. Even in methanol-rich mixts., a wide distribution of clusters exists where EAN preserves its bulk, sponge-like morphol. Accordingly, methanol does not succeed in fully dissocg. the ionic liq. that keeps on organizing in a bulk-like fashion. This behavior represents the premise to the more dramatic phenomenol. obsd. with longer alcs. that eventually phase sep. from EAN. These results challenge the commonly accepted polar and apolar moieties segregation in ionic liqs./mol. liqs. mixts. and the current understanding of technol. relevant solvation processes.
- 41Weingärtner, H. S.; Merkel, T.; Kashammer, S.; Schröer, W.; Wiegand, S. The Effect of Short-Range Hydrogen-Bonded Interactions on the Nature of the Critical-Point of Ionic Fluids. 1. General-Properties of the New System Ethylammonium Nitrate + n-Octanol with an Upper Consolute Point Near Room-Temperature Ber. Bunsen-Ges. Phys. Chem. 1993, 97, 970– 975Google ScholarThere is no corresponding record for this reference.
- 42Weingärtner, H. S.; Schröer, W. Liquid-Liquid Phase Separations and Critical-Behavior of Electrolyte-Solutions Driven by Long-Range and Short-Range Interactions J. Mol. Liq. 1995, 65–66, 107– 114Google ScholarThere is no corresponding record for this reference.
- 43Sahandzhieva, K.; Tuma, D.; Breyer, S.; Kamps, A. P. S.; Maurer, G. Liquid-Liquid Equilibrium in Mixtures of the Ionic Liquid 1-n-Butyl-3-methylimidazolium Hexafluorophosphate and an Alkanol J. Chem. Eng. Data 2006, 51, 1516– 1525Google ScholarThere is no corresponding record for this reference.
- 44Vale, V. R.; Will, S.; Schroer, W.; Rathke, B. The General Phase Behavior of Mixtures of 1-Alkyl-3-Methylimidazolium Bis[(trifluoromethyl)sulfonyl]amide Ionic Liquids with n-Alkyl Alcohols ChemPhysChem 2012, 13, 1860– 1867Google ScholarThere is no corresponding record for this reference.
- 45Greaves, T. L.; Kennedy, D. F.; Kirby, N.; Drummond, C. J. Nanostructure Changes in Protic Ionic Liquids (PILs) Through Adding Solutes and Mixing PILs Phys. Chem. Chem. Phys. 2011, 13, 13501– 13509Google ScholarThere is no corresponding record for this reference.
- 46Greaves, T. L.; Weerawardena, A.; Fong, C.; Krodkiewska, I.; Drummond, C. J. Protic Ionic Liquids: Solvents with Tunable Phase Behavior and Physicochemical Properties J. Phys. Chem. B 2006, 110, 22479– 22487Google Scholar46https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD28XhtVyhsbfI&md5=dff830a123cdc27a178c01d5b6159ca0Protic Ionic Liquids: Solvents with Tunable Phase Behavior and Physicochemical PropertiesGreaves, Tamar L.; Weerawardena, Asoka; Fong, Celesta; Krodkiewska, Irena; Drummond, Calum J.Journal of Physical Chemistry B (2006), 110 (45), 22479-22487CODEN: JPCBFK; ISSN:1520-6106. (American Chemical Society)The phase behavior, including glass, devitrification, solid crystal melting, and liq. boiling transitions, and physicochem. properties, including d., refractive index, viscosity, cond., and air-liq. surface tension, of a series of 25 protic ionic liqs. and protic fused salts are presented along with structure-property comparisons. The protic fused salts were mostly liq. at room temp., and many exhibited a glass transition occurring at low temps. between -114 and -44 °C, and high fragility, with many having low viscosities, down to as low as 17 mPa·s at 25 °C, and ionic conductivities up to 43.8 S/cm at 25 °C. These protic solvents are easily prepd. through the stoichiometric combination of a primary amine and Bronsted acid. They have poor ionic behavior when compared to the far more studied aprotic ionic liqs. However, some of the other physicochem. properties possessed by these solvents are highly promising and it is anticipated that these, or analogous protic solvents, will find applications beyond those already identified for aprotic ionic liqs. This series of protic fused salts was employed to det. the effect of structural changes on the physicochem. properties, including the effect of hydroxyl groups, increasing alkyl chain lengths, branching, and the differences between inorg. and org. anions. It was found that simple structural modifications provide a mechanism to manipulate, over a wide range, the temp. at which phase transitions occur and to specifically tailor physicochem. properties for potential end-use applications.
- 47Bicak, N. A New Ionic Liquid: 2-Hydroxy Ethylammonium Formate J. Mol. Liq. 2005, 116, 15– 18Google Scholar47https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2cXpsVSiu7c%253D&md5=44e2bb3f4b913ed1773a438984405598A new ionic liquid: 2-hydroxyethylammonium formateBicak, NiyaziJournal of Molecular Liquids (2004), 116 (1), 15-18CODEN: JMLIDT; ISSN:0167-7322. (Elsevier B.V)A new ionic liq. (2-hydroxyethylammonium formate), with extremely low melting temp. (-82°), is presented. This ionic liq. of equimolar mixt. of formic acid and 2-hydroxyethylamine shows reasonably high at ionic cond. (3.3 mS cm-1) room temp. and heat stability up to 150°. Both 1H-NMR and FT-IR spectra establish its simple salt structure. Due to its high polarity, the ionic liq. is able to dissolve many inorg. salts. Also some insol. polymers such as polyaniline and polypyrrole are highly sol. in the ionic liq. In the study, the phys. characteristics of the ionic liq., such as cond., viscosity and solvation abilities were investigated.
- 48Wakeham, D.; Niga, P.; Ridings, C.; Andersson, G.; Nelson, A.; Warr, G. G.; Baldelli, S.; Rutland, M. W.; Atkin, R. Surface Structure of a “Non-Amphiphilic” Protic Ionic Liquid Phys. Chem. Chem. Phys. 2012, 14, 5106– 5114Google ScholarThere is no corresponding record for this reference.
- 49Wakeham, D.; Warr, G. G.; Atkin, R. Surfactant Adsorption at the Surface of Mixed Ionic Liquids and Ionic Liquid Water Mixtures Langmuir 2012, 28, 13224– 13231Google Scholar49https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38Xht1egtrvI&md5=1cea2405847cd62de4930f51fabc6389Surfactant Adsorption at the Surface of Mixed Ionic Liquids and Ionic Liquid Water MixturesWakeham, Deborah; Warr, Gregory G.; Atkin, RobLangmuir (2012), 28 (37), 13224-13231CODEN: LANGD5; ISSN:0743-7463. (American Chemical Society)Surface tensiometry and neutron reflectivity were used to elucidate the structure of the adsorbed layer of nonionic surfactant tetraethylene glycol tetradecyl ether (C14E4) at the free surface of the ionic liqs. ethylammonium nitrate (EAN) and ethanolammonium nitrate (EtAN) and their binary mixts. with each other and with H2O. Surface tensions reveal that the crit. micelle concn. (cmc) depends strongly on solvent compn. The adsorbed surfactant structure elucidated by neutron reflectivity shows that the level of solvation of the ethylene oxide groups varies for both the pure and mixed solvents. This is attributed to solvent-solvent interactions dominating solvent-surfactant interactions.
- 50Greaves, T. L.; Mudie, S. T.; Drummond, C. J. 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. 2011, 13, 20441– 20452Google ScholarThere is no corresponding record for this reference.
- 51Wakeham, D.; Eschebach, D.; Webber, G. B.; Atkin, R.; Warr, G. G. Surface Composition of Mixtures of Ethylammonium Nitrate, Ethanolammonium Nitrate and Water Aust. J. Chem. 2012, 65, 1554– 1556Google ScholarThere is no corresponding record for this reference.
- 52Kashyap, H. K.; Hettige, J. J.; Annapureddy, H. V. R.; Margulis, C. J. SAXS Anti-peaks Reveal the Length-Scales of Dual Positive–Negative and Polar–Apolar Ordering in Room-Temperature Ionic Liquids Chem. Commun. 2012, 48, 5103– 5105Google ScholarThere is no corresponding record for this reference.
- 53Greaves, T. L.; Drummond, C. J. Solvent Nanostructure, the Solvophobic Effect and Amphiphile Self-Assembly in Ionic Liquids Chem. Soc. Rev. 2013, 42, 1096– 1120Google Scholar53https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3sXosV2iuw%253D%253D&md5=4426adee83b9bc3915570527a39e12abSolvent nanostructure, the solvophobic effect and amphiphile self-assembly in ionic liquidsGreaves, Tamar L.; Drummond, Calum J.Chemical Society Reviews (2013), 42 (3), 1096-1120CODEN: CSRVBR; ISSN:0306-0012. (Royal Society of Chemistry)A review. The ability of ionic liqs. (ILs) to support amphiphile self-assembly into a range of mesophase structures was established as a widespread phenomenon. From the ILs evaluated as self-assembly media, the vast majority have supported some lyotropic liq. crystal phase formation. Many neat ionic liqs. segregate into polar and nonpolar domains to form a nanostructured liq. A very strong correlation between the nanostructure of the ionic liq. and its characteristics as an amphiphile self-assembly solvent was found. In this review ionic liqs. as amphiphile self-assembly media, and identify trends that can be used to distinguish which ionic liqs. probably have good promotion properties as self-assembly media are discussed. In particular these trends focus on the nanostructure of neat ionic liqs., their solvent cohesive energy d., and the related solvophobic effect. The authors forecast that many more ILs will be identified as amphiphile self-assembly solvents in the future.
- 54Stewart, G. W.; Morrow, R. M. X-ray Diffraction in Liquids Primary Normal Alcohols Phys. Rev. 1927, 30, 232– 244Google ScholarThere is no corresponding record for this reference.
- 55Vahvaselka, K. S.; Serimaa, R.; Torkkeli, M. Determination of Liquid Structures of the Primary Alcohols Methanol, Ethanol, 1-Propanol, 1-Butanol and 1-Octanol By X-Ray-Scattering J. Appl. Crystallogr. 1995, 28, 189– 195Google Scholar55https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK2MXltVektb8%253D&md5=8a9cf507c6d6b681ae0e566a2dec1505Determination of liquid structures of the primary alcohols methanol, ethanol, 1-propanol, 1-butanol and 1-octanol by x-ray scatteringVahvaselka, Kaarlo Sakari; Serimaa, Ritva; Torkkeli, MikaJournal of Applied Crystallography (1995), 28 (2), 189-95CODEN: JACGAR; ISSN:0021-8898. (Munksgaard)The four first primary alcs. and 1-octanol were studied with the wide-angle X-ray scattering method using both transmission and reflection techniques at room temp. (293 K). The subsequent radial distribution anal. gives C-O and C-C distances that are in good agreement with the known bond lengths of single alc. mols. The shortest averaged C-H and O-H distances are much smaller than those from the gas-phase data. The shortened distance is caused by the deformed electron distribution between H and C atoms. The intermol. hydrogen bonding between hydroxyl groups occurs at a distance of 2.8 Å, and is a common feature for all the samples. The corresponding coordination no. of 1.7 indicates linear chains of about ten mols. The absence of sharp features in the intensity curves suggests that the arrangement of chains is irregular and that the length of the chains varies.
- 56Tomsic, M.; Jamnik, A.; Fritz-Popovski, G.; Glatter, O.; Vlcek, L. Structural Properties of Pure Simple Alcohols from Ethanol, Propanol, Butanol, Pentanol, to Hexanol: Comparing Monte Carlo Simulations with Experimental SAXS Data J. Phys. Chem. B 2007, 111, 1738– 1751Google ScholarThere is no corresponding record for this reference.
- 57Vrhovsek, A.; Gereben, O.; Jamnik, A.; Pusztai, L. Hydrogen Bonding and Molecular Aggregates in Liquid Methanol, Ethanol, and 1-Propanol J. Phys. Chem. B 2011, 115, 13473– 13488Google Scholar57https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3MXhtlKksLjI&md5=2d980442f347352f9f32e4ef925c8bd9Hydrogen Bonding and Molecular Aggregates in Liquid Methanol, Ethanol, and 1-PropanolVrhovsek, Aleksander; Gereben, Orsolya; Jamnik, Andrej; Pusztai, LaszloJournal of Physical Chemistry B (2011), 115 (46), 13473-13488CODEN: JPCBFK; ISSN:1520-5207. (American Chemical Society)The authors present a detailed and comprehensive structural study of mol. models of liq. methanol, ethanol, and 1-propanol that originate from reverse Monte Carlo (RMC), mol. dynamics (MD), and united-atom Monte Carlo (UA:MC) simulations. The authors compare several modeling approaches: RMC simulations that employ exptl. neutron and x-ray diffraction data as sole constraints, RMC with diffraction data complemented with partial radial distribution functions (PRDFs) from MD or UA:MC, and conventional MD and UA:MC simulations. The assessment is done in view of the structural parameters of the hydrogen bond and resulting morphol. characteristics of mol. aggregates. To achieve these tasks, a computer program for structural anal. of mol. configurations together with the appropriate aggregate classification scheme was developed. The authors analyzed the morphol. of clusters, their probability, and size distributions. Any cyclic structures that appeared in the configurations were extd. and characterized in the same manner. MD and UA:MC simulations resulted in configurations with bulkier, more threadlike aggregates that were not entirely consistent with the exptl. evidence from diffraction expts. A combination of neutron and x-ray diffraction data with PRDFs from MD simulations, simultaneously applied as constraints in the RMC procedure, proved to be a modeling approach with the most conclusive results.
- 58Dixit, S.; Crain, J.; Poon, W. C. K.; Finney, J. L.; Soper, A. K. Molecular Segregation Observed in a Concentrated Alcohol–Water Solution Nature 2002, 416, 829– 832Google ScholarThere is no corresponding record for this reference.
- 59Schröer, W.; Wiegand, S.; Weingärtner, H. The Effect of Short-Range Hydrogen-Bonded Interactions on the Nature of the Critical-Point of Ionic Fluids. 2. Static and Dynamic Light-Scattering on Solutions of Ethylammonium Nitrate in N-Octanol Ber. Bunsen-Ges. Phys. Chem. Chem. Phys. 1993, 97, 975– 982Google Scholar59https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK3sXmsFOgur8%253D&md5=dddd81df1ed9d2a243c6cdc882569ebfThe effect of short-range hydrogen-bonded interactions on the nature of the critical point of ionic fluids. Part II. Static and dynamic light scattering on solutions of ethylammonium nitrate in n-octanolSchroer, W.; Wiegand, S.; Weingaertner, H.Berichte der Bunsen-Gesellschaft (1993), 97 (8), 975-82CODEN: BBPCAX; ISSN:0005-9021.The authors report on static and dynamic light scattering measurements of near-crit. fluctuations of an ionic liq. The system is ethylammonium nitrate (EAN) dissolved in n-octanol which exhibits a liq.-liq. phase transition in the salt-rich regime with an upper crit. point at ∼315 K at a mole fraction of the salt x1 = 0.773. The correlation lengths and scattering intensities scale with temp. corresponding to an Ising-like crit. point. This contrasts with previous observations of mean-field behavior in electrolyte systems with liq.-liq. phase transitions driven by long-range Coulombic interactions. Rationalizations of this discrepancy are offered taking into account a) the competition of intermol. interactions with long-range and short-range character and b) the possibility of crossover from mean-field to Ising behavior as predicted by the Landau-Ginzburg theory.
- 60Atkin, R.; Bobillier, S. M. C.; Warr, G. G. Propylammonium Nitrate as a Solvent for Amphiphile Self-Assembly into Micelles, Lyotropic Liquid Crystals, and Microemulsions J. Phys. Chem. B 2010, 114, 1350– 1360Google Scholar60https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1MXhs1alurzL&md5=31e9e6905206e5eb3092ed1b5f5b2391Propylammonium Nitrate as a Solvent for Amphiphile Self-Assembly into Micelles, Lyotropic Liquid Crystals, and MicroemulsionsAtkin, Rob; Bobillier, Sophie M. C.; Warr, Gregory G.Journal of Physical Chemistry B (2010), 114 (3), 1350-1360CODEN: JPCBFK; ISSN:1520-6106. (American Chemical Society)The phase behavior and self-assembled microstructures of a range of oligo(oxyethylene)-n-alkyl ether (CiEj) surfactants was studied in propylammonium nitrate (PAN), a room temp. ionic liq. Micelles and single-phase microemulsions were all found to form at alkyl chain lengths from dodecyl to octadecyl, and lyotropic liq. crystals formed with hexadecyl chains or longer. Small-angle neutron scattering (SANS) shows that self-assembly occurs by solvophobic interactions driving the aggregation of the alkyl chains, but several results indicate that these are weaker in PAN than in water or ethylammonium nitrate, due chiefly to the hydrophobicity of PAN. Longer alkyl chains are needed for lyotropic liq. crystals to form, and higher surfactant concns. are needed to form a single phase microemulsion. Cond. shows these microemulsions to be weakly structured, and relatively insensitive to oil or surfactant mol. structure, unlike water-based systems. However, SANS contrast variation reveals a nanosegregation of oil from the alkyl tails of surfactants within the microemulsion, and may suggest a cosurfactant-like role for the propylammonium cation. Mol. areas within microemulsions and lamellar phases are larger than corresponding water- or ethylammonium nitrate-based systems due to the large mol. vol. of the solvating PANs.
- 61Topolnicki, I. L.; Atkin, R.; FitzGerald, P. A.; Warr, G. G. The Effect of Protic Ionic Liquid and Surfactant Structure on Partitioning of Polyoxyethylene Nonionic Surfactants. ChemPhysChem 2014, 15.Google ScholarThere is no corresponding record for this reference.
- 62Kline, S. R. Reduction and Analysis of SANS and USANS Data using Igor Pro J. Appl. Crystallogr. 2006, 39 (6) 895– 900Google Scholar62https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD28Xht1KrurjP&md5=51449389d831a41c38dd4145c342df4dReduction and analysis of SANS and USANS data using IGOR ProKline, Steven R.Journal of Applied Crystallography (2006), 39 (6), 895-900CODEN: JACGAR; ISSN:0021-8898. (International Union of Crystallography)A software package is presented for performing redn. and anal. of small-angle neutron scattering (SANS) and ultra-small-angle neutron scattering (USANS) data. A graphical interface was developed to visualize and quickly reduce raw SANS and USANS data into 1D or 2D formats for interpretation. The resulting reduced data can then be analyzed using model-independent methods or nonlinear fitting to one of a large and growing catalog of included structural models. The different instrumental smearing effects for slit-smeared USANS and pinhole-smeared SANS data are handled automatically during anal. In addn., any no. of SANS and USANS data sets can be analyzed simultaneously. The redn. operations and anal. models are written in a modular format for extensibility, allowing users to contribute code and models for distribution to all users. The software package is based on Igor Pro, providing freely distributable and modifiable code that runs on Macintosh and Windows operating systems.
- 63Teubner, M.; Strey, R. Origin of the Scattering Peak in Microemulsions J. Chem. Phys. 1987, 87, 3195– 3200Google Scholar63https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaL2sXlvVChs7w%253D&md5=2e671ec8ee923fd23f5f0275ff6836d3Origin of the scattering peak in microemulsionsTeubner, M.; Strey, R.Journal of Chemical Physics (1987), 87 (5), 3195-200CODEN: JCPSA6; ISSN:0021-9606.From a Landau theory the static scattering intensity distribution I(q) of microemulsions is obtained. It is shown that the scattering intensity relation describes exptl. literature data remarkably well, using only three fit parameters.
- 64Schubert, K. V.; Strey, R. Small-Angle Neutron Scattering from Microemulsions Near the Disorder Line in Water/Formamide--Octane-CiEj Systems J. Chem. Phys. 1991, 95, 8532– 8545Google ScholarThere is no corresponding record for this reference.
- 65Vonk, C. G.; Billman, J. F.; Kaler, E. W. Small Angle Scattering of Bicontinuous Structures in Microemulsions J. Chem. Phys. 1988, 88, 3970– 3975Google ScholarThere is no corresponding record for this reference.
- 66Schubert, K.-V.; Strey, R.; Kline, S. R.; Kaler, E. W. Small Angle Neutron Scattering Near Lifshitz Lines: Transition from Weakly Structured Mixtures to Microemulsions J. Chem. Phys. 1994, 101, 5343– 5355Google Scholar66https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK2cXmvVOjsLo%253D&md5=1041f5facf121cec63c8d98947b2191cSmall angle neutron scattering near Lifshitz lines: transition from weakly structured mixtures to microemulsionsSchubert, K.-V.; Strey, R.; Kline, S. R.; Kaler, E. W.Journal of Chemical Physics (1994), 101 (6), 5343-55CODEN: JCPSA6; ISSN:0021-9606.We have studied the phase behavior, wetting transitions, and small angle neutron scattering (SANS) of water, n-alkane, and n-alkyl polyglycol ether (CiEj) systems in order to locate the transition between weakly structured mixts. and microemulsions, and to provide a measure for the transition. We first detd. the wetting transition by macroscopic measurements and then measured the location of the Lifshitz lines by SANS. Starting with well-structured mixts. (exhibiting nonwetting middle phases and well-expressed scattering peaks, features that qualify them as microemulsions) the wetting transition was induced by increasing the chain length of the alkane or by changing the oil/water vol. ratio, and then the Lifshitz line was crossed. Further, starting with systems past the disorder line (weakly structured mixts. that display wetting middle phases and no scattering peaks), local structure was induced by either increasing the surfactant concn. or decreasing the oil/water vol. ratio or the temp. In each case a Lifshitz line was crossed. Analyzing the scattering expts. quant., allows detn. of the amphiphilicity factor, which is a measure of the strength of the surfactant. The results suggest there is a sequence of roughly parallel surfaces within the three-dimensional compn.-temp. space. As the amphiphilicity factor increases, first a disorder surface is encountered, then a Lifshitz surface, and finally a wetting transition surface. How and to what extent these surfaces move in the one-phase region toward smaller surfactant concns., and intersect there with the body of heterogeneous phases, depends on a no. of factors that are discussed in some detail.
- 67Barnes, I. S.; Corti, M.; Degiorgio, V.; Zemb, T. X-ray and Neutron-Scattering Measurements on Concentrated Non-Ionic Amphiphile Solutions Prog. Colloid Polym. Sci. 1993, 93, 205– 206Google Scholar67https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK2cXmvVOjs7g%253D&md5=c7da655991ee31f843df0648224adefcX-ray and neutron-scattering measurements on concentrated non-ionic amphiphile solutionsBarnes, I. S.; Corti, M.; Degiorgio, V.; Zemb, T.Progress in Colloid & Polymer Science (1993), 93 (TRENDS IN COLLOID AN), 205-6CODEN: PCPSD7; ISSN:0340-255X.The structure of aq. solns. of the nonionic amphiphile C12E8 is studied by small-angle x-ray and neutron scattering along the isothermal path across the single-phase region from 0% to 100% amphiphile vol. fraction. Soln. structuring is present even in the pure amphiphile, as shown by x-ray data.
- 68Laughlin, R. The Aqueous Phase Behavior of Surfactants; Academic Press: San Diego, 1994.Google ScholarThere is no corresponding record for this reference.
- 69Diat, O.; K, M. L.; Touraud, D.; Deme, B.; Grillo, I.; Kunz, W.; Zemb, T. Octanol-Rich and Water-Rich Domains in Dynamic Equilibrium in the Pre-Ouzo Region of Ternary Systems Containing a Hydrotrope J. Appl. Crystallogr. 2013, 46, 1665– 1669Google Scholar69https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3sXhsl2ntL3J&md5=7e7cb9f121dada6fd0bdec4b4747e090Octanol-rich and water-rich domains in dynamic equilibrium in the pre-ouzo region of ternary systems containing a hydrotropeDiat, Olivier; Klossek, Michael L.; Touraud, Didier; Deme, Bruno; Grillo, Isabelle; Kunz, Werner; Zemb, ThomasJournal of Applied Crystallography (2013), 46 (6), 1665-1669CODEN: JACGAR; ISSN:0021-8898. (International Union of Crystallography)Ternary mixts. of medium-chain fatty alcs., water and a hydrotrope (such as ethanol), near the immiscibility gap, make stable single phases at const. temp. Interestingly, in this 'pre-ouzo region' these single phases consist of two distinct nanoscopic pseudo-phases, one octanol-rich and one water-rich. This domain of compn., which is known to produce strong light scattering and to sep. under ultracentrifugation into two phases, has been studied using contrast variation in small-angle neutron scattering (SANS) combined with small- and wide-angle X-ray scattering (SWAXS). The existence of fatty alc.-rich domains of well defined size of the order of 2 nm radius is proven. The scattering can be approximated by an Ornstein-Zernike function, which is close to the general expression of Choi, Chen, Sottmann & Strey [Physica B, (1998), 241-243, 976-978] with vanishing quadratic Porod term. Exploitation of the relative intensities at the vanishing scattering angle in SANS demonstrates that the distribution coeff. of ethanol between the octanol-rich and the water-rich domains is close to one. WAXS of the two coexisting pseudo-phases is compared with the corresponding binary water-ethanol and octanol-ethanol samples.
- 70Tchakalova, V.; Zemb, T.; Benczédi, D. Evaporation Triggered Self-Assembly in Aqueous Fragrance–Ethanol Mixtures and Its Impact on Fragrance Performance. Colloids Surf., A 2014, in pressGoogle ScholarThere is no corresponding record for this reference.
- 71Clint, J. H. Micellization of Mixed Nonionic Surface Active Agents J. Chem. Soc., Faraday Trans. 1 1975, 71 (0) 1327– 1334Google ScholarThere is no corresponding record for this reference.
- 72Kahlweit, M.; Strey, R.; Busse, G. Weakly to Strongly Structured Mixtures Phys. Rev. E 1993, 47, 4197– 4209Google ScholarThere is no corresponding record for this reference.
- 73Greaves, T. L.; Drummond, C. J. Ionic Liquids as Amphiphile Self-Assembly Media Chem. Soc. Rev. 2008, 37, 1709– 1726Google Scholar73https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1cXoslOgtbk%253D&md5=512c82d5493b639e25979d578faed5c9Ionic liquids as amphiphile self-assembly mediaGreaves, Tamar L.; Drummond, Calum J.Chemical Society Reviews (2008), 37 (8), 1709-1726CODEN: CSRVBR; ISSN:0306-0012. (Royal Society of Chemistry)A review. In recent years, the no. of nonaq. solvents which mediate hydrocarbon-solvent interactions and promote the self-assembly of amphiphiles was markedly increased by the reporting of over 30 ionic liqs. which possess this previously unusual solvent characteristic. This new situation allows a different exploration of the mol. solvophobic effect and tests the current understanding of amphiphile self-assembly. Both protic and aprotic ionic liqs. support amphiphile self-assembly, indicating that it is not required for the solvents to be able to form a hydrogen bonded network. Here, the use of ionic liqs. as amphiphile self-assembly media is reviewed, including micelle and liq. cryst. mesophase formation, their use as a solvent phase in microemulsions and emulsions, and the emerging field of nanostructured inorg. materials synthesis. Surfactants, lipids and block co-polymers are the focus amphiphile classes in this crit. review (174 refs.).
- 74Beesley, A. H.; Evans, D. F.; Laughlin, R. G. Evidence For The Essential Role Of Hydrogen-Bonding In Promoting Amphiphilic Self-Assembly - Measurements in 3-Methylsydnone J. Phys. Chem. 1988, 92, 791– 793Google ScholarThere is no corresponding record for this reference.
- 75Ray, A. Micelle Formation in Pure Ethylene Glycol J. Am. Chem. Soc. 1969, 91 (23) 6511– 6512Google ScholarThere is no corresponding record for this reference.
- 76Ray, A. Solvophobic Interactions and Micelle Formation in Structure Forming Nonaqueous Solvents Nature 1971, 231, 313– 315Google Scholar76https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaE3MXksFSksrk%253D&md5=5e9f7c861f0b8352050969c6f20e9356Solvophobic interactions and micelle formation in structure forming nonaqueous solventsRay, AshokaNature (London, United Kingdom) (1971), 231 (5301), 313-15CODEN: NATUAS; ISSN:0028-0836.Solvophobic interactions occur in glycerol, HCONH2, H2N(CH2)2OH, HO(CH2)3OH, HO(CH2)4OH, HCO2H, H2N(CH2)2NH2, HS(CH2)2OH, HOCH2CH2CH(OH)Me, HOCH2CH(OH)Me, and H2NCH2CH(OH)Me (class I). Such interactions do not exist in HCONHMe, HCONMe2, and CH2I2 (class II), or in MeOH, EtOH, and PhMe (class III). A nonionic detergent [e.g., (p-tert-nonylphenoxy)polyethylene glycol] would aggregate into micelles only in class I solvents. The crit. micelle concns. of the detergent, and the free energies of micelle formation were calcd.
Cited By
This article is cited by 72 publications.
- Shurui Miao, Amaar Sardharwalla, Susan Perkin. Ion Diffusion Reveals Heterogeneous Viscosity in Nanostructured Ionic Liquids. The Journal of Physical Chemistry Letters 2024, 15
(47)
, 11855-11861. https://doi.org/10.1021/acs.jpclett.4c02996
- Xu Liu, Alessandro Mariani, Thomas Diemant, Maria Enrica Di Pietro, Xu Dong, Po-Hua Su, Andrea Mele, Stefano Passerini. PFAS-Free Locally Concentrated Ionic Liquid Electrolytes for Lithium Metal Batteries. ACS Energy Letters 2024, 9
(6)
, 3049-3057. https://doi.org/10.1021/acsenergylett.4c00814
- Navjot K. Kahlon, Emma L. Matthewman, Mohamad El Mohamad, Tamar L. Greaves, Cameron C. Weber. Small-Angle X-ray Scattering Study of the Amphiphilic Bulk Nanostructure of Tetraalkylammonium Deep Eutectic Solvents. The Journal of Physical Chemistry B 2024, 128
(18)
, 4566-4575. https://doi.org/10.1021/acs.jpcb.4c00943
- Stefan Paporakis, Kenny T.-C. Liu, Stuart J. Brown, Jason B. Harper, Andrew V. Martin, Tamar L. Greaves. Thermal Stability of Protic Ionic Liquids. The Journal of Physical Chemistry B 2024, 128
(17)
, 4208-4219. https://doi.org/10.1021/acs.jpcb.3c08011
- Jianan Wang, Joshua J. Buzolic, Jesse W. Mullen, Paul A. Fitzgerald, Zachary M. Aman, Maria Forsyth, Hua Li, Debbie S. Silvester, Gregory G. Warr, Rob Atkin. Nanostructure of Locally Concentrated Ionic Liquids in the Bulk and at Graphite and Gold Electrodes. ACS Nano 2023, 17
(21)
, 21567-21584. https://doi.org/10.1021/acsnano.3c06609
- Joshua B. Marlow, Rob Atkin, Gregory G. Warr. How Does Nanostructure in Ionic Liquids and Hybrid Solvents Affect Surfactant Self-Assembly?. The Journal of Physical Chemistry B 2023, 127
(7)
, 1490-1498. https://doi.org/10.1021/acs.jpcb.2c07458
- Katherine S. Lefroy, Brent S. Murray, Michael E. Ries. Effect of Oil on Cellulose Dissolution in the Ionic Liquid 1-Butyl-3-methyl Imidazolium Acetate. ACS Omega 2022, 7
(42)
, 37532-37545. https://doi.org/10.1021/acsomega.2c04311
- Shurui Miao, Ingo Hoffmann, Michael Gradzielski, Gregory G. Warr. Lipid Membrane Flexibility in Protic Ionic Liquids. The Journal of Physical Chemistry Letters 2022, 13
(23)
, 5240-5245. https://doi.org/10.1021/acs.jpclett.2c00980
- Oscar Nordness, Joan F. Brennecke. Ion Dissociation in Ionic Liquids and Ionic Liquid Solutions. Chemical Reviews 2020, 120
(23)
, 12873-12902. https://doi.org/10.1021/acs.chemrev.0c00373
- Saffron J. Bryant, Rob Atkin, Michael Gradzielski, Gregory G. Warr. Catanionic Surfactant Self-Assembly in Protic Ionic Liquids. The Journal of Physical Chemistry Letters 2020, 11
(15)
, 5926-5931. https://doi.org/10.1021/acs.jpclett.0c01608
- Yong-Lei Wang, Bin Li, Sten Sarman, Francesca Mocci, Zhong-Yuan Lu, Jiayin Yuan, Aatto Laaksonen, Michael D. Fayer. Microstructural and Dynamical Heterogeneities in Ionic Liquids. Chemical Reviews 2020, 120
(13)
, 5798-5877. https://doi.org/10.1021/acs.chemrev.9b00693
- Christiaan Ridings, Gregory G. Warr, and Gunther G. Andersson . Surface Ordering in Binary Mixtures of Protic Ionic Liquids. The Journal of Physical Chemistry Letters 2017, 8
(17)
, 4264-4267. https://doi.org/10.1021/acs.jpclett.7b01654
- Alessandro Mariani, Ruggero Caminiti, Fabio Ramondo, Giovanna Salvitti, Francesca Mocci, and Lorenzo Gontrani . Inhomogeneity in Ethylammonium Nitrate–Acetonitrile Binary Mixtures: The Highest “Low q Excess” Reported to Date. The Journal of Physical Chemistry Letters 2017, 8
(15)
, 3512-3522. https://doi.org/10.1021/acs.jpclett.7b01244
- Haihui Joy Jiang, Silvia Imberti, Rob Atkin, and Gregory G. Warr . Dichotomous Well-defined Nanostructure with Weakly Arranged Ion Packing Explains the Solvency of Pyrrolidinium Acetate. The Journal of Physical Chemistry B 2017, 121
(27)
, 6610-6617. https://doi.org/10.1021/acs.jpcb.7b03045
- Saffron J. Bryant, Rob Atkin, and Gregory G. Warr . Effect of Deep Eutectic Solvent Nanostructure on Phospholipid Bilayer Phases. Langmuir 2017, 33
(27)
, 6878-6884. https://doi.org/10.1021/acs.langmuir.7b01561
- Olga Russina, Fabrizio Lo Celso, Natalia V. Plechkova, and Alessandro Triolo . Emerging Evidences of Mesoscopic-Scale Complexity in Neat Ionic Liquids and Their Mixtures. The Journal of Physical Chemistry Letters 2017, 8
(6)
, 1197-1204. https://doi.org/10.1021/acs.jpclett.6b02811
- Alessandro Mariani, Rajeev Dattani, Ruggero Caminiti, and Lorenzo Gontrani . Nanoscale Density Fluctuations in Ionic Liquid Binary Mixtures with Nonamphiphilic Compounds: First Experimental Evidence. The Journal of Physical Chemistry B 2016, 120
(40)
, 10540-10546. https://doi.org/10.1021/acs.jpcb.6b07295
- Wolffram Schroer, Alessandro Triolo, and Olga Russina . Nature of Mesoscopic Organization in Protic Ionic Liquid–Alcohol Mixtures. The Journal of Physical Chemistry B 2016, 120
(9)
, 2638-2643. https://doi.org/10.1021/acs.jpcb.6b01422
- Sugosh R. Prabhu and G. B. Dutt . Rotational Diffusion of Charged and Nondipolar Solutes in Ionic Liquid–Organic Solvent Mixtures: Evidence for Stronger Specific Solute–Solvent Interactions in Presence of Organic Solvent. The Journal of Physical Chemistry B 2015, 119
(33)
, 10720-10726. https://doi.org/10.1021/acs.jpcb.5b06297
- Robert Hayes, Gregory G. Warr, and Rob Atkin . Structure and Nanostructure in Ionic Liquids. Chemical Reviews 2015, 115
(13)
, 6357-6426. https://doi.org/10.1021/cr500411q
- Sugosh R. Prabhu and G. B. Dutt . Effect of Low Viscous Nondipolar Solvent on the Rotational Diffusion of Structurally Similar Nondipolar Solutes in an Ionic Liquid. The Journal of Physical Chemistry B 2015, 119
(5)
, 2019-2025. https://doi.org/10.1021/jp512456c
- Thomas Murphy, Luis M. Varela, Grant B. Webber, Gregory G. Warr, and Rob Atkin . Nanostructure–Thermal Conductivity Relationships in Protic Ionic Liquids. The Journal of Physical Chemistry B 2014, 118
(41)
, 12017-12024. https://doi.org/10.1021/jp507408r
- Aurélien Perera. Site-site interaction model for alcohol models in two-dimension. Journal of Molecular Liquids 2025, 422 , 126944. https://doi.org/10.1016/j.molliq.2025.126944
- Yulia A. Kondratenko. From alkanolamines to protic alkanolammonium ionic liquids. Journal of Molecular Liquids 2024, 409 , 125460. https://doi.org/10.1016/j.molliq.2024.125460
- Stefan Paporakis, Stuart J. Brown, Connie Darmanin, Susanne Seibt, Patrick Adams, Michael Hassett, Andrew V. Martin, Tamar L. Greaves. Lyotropic liquid crystal phases of monoolein in protic ionic liquids. The Journal of Chemical Physics 2024, 160
(2)
https://doi.org/10.1063/5.0180420
- Shurui Miao, Michael Gradzielski, Gregory Warr. Micelle structure of nonionic surfactants containing carbon dioxide moieties in protic ionic liquids. Colloid and Polymer Science 2023, 301
(7)
, 813-820. https://doi.org/10.1007/s00396-023-05139-5
- Shurui Miao, Rob Atkin, Gregory Warr. Design and applications of biocompatible choline amino acid ionic liquids. Green Chemistry 2022, 24
(19)
, 7281-7304. https://doi.org/10.1039/D2GC02282F
- Igor A. Sedov, Timur I. Magsumov. Solvation properties of protic ionic liquids 2-methoxyethylammonium nitrate, propylammonium hydrogen sulfate, and butylammonium hydrogen sulfate. The Journal of Chemical Thermodynamics 2022, 170 , 106779. https://doi.org/10.1016/j.jct.2022.106779
- Shurui Miao, Silvia Imberti, Rob Atkin, Gregory Warr. Nanostructure in amino acid ionic molecular hybrid solvents. Journal of Molecular Liquids 2022, 351 , 118599. https://doi.org/10.1016/j.molliq.2022.118599
- Leon de Villiers Engelbrecht, Francesca Mocci, Yonglei Wang, Sergiy Perepelytsya, Tudor Vasiliu, Aatto Laaksonen. Molecular Perspective on Solutions and Liquid Mixtures from Modelling and Experiment. 2022, 53-84. https://doi.org/10.1007/978-3-030-80924-9_3
- Saffron J. Bryant, Alvaro Garcia, Ronald J. Clarke, Gregory G. Warr. Selective ion transport across a lipid bilayer in a protic ionic liquid. Soft Matter 2021, 17
(10)
, 2688-2694. https://doi.org/10.1039/D0SM02225J
- Alessandro Mariani, Leon Engelbrecht, Andrea Le Donne, Francesca Mocci, Enrico Bodo, Stefano Passerini. Disclosing the hierarchical structure of ionic liquid mixtures by multiscale computational methods. 2021, 1-67. https://doi.org/10.1016/B978-0-12-820280-7.00014-0
- Shurui Miao, Jared Wood, Haihui Joy Jiang, Silvia Imberti, Rob Atkin, Gregory Warr. Unusual origin of choline phenylalaninate ionic liquid nanostructure. Journal of Molecular Liquids 2020, 319 , 114327. https://doi.org/10.1016/j.molliq.2020.114327
- Dilek Yalcin, Ivan D. Welsh, Emma L. Matthewman, Seongmin Paul Jun, Mikkaila Mckeever-Willis, Iana Gritcan, Tamar L. Greaves, Cameron C. Weber. Structural investigations of molecular solutes within nanostructured ionic liquids. Physical Chemistry Chemical Physics 2020, 22
(20)
, 11593-11608. https://doi.org/10.1039/D0CP00783H
- S. Karlapudi, Z. Liu, Q. Bi, V. Govinda, I. Bahadur, O. Baswanth, Ch. Prasad. Molecular interactions in liquid mixtures containing o-cresol and 1-alkanols: Thermodynamics, FT-IR and computational studies. Journal of Molecular Liquids 2020, 305 , 112798. https://doi.org/10.1016/j.molliq.2020.112798
- Shurui Miao, Rob Atkin, Gregory G. Warr. Amphiphilic nanostructure in choline carboxylate and amino acid ionic liquids and solutions. Physical Chemistry Chemical Physics 2020, 22
(6)
, 3490-3498. https://doi.org/10.1039/C9CP06752C
- Gregory G. Warr, Rob Atkin. Solvophobicity and amphiphilic self-assembly in neoteric and nanostructured solvents. Current Opinion in Colloid & Interface Science 2020, 45 , 83-96. https://doi.org/10.1016/j.cocis.2019.12.009
- Alessandro Mariani, Matteo Bonomo, Stefano Passerini. Statistic-Driven Proton Transfer Affecting Nanoscopic Organization in an Ethylammonium Nitrate Ionic Liquid and 1,4-Diaminobutane Binary Mixture: A Steamy Pizza Model. Symmetry 2019, 11
(11)
, 1425. https://doi.org/10.3390/sym11111425
- Amzad Khan, Rashi Gusain, Manisha Sahai, Om P. Khatri. Fatty acids-derived protic ionic liquids as lubricant additive to synthetic lube base oil for enhancement of tribological properties. Journal of Molecular Liquids 2019, 293 , 111444. https://doi.org/10.1016/j.molliq.2019.111444
- Minh T. Lam, William D. Adamson, Shurui Miao, Rob Atkin, Gregory G. Warr. DTAB micelle formation in ionic liquid/water mixtures is determined by ionic liquid cation structure. Journal of Colloid and Interface Science 2019, 552 , 597-603. https://doi.org/10.1016/j.jcis.2019.05.082
- Igor A. Sedov, Timur M. Salikov, Boris N. Solomonov. Contrasting the solvation properties of protic ionic liquids with different nanoscale structure. Journal of Molecular Liquids 2019, 290 , 111361. https://doi.org/10.1016/j.molliq.2019.111361
- Andrei Filippov, Oleg I. Gnezdilov, Alexander G. Luchkin, Oleg N. Antzutkin. Self-diffusion of ethylammonium nitrate ionic liquid confined between modified polar glasses. Journal of Molecular Liquids 2019, 284 , 366-371. https://doi.org/10.1016/j.molliq.2019.04.021
- Saffron J. Bryant, Charl J. Jafta, Rob Atkin, Michael Gradzielski, Gregory G. Warr. Catanionic and chain-packing effects on surfactant self-assembly in the ionic liquid ethylammonium nitrate. Journal of Colloid and Interface Science 2019, 540 , 515-523. https://doi.org/10.1016/j.jcis.2019.01.048
- Haihui Joy Jiang, Silvia Imberti, Blake A. Simmons, Rob Atkin, Gregory G. Warr. Structural Design of Ionic Liquids for Optimizing Aromatic Dissolution. ChemSusChem 2019, 12
(1)
, 270-274. https://doi.org/10.1002/cssc.201802016
- Manuel Brunner, Hua Li, Zhezi Zhang, Dongke Zhang, Rob Atkin. Pinewood pyrolysis occurs at lower temperatures following treatment with choline-amino acid ionic liquids. Fuel 2019, 236 , 306-312. https://doi.org/10.1016/j.fuel.2018.09.004
- Andrei Filippov, Oleg I. Gnezdilov, Oleg N. Antzutkin. Static magnetic field alters properties of confined alkylammonium nitrate ionic liquids. Journal of Molecular Liquids 2018, 268 , 49-54. https://doi.org/10.1016/j.molliq.2018.07.027
- Haihui Joy Jiang, Rob Atkin, Gregory G. Warr. Nanostructured ionic liquids and their solutions: Recent advances and emerging challenges. Current Opinion in Green and Sustainable Chemistry 2018, 12 , 27-32. https://doi.org/10.1016/j.cogsc.2018.05.003
- Sraddha Agrawal, Hemant K. Kashyap. Structures of binary mixtures of ionic liquid 1-butyl-3-methylimidazolium bis(trifluoromethylsulfonyl)imide with primary alcohols: The role of hydrogen-bonding. Journal of Molecular Liquids 2018, 261 , 337-349. https://doi.org/10.1016/j.molliq.2018.03.124
- Fabrizio Lo Celso, Alessandro Triolo, Lorenzo Gontrani, Olga Russina. Communication: Anion-specific response of mesoscopic organization in ionic liquids upon pressurization. The Journal of Chemical Physics 2018, 148
(21)
https://doi.org/10.1063/1.5036588
- Marco Campetella, Alessandro Mariani, Claudia Sadun, Boning Wu, Edward W. Castner, Lorenzo Gontrani. Structure and dynamics of propylammonium nitrate-acetonitrile mixtures: An intricate multi-scale system probed with experimental and theoretical techniques. The Journal of Chemical Physics 2018, 148
(13)
https://doi.org/10.1063/1.5021868
- Andrei Filippov, Oleg N. Antzutkin. Magnetic field effects dynamics of ethylammonium nitrate ionic liquid confined between glass plates. Physical Chemistry Chemical Physics 2018, 20
(9)
, 6316-6320. https://doi.org/10.1039/C7CP06554J
- Vladimir A. Azov, Ksenia S. Egorova, Marina M. Seitkalieva, Alexey S. Kashin, Valentine P. Ananikov. “Solvent-in-salt” systems for design of new materials in chemistry, biology and energy research. Chemical Society Reviews 2018, 47
(4)
, 1250-1284. https://doi.org/10.1039/C7CS00547D
- S.M. Hosseini, M.M. Alavianmehr, A. Gutiérrez, R. Khalifeh, J. Moghadasi, S. Aparicio. On the properties and structure of 2-hydroxyethylammonium formate ionic liquid. Journal of Molecular Liquids 2018, 249 , 233-244. https://doi.org/10.1016/j.molliq.2017.10.122
- Olga Russina, Fabrizio Lo Celso, Natalia Plechkova, Charl J. Jafta, Giovanni Battista Appetecchi, Alessandro Triolo. Mesoscopic organization in ionic liquids. Topics in Current Chemistry 2017, 375
(3)
https://doi.org/10.1007/s41061-017-0147-2
- Hadrián Montes-Campos, José M. Otero-Mato, Trinidad Méndez-Morales, Elena López-Lago, Olga Russina, Oscar Cabeza, Luis J. Gallego, Luis M. Varela. Nanostructured solvation in mixtures of protic ionic liquids and long-chained alcohols. The Journal of Chemical Physics 2017, 146
(12)
https://doi.org/10.1063/1.4978943
- Gourav Shrivastav, Aditya Gupta, Aman Rastogi, Debdas Dhabal, Hemant K. Kashyap. Molecular dynamics study of nanoscale organization and hydrogen bonding in binary mixtures of butylammonium nitrate ionic liquid and primary alcohols. The Journal of Chemical Physics 2017, 146
(6)
https://doi.org/10.1063/1.4975172
- Olga Russina, Alessandro Triolo. Ionic Liquids and Neutron Scattering. 2017, 213-278. https://doi.org/10.1016/B978-0-12-805324-9.00004-2
- Saffron J. Bryant, Kathleen Wood, Rob Atkin, Gregory G. Warr. Effect of protic ionic liquid nanostructure on phospholipid vesicle formation. Soft Matter 2017, 13
(7)
, 1364-1370. https://doi.org/10.1039/C6SM02652D
- Andrei Filippov, Oleg I. Gnezdilov, Nicklas Hjalmarsson, Oleg N. Antzutkin, Sergei Glavatskih, István Furó, Mark W. Rutland. Acceleration of diffusion in ethylammonium nitrate ionic liquid confined between parallel glass plates. Physical Chemistry Chemical Physics 2017, 19
(38)
, 25853-25858. https://doi.org/10.1039/C7CP01772C
- Lang G. Chen, Stephen H. Strassburg, Harry Bermudez. Micelle co-assembly in surfactant/ionic liquid mixtures. Journal of Colloid and Interface Science 2016, 477 , 40-45. https://doi.org/10.1016/j.jcis.2016.05.020
- Thomas Zemb, Werner Kunz. Weak aggregation: State of the art, expectations and open questions. Current Opinion in Colloid & Interface Science 2016, 22 , 113-119. https://doi.org/10.1016/j.cocis.2016.04.002
- Thomas Murphy, Samantha K. Callear, Gregory G. Warr, Rob Atkin. Dissolved chloride markedly changes the nanostructure of the protic ionic liquids propylammonium and ethanolammonium nitrate. Physical Chemistry Chemical Physics 2016, 18
(26)
, 17169-17182. https://doi.org/10.1039/C5CP06947E
- Thomas Murphy, Robert Hayes, Silvia Imberti, Gregory G. Warr, Rob Atkin. Ionic liquid nanostructure enables alcohol self assembly. Physical Chemistry Chemical Physics 2016, 18
(18)
, 12797-12809. https://doi.org/10.1039/C6CP01739H
- A. Sanchez-Fernandez, K. J. Edler, T. Arnold, R. K. Heenan, L. Porcar, N. J. Terrill, A. E. Terry, A. J. Jackson. Micelle structure in a deep eutectic solvent: a small-angle scattering study. Physical Chemistry Chemical Physics 2016, 18
(20)
, 14063-14073. https://doi.org/10.1039/C6CP01757F
- Martina Požar, Bernarda Lovrinčević, Larisa Zoranić, Marijana Mijaković, Franjo Sokolić, Aurélien Perera. . The Journal of Chemical Physics 2016https://doi.org/10.1063/1.4960435
- Tamar L. Greaves, Danielle F. Kennedy, Yan Shen, Asoka Weerawardena, Adrian Hawley, Gonghua Song, Calum J. Drummond. Fluorous protic ionic liquid exhibits a series of lyotropic liquid crystalline mesophases upon water addition. Journal of Molecular Liquids 2015, 210 , 279-285. https://doi.org/10.1016/j.molliq.2015.03.037
- Thomas Murphy, Rob Atkin, Gregory G. Warr. Scattering from ionic liquids. Current Opinion in Colloid & Interface Science 2015, 20
(4)
, 282-292. https://doi.org/10.1016/j.cocis.2015.10.004
- O. Pizio, W. Rżysko, S. Sokołowski, Z. Sokołowska. Mixtures of ions and amphiphilic molecules in slit-like pores: A density functional approach. The Journal of Chemical Physics 2015, 142
(16)
https://doi.org/10.1063/1.4918640
- Adam H. Turner, John D. Holbrey. The Solution Structure of 1:2 Phenol/N-Methylpyridinium bis{(trifluoromethyl)sulfonyl}imide Liquid Mixtures. Journal of Solution Chemistry 2015, 44
(3-4)
, 621-633. https://doi.org/10.1007/s10953-015-0296-2
- Aaron Elbourne, Samuel Cronshaw, Kislon Voïtchovsky, Gregory G. Warr, Rob Atkin. Near surface properties of mixtures of propylammonium nitrate with n-alkanols 1. Nanostructure. Physical Chemistry Chemical Physics 2015, 17
(40)
, 26621-26628. https://doi.org/10.1039/C5CP04786B
- James Sweeney, Grant B. Webber, Rob Atkin. Near surface properties of mixtures of propylammonium nitrate with n-alkanols 2. Nanotribology and fluid dynamics. Physical Chemistry Chemical Physics 2015, 17
(40)
, 26629-26637. https://doi.org/10.1039/C5CP04787K
- Andrew Dolan, Rob Atkin, Gregory G. Warr. The origin of surfactant amphiphilicity and self-assembly in protic ionic liquids. Chemical Science 2015, 6
(11)
, 6189-6198. https://doi.org/10.1039/C5SC01202C
Article Views are the COUNTER-compliant sum of full text article downloads since November 2008 (both PDF and HTML) across all institutions and individuals. These metrics are regularly updated to reflect usage leading up to the last few days.
Citations are the number of other articles citing this article, calculated by Crossref and updated daily. Find more information about Crossref citation counts.
The Altmetric Attention Score is a quantitative measure of the attention that a research article has received online. Clicking on the donut icon will load a page at altmetric.com with additional details about the score and the social media presence for the given article. Find more information on the Altmetric Attention Score and how the score is calculated.
Recommended Articles
References
This article references 76 other publications.
- 1Rogers, R. D.; Seddon, K. R. Ionic Liquids - Solvents of the Future Science 2003, 302, 792– 793There is no corresponding record for this reference.
- 2MacFarlane, D. R.; Seddon, K. R. Ionic Liquids—Progress on the Fundamental Issues Aust. J. Chem. 2007, 60, 3– 5There is no corresponding record for this reference.
- 3Earle, M. J.; Seddon, K. R. Pure Appl. Chem. 2000, 72, 1391– 1398There is no corresponding record for this reference.
- 4Welton, T. Room-Temperature Ionic Liquids. Solvents for Synthesis and Catalysis Chem. Rev. 1999, 99, 2071– 20834https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK1MXkt1artrw%253D&md5=e17d4c2a7f45438755b34161e86f24e6Room-Temperature Ionic Liquids. Solvents for Synthesis and CatalysisWelton, ThomasChemical Reviews (Washington, D. C.) (1999), 99 (8), 2071-2083CODEN: CHREAY; ISSN:0009-2665. (American Chemical Society)A review with 124 refs. covering org. reactions in alkylhalo- and haloaluminate ionic liqs.
- 5Swatloski, R. P.; Visser, A. E.; Reichert, W. M.; Broker, G. A.; Farina, L. M.; Holbrey, J. D.; Rogers, R. D. Solvation of 1-Butyl-3-methylimidazolium Hexafluorophosphate in Aqueous Ethanol - a Green Solution for Dissolving ’Hydrophobic’ Ionic Liquids Chem. Commun. (Cambridge, U. K.) 2001, 2070– 20715https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3MXnsFWqu70%253D&md5=b1b8650a731fb7c3f970caf1839e6960Solvation of 1-butyl-3-methylimidazolium hexafluorophosphate in aqueous ethanol-a green solution for dissolving 'hydrophobic' ionic liquidsSwatloski, Richard P.; Visser, Ann E.; Reichert, W. Matthew; Broker, Grant A.; Farina, Lindsy M.; Holbrey, John D.; Rogers, Robin D.Chemical Communications (Cambridge, United Kingdom) (2001), (20), 2070-2071CODEN: CHCOFS; ISSN:1359-7345. (Royal Society of Chemistry)The relatively hydrophobic ionic liq. 1-butyl-3-methylimidazolium hexafluorophosphate has been found to be totally miscible with aq. ethanol between 0.5 and 0.9 mol fraction ethanol, whereas the ionic liq. is only partially miscible with either pure water or abs. ethanol. The ability to dissolve 1-butyl-3-methylimidazolium hexafluorophosphate in a 'green' aq. solvent system has important implications for cleaning, purifn., and sepns. using ionic liqs.
- 6Huddleston, J. G.; Visser, A. E.; Reichert, W. M.; Willauer, H. D.; Broker, G. A.; Rogers, R. D. Characterization and Comparison of Hydrophilic and Hydrophobic Room Temperature Ionic Liquids Incorporating the Imidazolium Cation Green Chem. 2001, 3, 156– 1646https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3MXmtVWhsr8%253D&md5=db13c3edffe333bc489b6ceeccf3bbbcCharacterization and comparison of hydrophilic and hydrophobic room temperature ionic liquids incorporating the imidazolium cationHuddleston, Jonathan G.; Visser, Ann E.; Reichert, W. Matthew; Willauer, Heather D.; Broker, Grant A.; Rogers, Robin D.Green Chemistry (2001), 3 (4), 156-164CODEN: GRCHFJ; ISSN:1463-9262. (Royal Society of Chemistry)A series of hydrophilic and hydrophobic 1-alkyl-3-methylimidazolium room temp. ionic liqs. (RTILs) have been prepd. and characterized to det. how water content, d., viscosity, surface tension, m.p., and thermal stability are affected by changes in alkyl chain length and anion. In the series of RTILs studied here, the choice of anion dets. water miscibility and has the most dramatic effect on the properties. Hydrophilic anions (e.g., chloride and iodide) produce ionic liqs. that are miscible in any proportion with water but, upon the removal of some water from the soln., illustrate how sensitive the phys. properties are to a change in water content. In comparison, for ionic liqs. contg. more hydrophobic anions (e.g., PF6- and N(SO2CF3)2-), the removal of water has a smaller affect on the resulting properties. For a series of 1-alkyl-3-methylimidazolium cations, increasing the alkyl chain length from Bu to hexyl to octyl increases the hydrophobicity and the viscosities of the ionic liqs. increase, whereas densities and surface tension values decrease. Thermal analyses indicate high temps. are attainable prior to decompn. and DSC studies reveal a glass transition for several samples. ILs incorporating PF6- have been used in liq./liq. partitioning of org. mols. from water and the results for two of these are also discussed here. On a cautionary note, the chem. of the individual cations and anions of the ILs should not be overlooked as, in the case of certain conditions for PF6- ILs, contact with an aq. phase may result in slow hydrolysis of the PF6- with the concomitant release of HF and other species.
- 7Anderson, J. L.; Pino, V.; Hagberg, E. C.; Sheares, V. V.; Armstrong, D. W. Surfactant Solvation Effects and Micelle Formation in Ionic Liquids Chem. Commun. (Cambridge, U. K.) 2003, 2444– 24457https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3sXntlWhsLo%253D&md5=d268c6f6fde4796153de37a3472bafa4Surfactant solvation effects and micelle formation in ionic liquidsAnderson, Jared L.; Pino, Veronica; Hagberg, Erik C.; Sheares, Valerie V.; Armstrong, Daniel W.Chemical Communications (Cambridge, United Kingdom) (2003), (19), 2444-2445CODEN: CHCOFS; ISSN:1359-7345. (Royal Society of Chemistry)The formation of micelles in 1-butyl-3-Me imidazolium chloride (BMIM-Cl) and hexafluorophosphate (BMIM-PF6) were explored using different surfactants and the solvation behavior of the new micellar-ionic liq. solns. examd. using inverse gas chromatog.
- 8Li, N.; Zhang, S.; Ma, H.; Zheng, L. Role of Solubilized Water in Micelles Formed by Triton X-100 in 1-Butyl-3-methylimidazolium Ionic Liquids Langmuir 2010, 26, 9315– 9320There is no corresponding record for this reference.
- 9He, Y.; Li, Z.; Simone, P.; Lodge, T. P. Self-Assembly of Block Copolymer Micelles in an Ionic Liquid J. Am. Chem. Soc. 2006, 128, 2745– 27509https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD28XpsVKisA%253D%253D&md5=aa9a5a20e2e14c37ce5ddb8de6bfcfb0Self-Assembly of Block Copolymer Micelles in an Ionic LiquidHe, Yiyong; Li, Zhibo; Simone, Peter; Lodge, Timothy P.Journal of the American Chemical Society (2006), 128 (8), 2745-2750CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)Four amphiphilic poly((1,2-butadiene)-block-ethylene oxide) (PB-PEO) diblock copolymers were shown to aggregate strongly and form micelles in an ionic liq., 1-butyl-3-methylimidazolium hexafluorophosphate ([BMIM][PF6]). The universal micellar structures (spherical micelle, wormlike micelle, and bilayered vesicle) were all accessed by varying the length of the corona block while holding the core block const. The nanostructures of the PB-PEO micelles formed in an ionic liq. were directly visualized by cryogenic transmission electron microscopy (cryo-TEM). Detailed micelle structural information was extd. from both cryo-TEM and dynamic light scattering measurements, with excellent agreement between the two techniques. Compared to aq. solns. of the same copolymers, [BMIM][PF6] solns. exhibit some distinct features, such as temp.-independent micellar morphologies between 25° and 100°. As in aq. solns., significant nonergodicity effects were also obsd. This work demonstrates the flexibility of amphiphilic block copolymers for controlling nanostructure in an ionic liq., with potential applications in many arenas.
- 10Sharma, S. C.; Atkin, R.; Warr, G. G. The Effect of Ionic Liquid Hydrophobicity and Solvent Miscibility on Pluronic Amphiphile Self-Assembly J. Phys. Chem. B 2013, 117, 14568– 14575There is no corresponding record for this reference.
- 11Yue, X.; Chen, X.; Wang, X.; Li, Z. Lyotropic Liquid Crystalline Phases Formed by Phyosterol Ethoxylates in Room-Temperature Ionic Liquids Colloids Surf., A 2011, 392, 225– 23211https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3MXhsFKrsrjO&md5=58b3dd8eedd05cbfcbba4069fe5184bdLyotropic liquid crystalline phases formed by phyosterol ethoxylates in room-temperature ionic liquidsYue, Xiu; Chen, Xiao; Wang, Xudong; Li, ZhihongColloids and Surfaces, A: Physicochemical and Engineering Aspects (2011), 392 (1), 225-232CODEN: CPEAEH; ISSN:0927-7757. (Elsevier B.V.)The phase behaviors of four phytosterol ethoxylates surfactants (BPS-n, n = 5, 10, 20, and 30) with different oxyethylene units in room temp. ionic liq., 1-butyl-3-methylimidazolium tetrafluoroborate ([Bmim]BF4), have been studied. The polarized optical microscopy and small-angle X-ray scattering techniques are used to characterize the phase structures of these binary systems at 25 °C. The structure and ordering of the liq. cryst. (LC) phases in such BPS-n/[Bmim]BF4 systems are found to be influenced by BPS-n concn. and the temp. Due to the bulky and rigid cholesterol group, the phytosterol ethoxylates surfactants exhibit different properties and interaction mechanism from the conventional CnEOm type nonionic surfactant systems. The rheol. measurements indicate a highly viscoelastic nature of these lyotropic LC phases and disclose a lamellar phase characteristic with a rather strong rigidity at high surfactant concns. The control expt. with Brij 97(polyoxyethylene (10) oleyl ether)/[Bmim]BF4 system and the FTIR measurements help to recognize that the solvophobic interaction combining with the hydrogen bonding are the main driving forces for the LC phases formation.
- 12Sakai, H.; Saitoh, T.; Misono, T.; Tsuchiya, K.; Sakai, K.; Abe, M. Phase Behavior of Phytosterol Ethoxylates in an Imidazolium-Type Room-Temperature Ionic Liquid J. Oleo Sci. 2012, 61, 135– 14112https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38XjsV2ltrk%253D&md5=e7cfefb1579bbb42f28d8cc597e8cabePhase behavior of phytosterol ethoxylates in an imidazolium-type room-temperature ionic liquidSakai, Hideki; Saitoh, Takanori; Misono, Takeshi; Tsuchiya, Koji; Sakai, Kenichi; Abe, MasahikoJournal of Oleo Science (2012), 61 (3), 135-141CODEN: JOSOAP; ISSN:1345-8957. (Japan Oil ChemistsÏ Society)The temp.-concn. phase behavior of nonionic surfactants in an aprotic imidazolium-type room-temp. ionic liq. (RT-IL) was evaluated on the basis of a combination of visual appearance, polarized optical microscopy, and small angle X-ray scattering data. Phytosterol ethoxylates (BPS-n, where n denotes oxyethylene chain lengths of 5, 10, 20, and 30) were used as surfactants in the RT-IL, 1-butyl-3-methylimidazolium hexafluorophosphate (BmimPF6). The two component mixts. yielded various phases such as discontinuous cubic, hexagonal, and lamellar phases. An increased tendency toward formation of lesser-curved mol. assemblies was obsd. at higher BPS-n concns., at lower temps., and for shorter oxyethylene chain surfactants. These trends are similar to those obsd. in aq. BPS-n systems; however, notable differences in the phase states of the aq. system vs. the BmimPF6 system were evident. Comparison with the water system showed that the BmimPF6 system yielded fewer phases and generally required higher BPS-n concns. to induce phase transitions. Evaluation of the effects of addn. of a third component (e.g., 1-dodecanol and dodecane) to the binary system on the phase behavior showed that at a given compn. ratio of BPS-20 to BmimPF6, the addn. of 1-dodecanol generally results in the phase transition to lesser-curved assemblies whereas dodecane generated no significant effects. The obsd. phase change is satisfactorily rationalized by localized solubilization of the third component into the binary surfactant assemblies.
- 13Wang, L. Y.; Chen, X.; Chai, Y. C.; Hao, J. C.; Sui, Z. M.; Zhuang, W. C.; Sun, Z. W. Lyotropic Liquid Crystalline Phases Formed in an Ionic Liquid Chem. Commun. (Cambridge, U. K.) 2004, 24, 2840– 2841There is no corresponding record for this reference.
- 14Anjum, N.; Guedeau-Boudeville, M.-A.; Stubenrauch, C.; Mourchid, A. Phase Behavior and Microstructure of Microemulsions Containing the Hydrophobic Ionic Liquid 1-Butyl-3-methylimidazolium Hexafluorophosphate J. Phys. Chem. B 2009, 113, 239– 244There is no corresponding record for this reference.
- 15Gao, H. X.; Li, J. C.; Han, B. X.; Chen, W. N.; Zhang, J. L.; Zhang, R.; Yan, D. D. Microemulsions with Ionic Liquid Polar Domains Phys. Chem. Chem. Phys. 2004, 6, 2914– 291615https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2cXkt1arsbs%253D&md5=6c6c5027c64ba1dc07fc5977b3a74a43Microemulsions with ionic liquid polar domainsGao, Haixiang; Li, Junchun; Han, Buxing; Chen, Wenna; Zhang, Jianling; Zhang, Rui; Yan, DadongPhysical Chemistry Chemical Physics (2004), 6 (11), 2914-2916CODEN: PPCPFQ; ISSN:1463-9076. (Royal Society of Chemistry)In this work 1-butyl-3-methylimidazolium tetrafluoroborate ([bmim][BF4], an ionic liq.)/Triton X-100 (TX-100)/cyclohexane microemulsions have been prepd. and characterized by phase behavior, cond. measurement, dynamic light scattering, freeze-fracturing electron microscopy, and UV-vis techniques, and our attention is concd. on the microemulsions with the ionic liq. as the nano-sized polar domains.
- 16Gao, Y. N.; Han, S. B.; Han, B. X.; Li, G. Z.; Shen, D.; Li, Z. H.; Du, J. M.; Hou, W. G.; Zhang, G. Y. TX-100/water/1-butyl-3-methylimidazolium Hexafluorophosphate Microemulsions Langmuir 2005, 21, 5681– 5684There is no corresponding record for this reference.
- 17Walden, P. Über die Molekulargröβe und elektrische Leitfähigkeit Einiger Geschmolzener Salze Bull. Acad. Imp. Sci. St.-Petersbourg 1914, 8, 405– 422There is no corresponding record for this reference.
- 18Belieres, J. P.; Angell, C. A. Protic Ionic Liquids: Preparation, Characterization, and Proton Free Energy Level Representation J. Phys. Chem. B 2007, 111, 4926– 493718https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2sXjvFCitb8%253D&md5=465b63065f29b7b0fcd88d76a1c83dd6Protic ionic liquids: preparation, characterization, and proton free energy level representationBelieres, Jean-Philippe; Angell, C. AustenJournal of Physical Chemistry B (2007), 111 (18), 4926-4937CODEN: JPCBFK; ISSN:1520-6106. (American Chemical Society)The authors give a perspective on the relations between inorg. and org. cation ionic liqs. (ILs), including members with m.ps. that overlap around the borderline 100 °C. The paper presents the synthesis and properties (melting, boiling, glass temps., etc.) of a large no. of an intermediate group of liqs. that cover the ground between equimolar mol. mixts. and ILs, depending on the energetics of transfer of a proton from one member of the pair to the other. These proton-transfer ILs have interesting properties, including the ability to serve as electrolytes in solvent-free fuel cell systems. This work provides a basis for assessing their relation to aprotic ILs by means of a Gurney-type proton-transfer free energy level diagram, with approx. values of the energy levels based on free energy of formation and pKa data. The energy level scheme allows verifying the relation between solvent-free acidic and basic electrolytes, and the familiar aq. variety, and to identify neutral protic electrolytes that are unavailable in the case of aq. systems.
- 19Yoshizawa, M.; Xu, W.; Angell, C. A. Ionic Liquids by Proton Transfer: Vapor Pressure, Conductivity, and the Relevance of Delta pKa from Aqueous Solutions J. Am. Chem. Soc. 2003, 125, 15411– 15419There is no corresponding record for this reference.
- 20Greaves, T. L.; Drummond, C. J. Protic Ionic Liquids: Properties and Applications Chem. Rev. 2008, 108, 206– 23720https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2sXhsVCrsb%252FL&md5=16a3f06eef103a54dc1e38f9e469506bProtic Ionic Liquids: Properties and ApplicationsGreaves, Tamar L.; Drummond, Calum J.Chemical Reviews (Washington, DC, United States) (2008), 108 (1), 206-237CODEN: CHREAY; ISSN:0009-2665. (American Chemical Society)This review covers the ionicity, physicochem. and thermal properties of protic ionic liqs., along with a description of their applications, and where they have been used, including org. synthesis, chromatog., biol. applications, fuel cells, explosives, and, very recently, industrial lubricants.
- 21Angell, C. A.; Byrne, N.; Belieres, J.-P. Parallel Developments in Aprotic and Protic Ionic Liquids: Physical Chemistry and Applications Acc. Chem. Res. 2007, 40, 1228– 123621https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2sXht1GmtrnJ&md5=8e0c1cb0803e1730718477540b13174aParallel Developments in Aprotic and Protic Ionic Liquids: Physical Chemistry and ApplicationsAngell, C. Austen; Byrne, Nolene; Belieres, Jean-PhilippeAccounts of Chemical Research (2007), 40 (11), 1228-1236CODEN: ACHRE4; ISSN:0001-4842. (American Chemical Society)A review. This account covers research dating from the early 1960s in the field of low-melting molten salts and hydrates, which has recently become popular under the rubric of "ionic liqs.". It covers understanding gained in the principal author's labs. (initially in Australia, but mostly in the U.S.A.) from spectroscopic, dynamic, and thermodn. studies and includes recent applications of this understanding in the fields of energy conversion and biopreservation. Both protic and aprotic varieties of ionic liqs. are included, but recent studies have focused on the protic class because of the special applications made possible by the highly variable proton activities available in these liqs.
- 22Evans, D. F.; Yamauchi, A.; Roman, R.; Casassa, E. Z. Micelle Formation in Ethylammonium Nitrate, a Low-Melting Fused Salt J. Colloid Interface Sci. 1982, 88, 89– 9622https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaL38XktFahtbs%253D&md5=25a5726d4d7af7f22422bfcc4eb17442Micelle formation in ethylammonium nitrate, a low-melting fused saltEvans, D. Fennell; Yamauchi, Akira; Roman, Ronald; Casassa, Ethel Z.Journal of Colloid and Interface Science (1982), 88 (1), 89-96CODEN: JCISA5; ISSN:0021-9797.Crit. micelle concns. (CMC) were detd. from surface tension measurements for alkyltrimethylammonium bromides and alkylpyridinium bromides at 50° and for Triton X-100 at 20 and 50° in ethylammonium nitrate, a low-melting anhyd. fused salt. The CMC's are ∼5 to 10 times larger than those obsd. in water. From the change of CMC with surfactant chain length, the free energy of transfer of a methylene group from the fused salt to the micelle interior is calcd. to be -370 cal/mol compared to -680 cal/mol for a similar transfer from water to the micelle. With respect to solvophobic behavior, ethylammonium nitrate and water show a no. of similarities.
- 23Evans, D. F.; Yamauchi, A.; Wei, G. J.; Bloomfield, V. A. Micelle Size in Ethylammonium Nitrate as Determined by Classical and Quasi-Elastic Light Scattering J. Phys. Chem. 1983, 87, 3537– 354123https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaL3sXkvFCit74%253D&md5=a3df15122ddfe5942e7c9fb033a43879Micelle size in ethylammonium nitrate as determined by classical and quasi-elastic light scatteringEvans, D. Fennell; Yamauchi, A.; Wei, G. Jason; Bloomfield, Victor A.Journal of Physical Chemistry (1983), 87 (18), 3537-41CODEN: JPCHAX; ISSN:0022-3654.The aggregation of surfactants to form micelles in ethylammonium nitrate, a low melting fused salt, was investigated by classical and quasi-elastic light scattering. For tetradecylpyridinium bromide and hexadecylpyridinium bromide, the following data are obtained: crit. micelle concns., 8.0 × 10-2 and 2.0 × 10-2 mol kg-1; micellar aggregation nos., 17 and 26; 2nd virial coeffs., 1.64 × 10-3 and 1.30 × 10-3 mol3 g-2; and hydrodynamic radii, 14 and 22 Å, resp. The results are consistent with either a small classical spherical micelle contg. only surfactant or a spherical mixed micelle contg. surfactant and ethylammonium ions as a cosurfactant. The 2nd virial coeffs. are almost equal to those calcd. for hard spheres and reflect highly screened electrostatic interactions in the totally ionized solvent.
- 24Araos, M. U.; Warr, G. G. Self-Assembly of Nonionic Surfactants into Lyotropic Liquid Crystals in Ethylammonium Nitrate, a Room-Temperature Ionic Liquid J. Phys. Chem. B 2005, 109, 14275– 1427724https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2MXmtVWisL4%253D&md5=9d6f0f5f1cfff23f7947cec06cd3d309Self-Assembly of Nonionic Surfactants into Lyotropic Liquid Crystals in Ethylammonium Nitrate, a Room-Temperature Ionic LiquidAraos, Miguel U.; Warr, Gregory G.Journal of Physical Chemistry B (2005), 109 (30), 14275-14277CODEN: JPCBFK; ISSN:1520-6106. (American Chemical Society)The stability of a variety of lyotropic liq. crystals formed by a no. of polyoxyethylene nonionic surfactants in the room-temp. ionic liq. ethylammonium nitrate (EAN) is surveyed and reported. The pattern of self-assembly behavior and mesophase formation is strikingly similar to that obsd. in water, even including the existence of a lower consolute boundary or cloud point. The only quant. difference from water is that longer alkyl chains are necessary to drive the formation of liq. cryst. mesophases in EAN, suggesting that a rich pattern of "solvophobic" self-assembly should exist in this solvent.
- 25Araos, M. U.; Warr, G. G. Structure of Nonionic Surfactant Micelles in the Ionic Liquid Ethylammonium Nitrate Langmuir 2008, 24, 9354– 936025https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1cXpt1yntr4%253D&md5=98b23022b0fcfdd828838b4926412e1cStructure of Nonionic Surfactant Micelles in the Ionic Liquid Ethylammonium NitrateAraos, Miguel U.; Warr, Gregory G.Langmuir (2008), 24 (17), 9354-9360CODEN: LANGD5; ISSN:0743-7463. (American Chemical Society)The structure of micelles formed by nonionic polyoxyethylene alkyl ether nonionic surfactants, CnEm, in room-temp. ionic liq., ethylammonium nitrate (EAN), was detd. by small-angle neutron scattering (SANS) as a function of alkyl and ethoxy chain length, concn., and temp. Micelles form for all surfactant alkyl chain lengths from dodecyl through to octadecyl. The dodecyl-chain surfactants have high crit. micelle concn., around 1%, and form weakly structured micelles. Surfactants with longer alkyl chains readily form micelles in EAN. The obsd. micelle structure changes systematically with alkyl and ethoxy chain length, in parallel with observations in aq. solns. Decreasing ethoxy chain length at const. alkyl chain length leads to a sphere to rod transition. These micelles also grow into rods with increasing temp. as their cloud point is approached in EAN.
- 26Atkin, R.; Warr, G. G. Phase Behavior and Microstructure of Microemulsions with a Room-Temperature Ionic Liquid as the Polar Phase J. Phys. Chem. B 2007, 111, 9309– 931626https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2sXns12ltb4%253D&md5=df6c503f69753f9b50aa817e4555795aPhase Behavior and Microstructure of Microemulsions with a Room-Temperature Ionic Liquid as the Polar PhaseAtkin, Rob; Warr, Gregory G.Journal of Physical Chemistry B (2007), 111 (31), 9309-9316CODEN: JPCBFK; ISSN:1520-6106. (American Chemical Society)Microemulsions of nonionic alkyl oligoethyleneoxide (CiEj) surfactants, alkanes, and ethylammonium nitrate (EAN), a room-temp. ionic liq., have been prepd. and characterized. Studies of phase behavior reveal that EAN microemulsions have many features in common with corresponding aq. systems, the primary difference being that higher surfactant concns. and longer surfactant tailgroups are required to offset the decreased solvophobicity the surfactant mols. in EAN compared with water. The response of the EAN microemulsions to variation in the length of the alkane, surfactant headgroup, and surfactant tailgroup has been found to parallel that obsd. in aq. systems in most instances. EAN microemulsions exhibit a single broad small-angle X-ray scattering peak, like aq. systems. These are well described by the Teubner-Strey model. A lamellar phase was also obsd. for surfactants with longer tails at lower temps. The scattering peaks of both microemulsion and lamellar phases move to lower wave vector on increasing temp. This is ascribed to a decrease in the interfacial area of the surfactant layer. Phase behavior, small-angle X-ray scattering, and cond. expts. have allowed the weakly to strongly structured transition to be identified for EAN systems.
- 27Atkin, R.; De Fina, L.-M.; Kiederling, U.; Warr, G. G. Structure and Self Assembly of Pluronic Amphiphiles in Ethylammonium Nitrate and at the Silica Surface J. Phys. Chem. B 2009, 113, 12201– 12213There is no corresponding record for this reference.
- 28Velasco, S. B.; Turmine, M.; Di Caprio, D.; Letellier, P. Micelle Formation in Ethyl-ammonium Nitrate (an Ionic Liquid) Colloids Surf., A 2006, 275, 50– 54There is no corresponding record for this reference.
- 29Evans, D. F.; Kaler, E. W.; Benton, W. J. Liquid Crystals in a Fused Salt: β,γ-Distearoylphosphatidylcholine in N-Ethylammonium Nitrate J. Phys. Chem. 1983, 87, 533– 535There is no corresponding record for this reference.
- 30Greaves, T. L.; Weerawardena, A.; Fong, C.; Drummond, C. J. Many Protic Ionic Liquids Mediate Hydrocarbon-Solvent Interactions and Promote Amphiphile Self-Assembly Langmuir 2007, 23, 402– 40430https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD28Xht12qt77M&md5=b664062a092ce3c7dd5ce1c6f13b2ef7Many Protic Ionic Liquids Mediate Hydrocarbon-Solvent Interactions and Promote Amphiphile Self-AssemblyGreaves, Tamar L.; Weerawardena, Asoka; Fong, Celesta; Drummond, Calum J.Langmuir (2007), 23 (2), 402-404CODEN: LANGD5; ISSN:0743-7463. (American Chemical Society)A large no. of protic ionic liqs. (PILs) have been found to mediate solvent-hydrocarbon interactions and promote amphiphile self-assembly. Hexagonal, cubic, and lamellar lyotropic liq. cryst. phases were obsd. in PIL-hexadecyltrimethylammonium bromide systems. The driving force for the formation of the self-assembled aggregate structures has been attributed to an entropic contribution to the free energy of assocn., analogous to the hydrophobic effect in water. The specific aggregate structures formed depend upon the cationic and anionic components of the PIL and their interactions with the amphiphiles.
- 31Evans, D. F.; Chen, S.-H.; Schriver, G. W.; Arnett, E. M. Thermodynamics of Solution of Nonpolar Gases in a Fused Salt. Hydrophobic Bonding Behavior in a Nonaqueous System J. Am. Chem. Soc. 1981, 103, 481– 482There is no corresponding record for this reference.
- 32Hayes, R.; Imberti, S.; Warr, G. G.; Atkin, R. The Nature of Hydrogen Bonding in Protic Ionic Liquids Angew. Chem., Int. Ed. 2013, 52, 4623– 462732https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3sXjtlaksLs%253D&md5=726d2360de55ca62fee0d747e702834fThe Nature of Hydrogen Bonding in Protic Ionic LiquidsHayes, Robert; Imberti, Silvia; Warr, Gregory G.; Atkin, RobAngewandte Chemie, International Edition (2013), 52 (17), 4623-4627CODEN: ACIEF5; ISSN:1433-7851. (Wiley-VCH Verlag GmbH & Co. KGaA)The ion arrangements in protic ionic liqs. (PILs) are a consequence of the balance between intermol. forces and the phys. dimensions of the ions. While similar bicontinuous sponge-like nanostructures form in all systems, there is significant variation in the H-bonding. The PIL nanostructure controls the H-bond strength and structure by dictating the relative orientations of the cations and anions to each other. The distribution of H-bond geometries is related to the ability of each PIL to accommodate H-bonds in the bicontinuous arrangement rather than inducing a different bulk structure. When the nanostructures have a relatively high proportion of H-bonds are linear (RX:HN ≈ 1), attractions between ions increase and phys. properties become more solid-like. At higher RX:HN ratios, bi-or trifurcated H-bonding results, leading to weaker, bent- H-bonds, decreasing cation-anion attractions and leading to a more fluid-like material.
- 33Atkin, R.; Warr, G. G. The Smallest Amphiphiles: Nanostructure in Protic Room-Temperature Ionic Liquids with Short Alkyl Groups J. Phys. Chem. B 2008, 112, 4164– 416633https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1cXjtlOjsb0%253D&md5=0f0a37486d4fb146d73c2e7435cc711dThe Smallest Amphiphiles: Nanostructure in Protic Room-Temperature Ionic Liquids with Short Alkyl GroupsAtkin, Rob; Warr, Gregory G.Journal of Physical Chemistry B (2008), 112 (14), 4164-4166CODEN: JPCBFK; ISSN:1520-6106. (American Chemical Society)Room-temp. ionic liqs. (ILs) are low-melting-point org. salts that, until recently, were thought to have homogeneous microstructure. In this work, we investigate nanoscale segregation of short (<C4) alkyl chain ILs using propylammonium nitrate (PAN) and ethylammonium nitrate (EAN). Structure peaks at q = 0.54 Å-1 for PAN and at q = 0.66 Å-1 for EAN, corresponding to Bragg spacings (D* = 2π/qmax) of 11.6 and 9.7 Å resp., provide the first exptl. evidence of nanoscale heterogeneity for ILs with alkyl chains less than C4. The observation that these ILs are not optically birefringent and the fits obtained suggest a disordered, locally smectic or sponge-like structure. Solvophobic interaction between alkyl groups is the most important factor for the prodn. of nanoscale heterogeneities, but electrostatic and hydrogen bonding attractions between the amine nitrogen and the nitrate anion will also play a role.
- 34Hayes, R.; Imberti, S.; Warr, G. G.; Atkin, R. Amphiphilicity Determines Nanostructure in Protic Ionic Liquids Phys. Chem. Chem. Phys. 2011, 13, 3237– 324734https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3MXhsFymtL4%253D&md5=aa28e5cc0144742a23aee635ad0f3567Amphiphilicity determines nanostructure in protic ionic liquidsHayes, Robert; Imberti, Silvia; Warr, Gregory G.; Atkin, RobPhysical Chemistry Chemical Physics (2011), 13 (8), 3237-3247CODEN: PPCPFQ; ISSN:1463-9076. (Royal Society of Chemistry)The bulk structure of the two oldest ionic liqs. (ILs), ethylammonium nitrate (EAN) and ethanolammonium nitrate (EtAN), is elucidated using neutron diffraction. The spectra were modelled using empirical potential structure refinement (EPSR). EAN exhibits a long-range structure of solvophobic origin, similar to a bicontinuous microemulsion or disordered L3-sponge phase, but with a domain size of only 1 nm. The alc. (-OH) moiety in EtAN interferes with solvophobic assocn. between cation alkyl chains resulting in small clusters of ions, rather than an extended network.
- 35Hayes, R.; Imberti, S.; Warr, G. G.; Atkin, R. Pronounced Sponge-like Nanostructure in Propylammonium Nitrate Phys. Chem. Chem. Phys. 2011, 13, 13544– 1355135https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3MXptVyrs78%253D&md5=3589b22ad9cfb74c67a2c2123077f587Pronounced sponge-like nanostructure in propylammonium nitrateHayes, Robert; Imberti, Silvia; Warr, Gregory G.; Atkin, RobPhysical Chemistry Chemical Physics (2011), 13 (30), 13544-13551CODEN: PPCPFQ; ISSN:1463-9076. (Royal Society of Chemistry)The bulk structure of the ionic liq. propylammonium nitrate (PAN) was detd. using neutron diffraction. Empirical potential structure refinement (EPSR) fits to the data show that PAN self-assembles into a quasi-periodic bicontinuous nanostructure reminiscent of an amphiphile L3-sponge phase. Atomic detail on the ion arrangements around the propylammonium cation and nitrate anion yields evidence of hydrogen bonding between ammonium and nitrate groups and of strong alkyl chain aggregation and interdigitation. The resultant amphiphilic PAN nanostructure is more pronounced than that previously detd. for ethylammonium nitrate (EAN) or ethanolammonium nitrate (EtAN).
- 36Greaves, T. L.; Kennedy, D. F.; Mudie, S. T.; Drummond, C. J. Diversity Observed in the Nanostructure of Protic Ionic Liquids J. Phys. Chem. B 2010, 114, 10022– 1003136https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3cXptVSmuro%253D&md5=a217468aa9c8707d0dcc9eaa5262b09bDiversity Observed in the Nanostructure of Protic Ionic LiquidsGreaves, Tamar L.; Kennedy, Danielle F.; Mudie, Stephen T.; Drummond, Calum J.Journal of Physical Chemistry B (2010), 114 (31), 10022-10031CODEN: JPCBFK; ISSN:1520-6106. (American Chemical Society)The nanostructure of a series of 20 protic ionic liqs. (PILs) has been investigated using small- and wide-angle X-ray scattering (SAXS and WAXS). The PILs contained alkylammonium, dialkylammonium, trialkylammonium, and cyclic ammonium cations combined with org. or inorg. anions. The presence of hydroxyl and methoxy substituents on the alkyl chains of the cations was also explored. Many of the PILs showed a nanostructure resulting from segregation of the polar and nonpolar components of the ionic liq. It was found that this segregation was enhanced for longer alkyl chains, with a corresponding increase in the length scale, whereas the presence of hydroxyl groups on the alkyl chains led to much less ordered liqs. The broad range of protic ionic liqs. studied allowed several structure-property relationships to be established. The solvophobic effect was shown to be dependent on the nanostructure of the PILs. These PILs support amphiphile self-assembly, and it was shown that the less structured PILs had more "water-like" behavior in the diversity of lyotropic liq.-crystal phases supported, and the thermal stability ranges for these phases.
- 37Hayes, R.; Imberti, S.; Warr, G. G.; Atkin, R. How Water Dissolves in Protic Ionic Liquids Angew. Chem., Int. Ed. 2012, 51, 7468– 747137https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38Xos1altbo%253D&md5=af19ae3e1fef9d34bd2f38b6e06d9d97How Water Dissolves in Protic Ionic LiquidsHayes, Robert; Imberti, Silvia; Warr, Gregory G.; Atkin, RobAngewandte Chemie, International Edition (2012), 51 (30), 7468-7471, S7468/1-S7468/13CODEN: ACIEF5; ISSN:1433-7851. (Wiley-VCH Verlag GmbH & Co. KGaA)When equal masses of water and the protic ionic liq. ethylammonium nitrate are mixed a bicontinuous nanostructure results. This nanostructure resembles aq. surfactant mesophases but has length scales at least an order of magnitude smaller. The local structure of both the water and the ionic liq. are strikingly similar to that found in the pure liqs.
- 38Greaves, T. L.; Kennedy, D. F.; Weerawardena, A.; Tse, N. M. K.; Kirby, N.; Drummond, C. J. Nanostructured Protic Ionic Liquids Retain Nanoscale Features in Aqueous Solution While Precursor Bronsted Acids and Bases Exhibit Different Behavior J. Phys. Chem. B 2011, 115, 2055– 2066There is no corresponding record for this reference.
- 39Hadded, M.; Mayaffre, A.; Letellier, P. Surface-Tension Of Ideal Solutions - Application To Binary-Mixtures Of Methanol And Molten Ethylammonium Nitrate At 298-K J. Chim. Phys. Phys.-Chim. Biol. 1989, 86, 525– 53739https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaL1MXlsVWit78%253D&md5=7aba3f1284aef3e535b6c8fd8a77bbcbSurface tension of ideal solutions: application to binary mixtures of methanol and molten ethylammonium nitrate at 298 KHadded, M.; Mayaffre, A.; Letellier, P.Journal de Chimie Physique et de Physico-Chimie Biologique (1989), 86 (3), 525-37CODEN: JCPBAN; ISSN:0021-7689.Surface tension and excess vols. measurements of MeOH-molten EtNH2.HNO3 mixts. were detd. at 298 K. In terms of excess vols., the behavior of these solns. can be described approx. by the laws of ideal solns. For the surface tensions, an unusual formalism of the Gibbs equation, which involves concns. of the soln. components and a distance parameter, Δh, is proposed. The parameter Δh does not depend on the choice of surface location. This formalism is able to define the behavior of the surface for an ideal mixt. Surface tensions of MeOH-EtNH2.HNO3 mixts. are described accurately by this equation over the entire concn. range.
- 40Russina, O.; Sferrazza, A.; Caminiti, R.; Triolo, A. Amphiphile Meets Amphiphile: Beyond the Polar–Apolar Dualism in Ionic Liquid/Alcohol Mixtures J. Phys. Chem. Lett. 2014, 5 (10) 1738– 174240https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2cXntlKnu74%253D&md5=babb2d72aaacfa008b1c8ef42b2892a1Amphiphile Meets Amphiphile: Beyond the Polar-Apolar Dualism in Ionic Liquid/Alcohol MixturesRussina, Olga; Sferrazza, Alessio; Caminiti, Ruggero; Triolo, AlessandroJournal of Physical Chemistry Letters (2014), 5 (10), 1738-1742CODEN: JPCLCD; ISSN:1948-7185. (American Chemical Society)The mesoscopic morphol. of binary mixts. of ethylammonium nitrate (EAN), the protic ionic liq. par excellence, and methanol is explored using neutron/X-ray diffraction and computational techniques. Both compds. are amphiphilic and characterized by an extended hydrogen bonding network: surprisingly, though macroscopically homogeneous, these mixts. turn out to be mesoscopically highly heterogeneous. Even in methanol-rich mixts., a wide distribution of clusters exists where EAN preserves its bulk, sponge-like morphol. Accordingly, methanol does not succeed in fully dissocg. the ionic liq. that keeps on organizing in a bulk-like fashion. This behavior represents the premise to the more dramatic phenomenol. obsd. with longer alcs. that eventually phase sep. from EAN. These results challenge the commonly accepted polar and apolar moieties segregation in ionic liqs./mol. liqs. mixts. and the current understanding of technol. relevant solvation processes.
- 41Weingärtner, H. S.; Merkel, T.; Kashammer, S.; Schröer, W.; Wiegand, S. The Effect of Short-Range Hydrogen-Bonded Interactions on the Nature of the Critical-Point of Ionic Fluids. 1. General-Properties of the New System Ethylammonium Nitrate + n-Octanol with an Upper Consolute Point Near Room-Temperature Ber. Bunsen-Ges. Phys. Chem. 1993, 97, 970– 975There is no corresponding record for this reference.
- 42Weingärtner, H. S.; Schröer, W. Liquid-Liquid Phase Separations and Critical-Behavior of Electrolyte-Solutions Driven by Long-Range and Short-Range Interactions J. Mol. Liq. 1995, 65–66, 107– 114There is no corresponding record for this reference.
- 43Sahandzhieva, K.; Tuma, D.; Breyer, S.; Kamps, A. P. S.; Maurer, G. Liquid-Liquid Equilibrium in Mixtures of the Ionic Liquid 1-n-Butyl-3-methylimidazolium Hexafluorophosphate and an Alkanol J. Chem. Eng. Data 2006, 51, 1516– 1525There is no corresponding record for this reference.
- 44Vale, V. R.; Will, S.; Schroer, W.; Rathke, B. The General Phase Behavior of Mixtures of 1-Alkyl-3-Methylimidazolium Bis[(trifluoromethyl)sulfonyl]amide Ionic Liquids with n-Alkyl Alcohols ChemPhysChem 2012, 13, 1860– 1867There is no corresponding record for this reference.
- 45Greaves, T. L.; Kennedy, D. F.; Kirby, N.; Drummond, C. J. Nanostructure Changes in Protic Ionic Liquids (PILs) Through Adding Solutes and Mixing PILs Phys. Chem. Chem. Phys. 2011, 13, 13501– 13509There is no corresponding record for this reference.
- 46Greaves, T. L.; Weerawardena, A.; Fong, C.; Krodkiewska, I.; Drummond, C. J. Protic Ionic Liquids: Solvents with Tunable Phase Behavior and Physicochemical Properties J. Phys. Chem. B 2006, 110, 22479– 2248746https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD28XhtVyhsbfI&md5=dff830a123cdc27a178c01d5b6159ca0Protic Ionic Liquids: Solvents with Tunable Phase Behavior and Physicochemical PropertiesGreaves, Tamar L.; Weerawardena, Asoka; Fong, Celesta; Krodkiewska, Irena; Drummond, Calum J.Journal of Physical Chemistry B (2006), 110 (45), 22479-22487CODEN: JPCBFK; ISSN:1520-6106. (American Chemical Society)The phase behavior, including glass, devitrification, solid crystal melting, and liq. boiling transitions, and physicochem. properties, including d., refractive index, viscosity, cond., and air-liq. surface tension, of a series of 25 protic ionic liqs. and protic fused salts are presented along with structure-property comparisons. The protic fused salts were mostly liq. at room temp., and many exhibited a glass transition occurring at low temps. between -114 and -44 °C, and high fragility, with many having low viscosities, down to as low as 17 mPa·s at 25 °C, and ionic conductivities up to 43.8 S/cm at 25 °C. These protic solvents are easily prepd. through the stoichiometric combination of a primary amine and Bronsted acid. They have poor ionic behavior when compared to the far more studied aprotic ionic liqs. However, some of the other physicochem. properties possessed by these solvents are highly promising and it is anticipated that these, or analogous protic solvents, will find applications beyond those already identified for aprotic ionic liqs. This series of protic fused salts was employed to det. the effect of structural changes on the physicochem. properties, including the effect of hydroxyl groups, increasing alkyl chain lengths, branching, and the differences between inorg. and org. anions. It was found that simple structural modifications provide a mechanism to manipulate, over a wide range, the temp. at which phase transitions occur and to specifically tailor physicochem. properties for potential end-use applications.
- 47Bicak, N. A New Ionic Liquid: 2-Hydroxy Ethylammonium Formate J. Mol. Liq. 2005, 116, 15– 1847https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2cXpsVSiu7c%253D&md5=44e2bb3f4b913ed1773a438984405598A new ionic liquid: 2-hydroxyethylammonium formateBicak, NiyaziJournal of Molecular Liquids (2004), 116 (1), 15-18CODEN: JMLIDT; ISSN:0167-7322. (Elsevier B.V)A new ionic liq. (2-hydroxyethylammonium formate), with extremely low melting temp. (-82°), is presented. This ionic liq. of equimolar mixt. of formic acid and 2-hydroxyethylamine shows reasonably high at ionic cond. (3.3 mS cm-1) room temp. and heat stability up to 150°. Both 1H-NMR and FT-IR spectra establish its simple salt structure. Due to its high polarity, the ionic liq. is able to dissolve many inorg. salts. Also some insol. polymers such as polyaniline and polypyrrole are highly sol. in the ionic liq. In the study, the phys. characteristics of the ionic liq., such as cond., viscosity and solvation abilities were investigated.
- 48Wakeham, D.; Niga, P.; Ridings, C.; Andersson, G.; Nelson, A.; Warr, G. G.; Baldelli, S.; Rutland, M. W.; Atkin, R. Surface Structure of a “Non-Amphiphilic” Protic Ionic Liquid Phys. Chem. Chem. Phys. 2012, 14, 5106– 5114There is no corresponding record for this reference.
- 49Wakeham, D.; Warr, G. G.; Atkin, R. Surfactant Adsorption at the Surface of Mixed Ionic Liquids and Ionic Liquid Water Mixtures Langmuir 2012, 28, 13224– 1323149https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38Xht1egtrvI&md5=1cea2405847cd62de4930f51fabc6389Surfactant Adsorption at the Surface of Mixed Ionic Liquids and Ionic Liquid Water MixturesWakeham, Deborah; Warr, Gregory G.; Atkin, RobLangmuir (2012), 28 (37), 13224-13231CODEN: LANGD5; ISSN:0743-7463. (American Chemical Society)Surface tensiometry and neutron reflectivity were used to elucidate the structure of the adsorbed layer of nonionic surfactant tetraethylene glycol tetradecyl ether (C14E4) at the free surface of the ionic liqs. ethylammonium nitrate (EAN) and ethanolammonium nitrate (EtAN) and their binary mixts. with each other and with H2O. Surface tensions reveal that the crit. micelle concn. (cmc) depends strongly on solvent compn. The adsorbed surfactant structure elucidated by neutron reflectivity shows that the level of solvation of the ethylene oxide groups varies for both the pure and mixed solvents. This is attributed to solvent-solvent interactions dominating solvent-surfactant interactions.
- 50Greaves, T. L.; Mudie, S. T.; Drummond, C. J. 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. 2011, 13, 20441– 20452There is no corresponding record for this reference.
- 51Wakeham, D.; Eschebach, D.; Webber, G. B.; Atkin, R.; Warr, G. G. Surface Composition of Mixtures of Ethylammonium Nitrate, Ethanolammonium Nitrate and Water Aust. J. Chem. 2012, 65, 1554– 1556There is no corresponding record for this reference.
- 52Kashyap, H. K.; Hettige, J. J.; Annapureddy, H. V. R.; Margulis, C. J. SAXS Anti-peaks Reveal the Length-Scales of Dual Positive–Negative and Polar–Apolar Ordering in Room-Temperature Ionic Liquids Chem. Commun. 2012, 48, 5103– 5105There is no corresponding record for this reference.
- 53Greaves, T. L.; Drummond, C. J. Solvent Nanostructure, the Solvophobic Effect and Amphiphile Self-Assembly in Ionic Liquids Chem. Soc. Rev. 2013, 42, 1096– 112053https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3sXosV2iuw%253D%253D&md5=4426adee83b9bc3915570527a39e12abSolvent nanostructure, the solvophobic effect and amphiphile self-assembly in ionic liquidsGreaves, Tamar L.; Drummond, Calum J.Chemical Society Reviews (2013), 42 (3), 1096-1120CODEN: CSRVBR; ISSN:0306-0012. (Royal Society of Chemistry)A review. The ability of ionic liqs. (ILs) to support amphiphile self-assembly into a range of mesophase structures was established as a widespread phenomenon. From the ILs evaluated as self-assembly media, the vast majority have supported some lyotropic liq. crystal phase formation. Many neat ionic liqs. segregate into polar and nonpolar domains to form a nanostructured liq. A very strong correlation between the nanostructure of the ionic liq. and its characteristics as an amphiphile self-assembly solvent was found. In this review ionic liqs. as amphiphile self-assembly media, and identify trends that can be used to distinguish which ionic liqs. probably have good promotion properties as self-assembly media are discussed. In particular these trends focus on the nanostructure of neat ionic liqs., their solvent cohesive energy d., and the related solvophobic effect. The authors forecast that many more ILs will be identified as amphiphile self-assembly solvents in the future.
- 54Stewart, G. W.; Morrow, R. M. X-ray Diffraction in Liquids Primary Normal Alcohols Phys. Rev. 1927, 30, 232– 244There is no corresponding record for this reference.
- 55Vahvaselka, K. S.; Serimaa, R.; Torkkeli, M. Determination of Liquid Structures of the Primary Alcohols Methanol, Ethanol, 1-Propanol, 1-Butanol and 1-Octanol By X-Ray-Scattering J. Appl. Crystallogr. 1995, 28, 189– 19555https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK2MXltVektb8%253D&md5=8a9cf507c6d6b681ae0e566a2dec1505Determination of liquid structures of the primary alcohols methanol, ethanol, 1-propanol, 1-butanol and 1-octanol by x-ray scatteringVahvaselka, Kaarlo Sakari; Serimaa, Ritva; Torkkeli, MikaJournal of Applied Crystallography (1995), 28 (2), 189-95CODEN: JACGAR; ISSN:0021-8898. (Munksgaard)The four first primary alcs. and 1-octanol were studied with the wide-angle X-ray scattering method using both transmission and reflection techniques at room temp. (293 K). The subsequent radial distribution anal. gives C-O and C-C distances that are in good agreement with the known bond lengths of single alc. mols. The shortest averaged C-H and O-H distances are much smaller than those from the gas-phase data. The shortened distance is caused by the deformed electron distribution between H and C atoms. The intermol. hydrogen bonding between hydroxyl groups occurs at a distance of 2.8 Å, and is a common feature for all the samples. The corresponding coordination no. of 1.7 indicates linear chains of about ten mols. The absence of sharp features in the intensity curves suggests that the arrangement of chains is irregular and that the length of the chains varies.
- 56Tomsic, M.; Jamnik, A.; Fritz-Popovski, G.; Glatter, O.; Vlcek, L. Structural Properties of Pure Simple Alcohols from Ethanol, Propanol, Butanol, Pentanol, to Hexanol: Comparing Monte Carlo Simulations with Experimental SAXS Data J. Phys. Chem. B 2007, 111, 1738– 1751There is no corresponding record for this reference.
- 57Vrhovsek, A.; Gereben, O.; Jamnik, A.; Pusztai, L. Hydrogen Bonding and Molecular Aggregates in Liquid Methanol, Ethanol, and 1-Propanol J. Phys. Chem. B 2011, 115, 13473– 1348857https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3MXhtlKksLjI&md5=2d980442f347352f9f32e4ef925c8bd9Hydrogen Bonding and Molecular Aggregates in Liquid Methanol, Ethanol, and 1-PropanolVrhovsek, Aleksander; Gereben, Orsolya; Jamnik, Andrej; Pusztai, LaszloJournal of Physical Chemistry B (2011), 115 (46), 13473-13488CODEN: JPCBFK; ISSN:1520-5207. (American Chemical Society)The authors present a detailed and comprehensive structural study of mol. models of liq. methanol, ethanol, and 1-propanol that originate from reverse Monte Carlo (RMC), mol. dynamics (MD), and united-atom Monte Carlo (UA:MC) simulations. The authors compare several modeling approaches: RMC simulations that employ exptl. neutron and x-ray diffraction data as sole constraints, RMC with diffraction data complemented with partial radial distribution functions (PRDFs) from MD or UA:MC, and conventional MD and UA:MC simulations. The assessment is done in view of the structural parameters of the hydrogen bond and resulting morphol. characteristics of mol. aggregates. To achieve these tasks, a computer program for structural anal. of mol. configurations together with the appropriate aggregate classification scheme was developed. The authors analyzed the morphol. of clusters, their probability, and size distributions. Any cyclic structures that appeared in the configurations were extd. and characterized in the same manner. MD and UA:MC simulations resulted in configurations with bulkier, more threadlike aggregates that were not entirely consistent with the exptl. evidence from diffraction expts. A combination of neutron and x-ray diffraction data with PRDFs from MD simulations, simultaneously applied as constraints in the RMC procedure, proved to be a modeling approach with the most conclusive results.
- 58Dixit, S.; Crain, J.; Poon, W. C. K.; Finney, J. L.; Soper, A. K. Molecular Segregation Observed in a Concentrated Alcohol–Water Solution Nature 2002, 416, 829– 832There is no corresponding record for this reference.
- 59Schröer, W.; Wiegand, S.; Weingärtner, H. The Effect of Short-Range Hydrogen-Bonded Interactions on the Nature of the Critical-Point of Ionic Fluids. 2. Static and Dynamic Light-Scattering on Solutions of Ethylammonium Nitrate in N-Octanol Ber. Bunsen-Ges. Phys. Chem. Chem. Phys. 1993, 97, 975– 98259https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK3sXmsFOgur8%253D&md5=dddd81df1ed9d2a243c6cdc882569ebfThe effect of short-range hydrogen-bonded interactions on the nature of the critical point of ionic fluids. Part II. Static and dynamic light scattering on solutions of ethylammonium nitrate in n-octanolSchroer, W.; Wiegand, S.; Weingaertner, H.Berichte der Bunsen-Gesellschaft (1993), 97 (8), 975-82CODEN: BBPCAX; ISSN:0005-9021.The authors report on static and dynamic light scattering measurements of near-crit. fluctuations of an ionic liq. The system is ethylammonium nitrate (EAN) dissolved in n-octanol which exhibits a liq.-liq. phase transition in the salt-rich regime with an upper crit. point at ∼315 K at a mole fraction of the salt x1 = 0.773. The correlation lengths and scattering intensities scale with temp. corresponding to an Ising-like crit. point. This contrasts with previous observations of mean-field behavior in electrolyte systems with liq.-liq. phase transitions driven by long-range Coulombic interactions. Rationalizations of this discrepancy are offered taking into account a) the competition of intermol. interactions with long-range and short-range character and b) the possibility of crossover from mean-field to Ising behavior as predicted by the Landau-Ginzburg theory.
- 60Atkin, R.; Bobillier, S. M. C.; Warr, G. G. Propylammonium Nitrate as a Solvent for Amphiphile Self-Assembly into Micelles, Lyotropic Liquid Crystals, and Microemulsions J. Phys. Chem. B 2010, 114, 1350– 136060https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1MXhs1alurzL&md5=31e9e6905206e5eb3092ed1b5f5b2391Propylammonium Nitrate as a Solvent for Amphiphile Self-Assembly into Micelles, Lyotropic Liquid Crystals, and MicroemulsionsAtkin, Rob; Bobillier, Sophie M. C.; Warr, Gregory G.Journal of Physical Chemistry B (2010), 114 (3), 1350-1360CODEN: JPCBFK; ISSN:1520-6106. (American Chemical Society)The phase behavior and self-assembled microstructures of a range of oligo(oxyethylene)-n-alkyl ether (CiEj) surfactants was studied in propylammonium nitrate (PAN), a room temp. ionic liq. Micelles and single-phase microemulsions were all found to form at alkyl chain lengths from dodecyl to octadecyl, and lyotropic liq. crystals formed with hexadecyl chains or longer. Small-angle neutron scattering (SANS) shows that self-assembly occurs by solvophobic interactions driving the aggregation of the alkyl chains, but several results indicate that these are weaker in PAN than in water or ethylammonium nitrate, due chiefly to the hydrophobicity of PAN. Longer alkyl chains are needed for lyotropic liq. crystals to form, and higher surfactant concns. are needed to form a single phase microemulsion. Cond. shows these microemulsions to be weakly structured, and relatively insensitive to oil or surfactant mol. structure, unlike water-based systems. However, SANS contrast variation reveals a nanosegregation of oil from the alkyl tails of surfactants within the microemulsion, and may suggest a cosurfactant-like role for the propylammonium cation. Mol. areas within microemulsions and lamellar phases are larger than corresponding water- or ethylammonium nitrate-based systems due to the large mol. vol. of the solvating PANs.
- 61Topolnicki, I. L.; Atkin, R.; FitzGerald, P. A.; Warr, G. G. The Effect of Protic Ionic Liquid and Surfactant Structure on Partitioning of Polyoxyethylene Nonionic Surfactants. ChemPhysChem 2014, 15.There is no corresponding record for this reference.
- 62Kline, S. R. Reduction and Analysis of SANS and USANS Data using Igor Pro J. Appl. Crystallogr. 2006, 39 (6) 895– 90062https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD28Xht1KrurjP&md5=51449389d831a41c38dd4145c342df4dReduction and analysis of SANS and USANS data using IGOR ProKline, Steven R.Journal of Applied Crystallography (2006), 39 (6), 895-900CODEN: JACGAR; ISSN:0021-8898. (International Union of Crystallography)A software package is presented for performing redn. and anal. of small-angle neutron scattering (SANS) and ultra-small-angle neutron scattering (USANS) data. A graphical interface was developed to visualize and quickly reduce raw SANS and USANS data into 1D or 2D formats for interpretation. The resulting reduced data can then be analyzed using model-independent methods or nonlinear fitting to one of a large and growing catalog of included structural models. The different instrumental smearing effects for slit-smeared USANS and pinhole-smeared SANS data are handled automatically during anal. In addn., any no. of SANS and USANS data sets can be analyzed simultaneously. The redn. operations and anal. models are written in a modular format for extensibility, allowing users to contribute code and models for distribution to all users. The software package is based on Igor Pro, providing freely distributable and modifiable code that runs on Macintosh and Windows operating systems.
- 63Teubner, M.; Strey, R. Origin of the Scattering Peak in Microemulsions J. Chem. Phys. 1987, 87, 3195– 320063https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaL2sXlvVChs7w%253D&md5=2e671ec8ee923fd23f5f0275ff6836d3Origin of the scattering peak in microemulsionsTeubner, M.; Strey, R.Journal of Chemical Physics (1987), 87 (5), 3195-200CODEN: JCPSA6; ISSN:0021-9606.From a Landau theory the static scattering intensity distribution I(q) of microemulsions is obtained. It is shown that the scattering intensity relation describes exptl. literature data remarkably well, using only three fit parameters.
- 64Schubert, K. V.; Strey, R. Small-Angle Neutron Scattering from Microemulsions Near the Disorder Line in Water/Formamide--Octane-CiEj Systems J. Chem. Phys. 1991, 95, 8532– 8545There is no corresponding record for this reference.
- 65Vonk, C. G.; Billman, J. F.; Kaler, E. W. Small Angle Scattering of Bicontinuous Structures in Microemulsions J. Chem. Phys. 1988, 88, 3970– 3975There is no corresponding record for this reference.
- 66Schubert, K.-V.; Strey, R.; Kline, S. R.; Kaler, E. W. Small Angle Neutron Scattering Near Lifshitz Lines: Transition from Weakly Structured Mixtures to Microemulsions J. Chem. Phys. 1994, 101, 5343– 535566https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK2cXmvVOjsLo%253D&md5=1041f5facf121cec63c8d98947b2191cSmall angle neutron scattering near Lifshitz lines: transition from weakly structured mixtures to microemulsionsSchubert, K.-V.; Strey, R.; Kline, S. R.; Kaler, E. W.Journal of Chemical Physics (1994), 101 (6), 5343-55CODEN: JCPSA6; ISSN:0021-9606.We have studied the phase behavior, wetting transitions, and small angle neutron scattering (SANS) of water, n-alkane, and n-alkyl polyglycol ether (CiEj) systems in order to locate the transition between weakly structured mixts. and microemulsions, and to provide a measure for the transition. We first detd. the wetting transition by macroscopic measurements and then measured the location of the Lifshitz lines by SANS. Starting with well-structured mixts. (exhibiting nonwetting middle phases and well-expressed scattering peaks, features that qualify them as microemulsions) the wetting transition was induced by increasing the chain length of the alkane or by changing the oil/water vol. ratio, and then the Lifshitz line was crossed. Further, starting with systems past the disorder line (weakly structured mixts. that display wetting middle phases and no scattering peaks), local structure was induced by either increasing the surfactant concn. or decreasing the oil/water vol. ratio or the temp. In each case a Lifshitz line was crossed. Analyzing the scattering expts. quant., allows detn. of the amphiphilicity factor, which is a measure of the strength of the surfactant. The results suggest there is a sequence of roughly parallel surfaces within the three-dimensional compn.-temp. space. As the amphiphilicity factor increases, first a disorder surface is encountered, then a Lifshitz surface, and finally a wetting transition surface. How and to what extent these surfaces move in the one-phase region toward smaller surfactant concns., and intersect there with the body of heterogeneous phases, depends on a no. of factors that are discussed in some detail.
- 67Barnes, I. S.; Corti, M.; Degiorgio, V.; Zemb, T. X-ray and Neutron-Scattering Measurements on Concentrated Non-Ionic Amphiphile Solutions Prog. Colloid Polym. Sci. 1993, 93, 205– 20667https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK2cXmvVOjs7g%253D&md5=c7da655991ee31f843df0648224adefcX-ray and neutron-scattering measurements on concentrated non-ionic amphiphile solutionsBarnes, I. S.; Corti, M.; Degiorgio, V.; Zemb, T.Progress in Colloid & Polymer Science (1993), 93 (TRENDS IN COLLOID AN), 205-6CODEN: PCPSD7; ISSN:0340-255X.The structure of aq. solns. of the nonionic amphiphile C12E8 is studied by small-angle x-ray and neutron scattering along the isothermal path across the single-phase region from 0% to 100% amphiphile vol. fraction. Soln. structuring is present even in the pure amphiphile, as shown by x-ray data.
- 68Laughlin, R. The Aqueous Phase Behavior of Surfactants; Academic Press: San Diego, 1994.There is no corresponding record for this reference.
- 69Diat, O.; K, M. L.; Touraud, D.; Deme, B.; Grillo, I.; Kunz, W.; Zemb, T. Octanol-Rich and Water-Rich Domains in Dynamic Equilibrium in the Pre-Ouzo Region of Ternary Systems Containing a Hydrotrope J. Appl. Crystallogr. 2013, 46, 1665– 166969https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3sXhsl2ntL3J&md5=7e7cb9f121dada6fd0bdec4b4747e090Octanol-rich and water-rich domains in dynamic equilibrium in the pre-ouzo region of ternary systems containing a hydrotropeDiat, Olivier; Klossek, Michael L.; Touraud, Didier; Deme, Bruno; Grillo, Isabelle; Kunz, Werner; Zemb, ThomasJournal of Applied Crystallography (2013), 46 (6), 1665-1669CODEN: JACGAR; ISSN:0021-8898. (International Union of Crystallography)Ternary mixts. of medium-chain fatty alcs., water and a hydrotrope (such as ethanol), near the immiscibility gap, make stable single phases at const. temp. Interestingly, in this 'pre-ouzo region' these single phases consist of two distinct nanoscopic pseudo-phases, one octanol-rich and one water-rich. This domain of compn., which is known to produce strong light scattering and to sep. under ultracentrifugation into two phases, has been studied using contrast variation in small-angle neutron scattering (SANS) combined with small- and wide-angle X-ray scattering (SWAXS). The existence of fatty alc.-rich domains of well defined size of the order of 2 nm radius is proven. The scattering can be approximated by an Ornstein-Zernike function, which is close to the general expression of Choi, Chen, Sottmann & Strey [Physica B, (1998), 241-243, 976-978] with vanishing quadratic Porod term. Exploitation of the relative intensities at the vanishing scattering angle in SANS demonstrates that the distribution coeff. of ethanol between the octanol-rich and the water-rich domains is close to one. WAXS of the two coexisting pseudo-phases is compared with the corresponding binary water-ethanol and octanol-ethanol samples.
- 70Tchakalova, V.; Zemb, T.; Benczédi, D. Evaporation Triggered Self-Assembly in Aqueous Fragrance–Ethanol Mixtures and Its Impact on Fragrance Performance. Colloids Surf., A 2014, in pressThere is no corresponding record for this reference.
- 71Clint, J. H. Micellization of Mixed Nonionic Surface Active Agents J. Chem. Soc., Faraday Trans. 1 1975, 71 (0) 1327– 1334There is no corresponding record for this reference.
- 72Kahlweit, M.; Strey, R.; Busse, G. Weakly to Strongly Structured Mixtures Phys. Rev. E 1993, 47, 4197– 4209There is no corresponding record for this reference.
- 73Greaves, T. L.; Drummond, C. J. Ionic Liquids as Amphiphile Self-Assembly Media Chem. Soc. Rev. 2008, 37, 1709– 172673https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1cXoslOgtbk%253D&md5=512c82d5493b639e25979d578faed5c9Ionic liquids as amphiphile self-assembly mediaGreaves, Tamar L.; Drummond, Calum J.Chemical Society Reviews (2008), 37 (8), 1709-1726CODEN: CSRVBR; ISSN:0306-0012. (Royal Society of Chemistry)A review. In recent years, the no. of nonaq. solvents which mediate hydrocarbon-solvent interactions and promote the self-assembly of amphiphiles was markedly increased by the reporting of over 30 ionic liqs. which possess this previously unusual solvent characteristic. This new situation allows a different exploration of the mol. solvophobic effect and tests the current understanding of amphiphile self-assembly. Both protic and aprotic ionic liqs. support amphiphile self-assembly, indicating that it is not required for the solvents to be able to form a hydrogen bonded network. Here, the use of ionic liqs. as amphiphile self-assembly media is reviewed, including micelle and liq. cryst. mesophase formation, their use as a solvent phase in microemulsions and emulsions, and the emerging field of nanostructured inorg. materials synthesis. Surfactants, lipids and block co-polymers are the focus amphiphile classes in this crit. review (174 refs.).
- 74Beesley, A. H.; Evans, D. F.; Laughlin, R. G. Evidence For The Essential Role Of Hydrogen-Bonding In Promoting Amphiphilic Self-Assembly - Measurements in 3-Methylsydnone J. Phys. Chem. 1988, 92, 791– 793There is no corresponding record for this reference.
- 75Ray, A. Micelle Formation in Pure Ethylene Glycol J. Am. Chem. Soc. 1969, 91 (23) 6511– 6512There is no corresponding record for this reference.
- 76Ray, A. Solvophobic Interactions and Micelle Formation in Structure Forming Nonaqueous Solvents Nature 1971, 231, 313– 31576https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaE3MXksFSksrk%253D&md5=5e9f7c861f0b8352050969c6f20e9356Solvophobic interactions and micelle formation in structure forming nonaqueous solventsRay, AshokaNature (London, United Kingdom) (1971), 231 (5301), 313-15CODEN: NATUAS; ISSN:0028-0836.Solvophobic interactions occur in glycerol, HCONH2, H2N(CH2)2OH, HO(CH2)3OH, HO(CH2)4OH, HCO2H, H2N(CH2)2NH2, HS(CH2)2OH, HOCH2CH2CH(OH)Me, HOCH2CH(OH)Me, and H2NCH2CH(OH)Me (class I). Such interactions do not exist in HCONHMe, HCONMe2, and CH2I2 (class II), or in MeOH, EtOH, and PhMe (class III). A nonionic detergent [e.g., (p-tert-nonylphenoxy)polyethylene glycol] would aggregate into micelles only in class I solvents. The crit. micelle concns. of the detergent, and the free energies of micelle formation were calcd.