Zeta Potentials of Cotton Membranes in Acetonitrile SolutionsClick to copy article linkArticle link copied!
- Yuki Uematsu*Yuki Uematsu*Email: [email protected]Department of Physics and Information Technology, Kyushu Institute of Technolohy, Iizuka 820-8502, JapanPRESTO, Japan Science and Technology Agency, 4-1-8 Honcho, Kawaguchi, Saitama 332-0012, JapanMore by Yuki Uematsu
- Suguru IwaiSuguru IwaiDepartment of Chemical Science and Engineering, School of Materials and Chemical Technology, Tokyo Institute of Technology, 4259 Nagatsuta-cho, Midori-ku, Yokohama 226-8502, JapanMore by Suguru Iwai
- Mariko KonishiMariko KonishiDepartment of Chemical Science and Engineering, School of Materials and Chemical Technology, Tokyo Institute of Technology, 4259 Nagatsuta-cho, Midori-ku, Yokohama 226-8502, JapanMore by Mariko Konishi
- Shinsuke InagiShinsuke InagiDepartment of Chemical Science and Engineering, School of Materials and Chemical Technology, Tokyo Institute of Technology, 4259 Nagatsuta-cho, Midori-ku, Yokohama 226-8502, JapanMore by Shinsuke Inagi
Abstract
Solid surfaces in contact with nonaqueous solvents play a key role in electrochemistry, analytical chemistry, and industrial chemistry. In this work, the zeta potentials of cotton membranes in acetonitrile solutions were determined by streaming potential and bulk conductivity measurements. By applying the Gouy–Chapman theory and the Langmuir adsorption isotherm of ions to the experimental data, the mechanism of the electrification at the cotton/acetonitrile interface is revealed for the first time to be solely due to ion adsorption on the surface, rather than proton dissociation at the interface. Different salts were found to produce opposite signs of the zeta potentials. This behavior can be attributed to ion solvation effects and the strong ordering of acetonitrile molecules at the interface. Furthermore, a trend of the electroviscous effect was observed, in agreement with the standard electrokinetic theory. These findings demonstrate that electrokinetics in acetonitrile, a polar aprotic solvent, can be treated in the same manner as in water.
This publication is licensed under
License Summary*
You are free to share(copy and redistribute) this article in any medium or format within the parameters below:
Creative Commons (CC): This is a Creative Commons license.
Attribution (BY): Credit must be given to the creator.
Non-Commercial (NC): Only non-commercial uses of the work are permitted.
No Derivatives (ND): Derivative works may be created for non-commercial purposes, but sharing is prohibited.
*Disclaimer
This summary highlights only some of the key features and terms of the actual license. It is not a license and has no legal value. Carefully review the actual license before using these materials.
License Summary*
You are free to share(copy and redistribute) this article in any medium or format within the parameters below:
Creative Commons (CC): This is a Creative Commons license.
Attribution (BY): Credit must be given to the creator.
Non-Commercial (NC): Only non-commercial uses of the work are permitted.
No Derivatives (ND): Derivative works may be created for non-commercial purposes, but sharing is prohibited.
*Disclaimer
This summary highlights only some of the key features and terms of the actual license. It is not a license and has no legal value. Carefully review the actual license before using these materials.
License Summary*
You are free to share(copy and redistribute) this article in any medium or format within the parameters below:
Creative Commons (CC): This is a Creative Commons license.
Attribution (BY): Credit must be given to the creator.
Non-Commercial (NC): Only non-commercial uses of the work are permitted.
No Derivatives (ND): Derivative works may be created for non-commercial purposes, but sharing is prohibited.
*Disclaimer
This summary highlights only some of the key features and terms of the actual license. It is not a license and has no legal value. Carefully review the actual license before using these materials.
License Summary*
You are free to share(copy and redistribute) this article in any medium or format within the parameters below:
Creative Commons (CC): This is a Creative Commons license.
Attribution (BY): Credit must be given to the creator.
Non-Commercial (NC): Only non-commercial uses of the work are permitted.
No Derivatives (ND): Derivative works may be created for non-commercial purposes, but sharing is prohibited.
*Disclaimer
This summary highlights only some of the key features and terms of the actual license. It is not a license and has no legal value. Carefully review the actual license before using these materials.
License Summary*
You are free to share(copy and redistribute) this article in any medium or format within the parameters below:
Creative Commons (CC): This is a Creative Commons license.
Attribution (BY): Credit must be given to the creator.
Non-Commercial (NC): Only non-commercial uses of the work are permitted.
No Derivatives (ND): Derivative works may be created for non-commercial purposes, but sharing is prohibited.
*Disclaimer
This summary highlights only some of the key features and terms of the actual license. It is not a license and has no legal value. Carefully review the actual license before using these materials.
Introduction
Experimental Section
Materials
Methods
Results with Discussion
Zeta Potential
Electroviscous Effect
Conclusions
Acknowledgments
This work was supported by JST PRESTO Grant No. JPMJPR21O2 and JST FOREST Grant No. JPMJFR211G.
References
This article references 62 other publications.
- 1Russel, W. B.; Saville, D. A.; Schowalter, W. R. Colloidal Dispersions; Cambridge University Press, 1989.Google ScholarThere is no corresponding record for this reference.
- 2Uematsu, Y. Electrification of water interface. J. Phys.: Condens. Matter 2021, 33, 423001, DOI: 10.1088/1361-648X/ac15d5Google Scholar2https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB3MXhvF2hs77O&md5=869f91e31de4b7966eb00e86a0bcf12aElectrification of water interfaceUematsu, YukiJournal of Physics: Condensed Matter (2021), 33 (42), 423001CODEN: JCOMEL; ISSN:0953-8984. (IOP Publishing Ltd.)A review. The surface charge of a water interface dets. many fundamental processes in phys. chem. and interface science, and it has been intensively studied for over a hundred years. We summarize exptl. methods to characterize the surface charge densities developed so far: electrokinetics, double-layer force measurements, potentiometric titrn., surface-sensitive nonlinear spectroscopy, and surface-sensitive mass spectrometry. Then, we elucidate phys. ion adsorption and chem. electrification as examples of electrification mechanisms. In the end, novel effects on surface electrification are discussed in detail. We believe that this clear overview of state of the art in a charged water interface will surely help the fundamental progress of physics and chem. at interfaces in the future.
- 3Siria, A.; Poncharal, P.; Biance, A.-L.; Fulcrand, R.; Blase, X.; Purcell, S. T.; Bocquet, L. Giant osmotic energy conversion measured in a single transmembrane boron nitride nanotube. Nature 2013, 494, 455– 458, DOI: 10.1038/nature11876Google Scholar3https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3sXjtFSrt7Y%253D&md5=6b675bc4d36dfb57cb309bad50613859Giant osmotic energy conversion measured in a single transmembrane boron nitride nanotubeSiria, Alessandro; Poncharal, Philippe; Biance, Anne-Laure; Fulcrand, Remy; Blase, Xavier; Purcell, Stephen T.; Bocquet, LydericNature (London, United Kingdom) (2013), 494 (7438), 455-458CODEN: NATUAS; ISSN:0028-0836. (Nature Publishing Group)New models of fluid transport are expected to emerge from the confinement of liqs. at the nanoscale, with potential applications in ultrafiltration, desalination and energy conversion. Nevertheless, advancing our fundamental understanding of fluid transport on the smallest scales requires mass and ion dynamics to be ultimately characterized across an individual channel to avoid averaging over many pores. A major challenge for nanofluidics thus lies in building distinct and well-controlled nanochannels, amenable to the systematic exploration of their properties. Here we describe the fabrication and use of a hierarchical nanofluidic device made of a boron nitride nanotube that pierces an ultrathin membrane and connects two fluid reservoirs. Such a transmembrane geometry allows the detailed study of fluidic transport through a single nanotube under diverse forces, including elec. fields, pressure drops and chem. gradients. Using this device, we discover very large, osmotically induced elec. currents generated by salinity gradients, exceeding by two orders of magnitude their pressure-driven counterpart. We show that this result originates in the anomalously high surface charge carried by the nanotube's internal surface in water at large pH, which we independently quantify in conductance measurements. The nano-assembly route using nanostructures as building blocks opens the way to studying fluid, ionic and mol. transport on the nanoscale, and may lead to biomimetic functionalities. Our results furthermore suggest that boron nitride nanotubes could be used as membranes for osmotic power harvesting under salinity gradients.
- 4Gortner, R. A. Electrokinetics XXIII. Electrokinetics as a tool for the study of the molecular structure of organic compounds. Trans. Faraday Soc. 1940, 35, 63, DOI: 10.1039/TF9403500063Google ScholarThere is no corresponding record for this reference.
- 5Westermann-Clark, G.; Christoforou, C. Note on nonaqueous electrokinetic transport in charged porous media. J. Membr. Sci. 1984, 20, 325– 338, DOI: 10.1016/S0376-7388(00)82009-8Google Scholar5https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaL2cXlvFyktrw%253D&md5=8117a095c3950f93c6a60a515684abfeNote on nonaqueous electrokinetic transport in charged porous mediaWestermann-Clark, G. B.; Christoforou, C. C.Journal of Membrane Science (1984), 20 (3), 325-39CODEN: JMESDO; ISSN:0376-7388.A model for electrokinetic transport in charged capillaries is compared with expts. by using nonaq. LiCl solns. The electrokinetic parameters considered are the pore fluid cond. and the concn. potential. Methanol/water mixts. were the solvents and track-etched mica membranes with a well-characterized pore structure were the porous medium. The electrolyte concns. used were such that the Debye lengths of solns. in pores ranged from much smaller to much larger than the radii of the pores. The space-charge model is capable of describing qual. the trend of the electrokinetic data, but as expected, at higher concns. the model fails, probably because the assumption that ion-ion interactions are negligible no longer holds. The pore fluid cond. depends strongly on the dielec. const. of the solvent; the abs. value of the pore wall charge tends to decrease with the lowering of the solvent dielec. const., and the wall charge tends to increase with the concn. of the chloride ion.
- 6Lucas, R. A.; Lin, C.-Y.; Siwy, Z. S. Electrokinetic Phenomena in Organic Solvents. J. Phys. Chem. B 2019, 123, 6123– 6131, DOI: 10.1021/acs.jpcb.9b04969Google Scholar6https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1MXhtF2ks7fM&md5=e06b885a013b890be5cabf431cf8877aElectrokinetic Phenomena in Organic SolventsLucas, Rachel A.; Lin, Chih-Yuan; Siwy, Zuzanna S.Journal of Physical Chemistry B (2019), 123 (28), 6123-6131CODEN: JPCBFK; ISSN:1520-5207. (American Chemical Society)Solid/liq. interfaces play a key role in sepn. processes, energy storage devices, and transport in nanoscale systems. Nanopores and mesopores with well-defined geometry and chem. characteristics have been a valuable tool to unravel electrochem. properties of interfaces, but the majority of studies have been focused on aq. solns. Here, we present expts. and numerical modeling aimed at characterizing effective surface charge of polymer pores in mixts. of water and alcs. as well as in propylene carbonate and acetone. The charge properties of pore walls are probed through anal. of current-voltage curves recorded in the presence of salt concn. gradients. The presence and direction of electro-osmotic flow lead to asym. current-voltage curves, with rectification characteristics detd. by the polarity of surface charge. The results suggest that the effective surface charge of the pore walls depends not only on the type of solvent but also on the concn. of the electrolyte and voltage. We identified conditions at which polymer pores that are neg. charged in aq. solns. become pos. charged in propylene carbonate and acetone. The findings are of importance for nonaq. sepns., fundamental knowledge on solid/liq. interfaces in org. media, and prepn. of porous devices with tunable surface charge characteristics.
- 7Russell, W. S.; Siwy, Z. Enhanced electro-osmosis in propylene carbonate salt solutions. J. Chem. Phys. 2021, 154, 134707, DOI: 10.1063/5.0044402Google Scholar7https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB3MXotlGnsL0%253D&md5=5f659f27bbe30ab75e421d54494131a6Enhanced electro-osmosis in propylene carbonate salt solutionsRussell, Wilfred Shelby; Siwy, ZuzannaJournal of Chemical Physics (2021), 154 (13), 134707CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)Properties of solid-liq. interfaces and surface charge characteristics mediate ionic and mol. transport through porous systems, affecting many processes such as sepns. Herein, the authors report expts. designed to probe the electrochem. properties of solid-liq. interfaces using a model system of a single polyethylene terephthalate (PET) pore in contact with aq. and propylene carbonate solns. of LiClO4. First, the existence and polarity of surface charges were inferred from current-voltage curves recorded when a pore was placed in contact with a LiClO4 concn. gradient. Second, the electroosmotic transport of uncharged polystyrene particles through the PET pore provided information on the polarity and the magnitude of the pore walls' zeta potential. The authors' expts. show that the PET pores become effectively pos. charged when in contact with LiClO4 solns. in propylene carbonate, even though in aq. LiClO4, the same pores are neg. charged. Addnl., the electroosmotic velocity of the particles revealed a significantly higher magnitude of the pos. zeta potential of the pores in propylene carbonate compared to the magnitude of the neg. zeta potential in H2O. The presented methods of probing the properties of solid-liq. interfaces are expected to be applicable to a wide variety of solid and liq. systems. (c) 2021 American Institute of Physics.
- 8Suzuki, Y.; Mizuhata, M. Predictive Zeta Potential Measurement Method Applicable to Nonaqueous Solvents in High-concentration Dispersion Systems for the System of LiClO4–Propylene Carbonate Solution and LiCoO2 Powder Sheet. Electrochemistry 2022, 90, 103001– 103001, DOI: 10.5796/electrochemistry.22-66050Google ScholarThere is no corresponding record for this reference.
- 9Schwer, C.; Kenndler, E. Electrophoresis in fused-silica capillaries: the influence of organic solvents on the electroosmotic velocity and the zeta potential. Anal. Chem. 1991, 63, 1801– 1807, DOI: 10.1021/ac00017a026Google Scholar9https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK3MXltVSitro%253D&md5=6b5df657d236b8d895b4a924bba8eb6cElectrophoresis in fused-silica capillaries: the influence of organic solvents on the electroosmotic velocity and the ζ potentialSchwer, Christine; Kenndler, ErnstAnalytical Chemistry (1991), 63 (17), 1801-7CODEN: ANCHAM; ISSN:0003-2700.The influence of pH and solvent compn. on the electroosmotic flow in fused-silica capillaries was studied. Binary mixts. of H2O and a protic (MeOH, EtOH, iso-PrOH) or an aprotic dipolar solvent (MeCN, Me2CO, DMSO) were used for the buffer electrolyte. The dependence of the electroosmotic flow on the pH was detd. in the pH range from 3 to 11 for pure aq. soln. and for solns. contg. 50 vol.% org. solvent. Adding org. solvents to the electrolyte, the inflection points of the resulting curves, corresponding to the pK values of the surface silanol groups, are shifted to higher values. At high pH, increasing the fraction of org. cosolvent generally decreases the electroosmotic velocity. Zeta potentials were calcd. from the Smoluchowski equation by using electroosmotic velocity data and values for the ratio of the viscosity coeff. and dielec. const. for binary mixts., taken from the authors own measurements and from the literature. With the exception of Me2CO-H2O mixts., the ζ potentials obtained show a similar trend to that of the electroosmotic velocity; they decrease with increasing content of org. solvent. This trend is explained by changes of the dielec. properties of the elec. double layer and of the charge generation on the fused-silica surface.
- 10Wright, P. B.; Lister, A. S.; Dorsey, J. G. Behavior and Use of Nonaqueous Media without Supporting Electrolyte in Capillary Electrophoresis and Capillary Electrochromatography. Anal. Chem. 1997, 69, 3251– 3259, DOI: 10.1021/ac9613186Google Scholar10https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK2sXkslKmtLc%253D&md5=12548b0d583d659dcf0685819dfc1d56Behavior and Use of Nonaqueous Media without Supporting Electrolyte in Capillary Electrophoresis and Capillary ElectrochromatographyWright, Paul B.; Lister, Ashley S.; Dorsey, John G.Analytical Chemistry (1997), 69 (16), 3251-3259CODEN: ANCHAM; ISSN:0003-2700. (American Chemical Society)Five nonaq. solvents (acetonitrile, methanol, DMF, DMSO, formamide) and deionized water were studied for their ability to support electroosmotic flow (EOF) without electrolytic additives. In general, flow is equal to or greater than flow with typical CE buffer systems. The magnitude of EOF was detd. for each solvent by open tubular capillary electrophoresis (CE) and related to viscosity (η), dielec. const. (ε), and the ratio of dielec. const. to viscosity (ε/η). Zeta potentials (ζ) were derived indirectly from flow data and tabulated. Comparisons of flow behavior and ζ were made between pure solvents and conventional CE buffers, and questions of equilibration and reproducibility were addressed. Similar expts. were performed using hydroorg. mobile phases (ACN/water, MeOH/water) across the complete compositional range (100% water-100% org.), with flow characteristics and ζ reported for each mobile phase system. Packed capillary columns (5-μm ODS) were evaluated for flow and retention stability under capillary electrochromatog. (CEC) conditions. A sepn. of 11 polycyclic arom. hydrocarbons was performed in under 13 min by CEC with an ACN/water mobile phase. Reduced plate heights (h) were calcd. between 2.5 and 3.0 for solutes with capacity factors (k') up to 4.5 for the most retained solute.
- 11Tjo̷rnelund, J.; Hansen, S. H. The effect of water on separations in non-aqueous capillary electrophoresis systems. Chromatographia 1997, 44, 5– 9, DOI: 10.1007/BF02466508Google ScholarThere is no corresponding record for this reference.
- 12Valkó, I. E.; Sirén, H.; Riekkola, M.-L. Characteristics of electroosmotic flow in capillary electrophoresis in water and in organic solvents without added ionic species. J. Microcolumn Sep. 1999, 11, 199– 208, DOI: 10.1002/(SICI)1520-667X(1999)11:3<199::AID-MCS5>3.0.CO;2-DGoogle Scholar12https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK1MXhtlGntLo%253D&md5=5565670086434868d01b1c6931468c0fCharacteristics of electroosmotic flow in capillary electrophoresis in water and in organic solvents without added ionic speciesValko, Istvan E.; Siren, Heli; Riekkola, Marja-LiisaJournal of Microcolumn Separations (1999), 11 (3), 199-208CODEN: JMSEEJ; ISSN:1040-7685. (John Wiley & Sons, Inc.)Although proper buffering is essential for repeatable sepns. in capillary electrophoresis, electroosmotic flow can be developed in many solvents without the addn. of buffer components. In this study we detected electroosmotic flow in fused silica capillary in the following solvents without the addn. of ionic species: water, deuterium oxide, acetonitrile, acetone, 2-butanone, formamide, N-methylformamide, N,N-dimethylformamide, methanol, ethanol, 1-propanol, DMSO, Et acetate, THF, and morpholine. The electroosmotic flow was in the range of 3.4 X 10-9-1.8 X 10-7 m2/V s. About 100 times weaker electroosmosis was detected in glacial acetic acid. We also detected electroosmotic flow in N-methylacetamide, which is solid at room temp. (melting range 28-30°C). No electroosmotic flow was found in inert solvents such as n-hexane and chloroform. The zeta potential of the fused silica capillary was calcd. for the solvents where electroosmotic flow was found. The zeta potential of the capillary wall was also studied in water-methanol mixts., where it has a max. at about 40% methanol concn.
- 13Riekkola, M.-L.; Jussila, M.; Porras, S. P.; Valkó, I. E. Non-aqueous capillary electrophoresis. J. Chromatogr. A 2000, 892, 155– 170, DOI: 10.1016/S0021-9673(00)00108-4Google ScholarThere is no corresponding record for this reference.
- 14Geiser, L.; Mirgaldi, M.; Veuthey, J.-L. Determination of electroosmotic flow in nonaqueous capillary electrophoresis. J. Chromatogr. A 2005, 1068, 75– 81, DOI: 10.1016/j.chroma.2005.02.001Google ScholarThere is no corresponding record for this reference.
- 15Kitahara, A. Zeta potential in nonaqueous media and its effect on dispersion stability. Prog. Org. Coat. 1973, 2, 81– 98, DOI: 10.1016/0300-9440(73)80001-3Google Scholar15https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaE2cXktFKqtrs%253D&md5=0e7f651e29ac8a76fc64d605281265a4Zeta potential in nonaqueous media and its effect on dispersion stabilityKitahara, AyaoProgress in Organic Coatings (1973), 2 (2), 81-98CODEN: POGCAT; ISSN:0300-9440.A review is given with 38 refs. The properties and stability of org. coatings in relation to zeta potential are also discussed.
- 16Fowkes, F. Dispersions of ceramic powders in organic media. Adv. Ceram.: Ceram. Powder Sci. 1987, 21, 411– 421Google ScholarThere is no corresponding record for this reference.
- 17Van Der Hoeven, P.; Lyklema, J. Electrostatic stabilization in non-aqueous media. Adv. Colloid Interface Sci. 1992, 42, 205– 277, DOI: 10.1016/0001-8686(92)80024-RGoogle Scholar17https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK38XmsFyksbo%253D&md5=5c29bb91e393773edc986c1ba9396877Electrostatic stabilization in non-aqueous mediaVan der Hoeven, Ph. C.; Lyklema, J.Advances in Colloid and Interface Science (1992), 42 (), 205-77CODEN: ACISB9; ISSN:0001-8686.A review on electrostatic stabilization of non-aq. dispersions with particular emphasis on the roles of dielec. const. and ionic strength. The liq. dielec. const. controls the degree of ionization of the stabilizing electrolyte and 3 regimes (semipolar, low-polar, apolar) can be distinguished. When attraction between suspended salt particles in nonionic liqs. is weak, electrostatic stabilization is achieved easily. 107 Refs.
- 18Siffert, B.; Jada, A.; Eleli-Letsango, J. Stability Calculations of TiO2 Nonaqueous Suspensions: Thickness of the Electrical Double Layer. J. Colloid Interface Sci. 1994, 167, 281– 286, DOI: 10.1006/jcis.1994.1362Google ScholarThere is no corresponding record for this reference.
- 19Hughes, D. F. K.; Robb, I. D.; Dowding, P. J. Stability of Copper Phthalocyanine Dispersions in Organic Media. Langmuir 1999, 15, 5227– 5231, DOI: 10.1021/la981389vGoogle Scholar19https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK1MXkt1aqsrw%253D&md5=647f09165ba8da13704defab831405adStability of Copper Phthalocyanine Dispersions in Organic MediaHughes, D. F. K.; Robb, Ian D.; Dowding, Peter J.Langmuir (1999), 15 (16), 5227-5231CODEN: LANGD5; ISSN:0743-7463. (American Chemical Society)It was found that quite small surface potentials (a few millivolts) were sufficient to stabilize the dil. dispersions of the particles in solvents of low dielec. const. (possibly because this would be a const. charge system) though insufficient to stabilize particles in solvents with dielec. consts. above about 20. Addn. of org. acids enhanced the particles' stability, probably by a proton-exchange mechanism, and addn. of an org.-sol. electrolyte, tetrabutylammonium bromide, caused all systems to flocculate. The presence of a polymer at concns. above about 1% caused flocculation to occur, probably by a depletion mechanism. The attractive depletion energy increased with polymer concn., resulting in the sedimentation vol. increasing with polymer concn.
- 20Kosmulski, M. Zeta potentials in nonaqueous media: how to measure and control them. Colloids Surf., A 1999, 159, 277– 281, DOI: 10.1016/S0927-7757(99)00273-3Google Scholar20https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK1MXnsFertbY%253D&md5=30b927e714a0130eb179580f2eb4616bZeta potentials in nonaqueous media. How to measure and control themKosmulski, M.Colloids and Surfaces, A: Physicochemical and Engineering Aspects (1999), 159 (2-3), 277-281CODEN: CPEAEH; ISSN:0927-7757. (Elsevier Science B.V.)A review, with 20 refs. The effect of traces of water in nonaq. solvents on the ζ-potential is overrated, and the role of impurities in presumably pure org. solvents is often overlooked. Oxides can serve as potentiometric sensors to det. the concn. and nature of impurities. Even in very inert solvents the counter-charge in soln. must exist. Small ions are stabilized by homoconjugation and by amphiphilic nonionic compds. The electrokinetic phenomena of the second kind are more common in nonaq. solvents and they lead to very high abs. values of mobility at high field strengths.
- 21Kosmulski, M.; Eriksson, P.; Brancewicz, C.; Rosenholm, J. B. Zeta potentials of monodispersed, spherical silica particles in mixed solvents as a function of cesium chloride concentration. Colloids Surf., A 2000, 162, 37– 48, DOI: 10.1016/S0927-7757(99)00027-8Google Scholar21https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3cXhtFeltL4%253D&md5=de877254cdeaede02f8f4c7684abf0d1Zeta potentials of monodispersed, spherical silica particles in mixed solvents as a function of cesium chloride concentrationKosmulski, Marek; Eriksson, Patrik; Brancewicz, Chris; Rosenholm, Jarl B.Colloids and Surfaces, A: Physicochemical and Engineering Aspects (2000), 162 (1-3), 37-48CODEN: CPEAEH; ISSN:0927-7757. (Elsevier Science B.V.)Zeta potentials of silica are neg. in most mixed solvents (99% org. co-solvent, 1% water), but in t-butanol and pyridine they are pos. When CsCl is added to the suspension as a probe substance, two types of behavior are obsd. In solvents of ε >25, the zeta potential asymptotically approaches zero, while in solvents of lower dielec. const., the sign of zeta potential is reversed to pos. at a sufficiently high CsCl concn., crev. To evaluate the nature of interaction leading to different solvent responses, a range of frequently used solvent scales relating to specific interactions between the components of the system were compared. A moderate correlation was found between crev for silica and one solvent property alone, and those previously found for anatase, resp. Using linear combinations of 2 solvent scales improved the correlation, but it was still not satisfactory. However, considering sep. the contributions from both solvent components, a nearly statistically significant correlation was found. The results are evaluated considering the particular behavior of each solvent group, and the specific influence of the material properties of silica and of possible impurities present in the system.
- 22Zhukov, A. Integrated investigations of the electrosurface properties of nonaqueous disperse and capillary systems. Adv. Colloid Interface Sci. 2007, 134–135, 330– 345, DOI: 10.1016/j.cis.2007.04.014Google ScholarThere is no corresponding record for this reference.
- 23Hamieh, T.; Toufaily, J.; Alloul, H. Physicochemical Properties of the Dispersion of Titanium Dioxide in Organic Media by Using Zetametry Technique. J. Dispers. Sci. Technol. 2008, 29, 1181– 1188, DOI: 10.1080/01932690701856626Google ScholarThere is no corresponding record for this reference.
- 24Kosmulski, M.; Prochniak, P.; Rosenholm, J. B. Control of the Zeta Potential in Semiconcentrated Dispersions of Titania in Polar Organic Solvents. J. Phys. Chem. C 2009, 113, 12806– 12810, DOI: 10.1021/jp903845eGoogle Scholar24https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1MXnsVWgtbg%253D&md5=796d406d0c2347c33f35537538c4ce80Control of the Zeta Potential in Semiconcentrated Dispersions of Titania in Polar Organic SolventsKosmulski, Marek; Prochniak, Piotr; Rosenholm, Jarl B.Journal of Physical Chemistry C (2009), 113 (29), 12806-12810CODEN: JPCCCK; ISSN:1932-7447. (American Chemical Society)The electrokinetic potential of com. titania powder (chiefly anatase) in its 1-10% dispersions in water, lower aliph. alcs., and DMSO is pos., and rather insensitive to the nature of the solvent. The sign of the electrokinetic potential can be reversed to neg. by addn. of phosphoric acid, or of combination of phosphoric acid with triethylamine or with alkali hydroxide. The crit. concn. of surface-active compds., which induces a sign reversal, depends on the nature of the solvent, and it is higher in org. solvents than in water. The crit. surface concn. of surface-active compds. for given org. solvent is rather insensitive to the solid-to-liq. ratio, and typically it is in the range of a few micromoles per square meter.
- 25Cihlar, J.; Drdlik, D.; Cihlarova, Z.; Hadraba, H. Effect of acids and bases on electrophoretic deposition of alumina and zirconia particles in 2-propanol. J. Eur. Ceram. Soc. 2013, 33, 1885– 1892, DOI: 10.1016/j.jeurceramsoc.2013.02.017Google ScholarThere is no corresponding record for this reference.
- 26Lyklema, J. Principles of interactions in non-aqueous electrolyte solutions. Curr. Opin. Colloid Interface Sci. 2013, 18, 116– 128, DOI: 10.1016/j.cocis.2013.02.002Google Scholar26https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3sXivFylsrY%253D&md5=296ea4b32938cacda3d79800931810b7Principles of interactions in non-aqueous electrolyte solutionsLyklema, JohannesCurrent Opinion in Colloid & Interface Science (2013), 18 (2), 116-128CODEN: COCSFL; ISSN:1359-0294. (Elsevier Ltd.)In this paper a review is presented on the mol. interactions in non-aq. media of low dielec. permittivity. Qual. and quant. distinctions with aq. solns. are emphasized. The reviewed themes include dispersion forces, dissocn. and assocn. equil., discrimination between electrostatic and non-electrostatic interactions, ionic specificity, cond., electrokinetics and colloid interaction. Distinctions between so-called primitive and non-primitive interpretations and between individual and collective behavior are discussed; in these respects the colloid stability phenomena behave differently from the corresponding ones in aq. solvents.
- 27Mazzini, V.; Craig, V. S. Specific-ion effects in non-aqueous systems. Curr. Opin. Colloid Interface Sci. 2016, 23, 82– 93, DOI: 10.1016/j.cocis.2016.06.009Google Scholar27https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC28XhtFCitbfK&md5=2aacad84502b94640a7f3075129d85ecSpecific-ion effects in non-aqueous systemsMazzini, Virginia; Craig, Vincent S. J.Current Opinion in Colloid & Interface Science (2016), 23 (), 82-93CODEN: COCSFL; ISSN:1359-0294. (Elsevier Ltd.)A review. It is widely acknowledged that specific-ion effects are ubiquitous in aq. systems and undoubtedly are essential to the fundamental processes of life, although a deep fundamental understanding of specific-ion effects remains an important challenge. Specific-ion effects in non-aq. solvents are known but have attracted far less attention, yet knowledge of specific-ion effects in non-aq. systems is likely to provide important information for guiding, evaluating and testing our theories of specific-ion effects. Here, the literature on specific-ion effects in non-aq. solvents is surveyed with a view to detg. if the Hofmeister series or lyotropic series so universally obsd. in aq. systems is widely evident in non-aq. systems. Particular attention has been applied to expts. on non-aq. systems that are known to exhibit Hofmeister series in aq. systems with the aim of detg. if a consistent ion ordering in the strength of specific-ion effects is obsd. in other solvents. We find that specific-ion effects are ubiquitous in non-aq. solvents, that both Hofmeister and lyotropic series are widely obsd., although not necessarily for the same class of expt. Moreover, we find that Hofmeister and lyotropic series are obsd. in non-aq. solvents even for expts. in which these series are not obsd. for water. Addnl., series reversal is seen for a given expt. when the solvent is changed. All this poses significant challenges for our understanding of specific-ion effects in aq. and non-aq. systems and also provides guideposts for future investigations.
- 28Rosenholm, J. B. Evaluation of particle charging in non-aqueous suspensions. Adv. Colloid Interface Sci. 2018, 259, 21– 43, DOI: 10.1016/j.cis.2018.06.004Google Scholar28https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1cXhtlWht7vN&md5=e3f9922bedc0c19a5d2d175e4a7b98b4Evaluation of particle charging in non-aqueous suspensionsRosenholm, Jarl B.Advances in Colloid and Interface Science (2018), 259 (), 21-43CODEN: ACISB9; ISSN:0001-8686. (Elsevier B.V.)Factors influencing the sign and size of effective surface (zeta) potential in suspensions of very low dielec. consts. are evaluated. For non-aq. suspensions it was found that Gutmann's donor no. (DN = neg. Lewis type molar acid-base adduct formation enthalpy) was successfully related to zeta potential changes, similarly as pH is optimal for aq. suspensions. Neg. molar proton dissocn. enthalpy (Br.vphi.nsted type HD no.), neg. hydrogen bond enthalpy (HB no.), logarithmic hydrogen bond equil. const. (molar Gibbs free energy), std. redn. potential of solvated protons (Eo(H+L/H2)), electrolytic dissocn. potential of water (Eo(H2O/H2, O2)) and electron exchange Fermi potentials could equally well be related to zeta potential changes. All these properties were linearly dependent on each other. When evaluating factors contributing to attractive and repulsive interaction energies, it is found that in order for the models to be relevant the extension of diffuse charging has to be much larger than the distance to repulsive barrier ensuring suspension stability. At this limit and at high surface potentials, the repulsive energy grows exceptionally large being in the range of lattice energy of each solid. The models fail when surface potential is low and the extension of diffuse charging is much smaller than the distance to repulsive barrier.
- 29Barthel, J.; Gores, H.-J.; Schmeer, G.; Wachter, R. Topics in Current Chemistry; Springer: Berlin Heidelberg, 1983; pp 33– 144.Google ScholarThere is no corresponding record for this reference.
- 30Izutsu, K. Electrochemistry in Nonaqueous Solutions; Wiley-VCH Verlag GmbH & Co. KGaA, 2002.Google ScholarThere is no corresponding record for this reference.
- 31Iwasaki, H.; Kimura, Y.; Uematsu, Y. Ubiquitous Preferential Water Adsorption to Electrodes in Water/1-Propanol Mixtures Detected by Electrochemical Impedance Spectroscopy. J. Phys. Chem. C 2023, 127, 23382– 23389, DOI: 10.1021/acs.jpcc.3c05320Google ScholarThere is no corresponding record for this reference.
- 32Iwai, S.; Suzuki, T.; Sakagami, H.; Miyamoto, K.; Chen, Z.; Konishi, M.; Villani, E.; Shida, N.; Tomita, I.; Inagi, S. Electropolymerization without an electric power supply. Commun. Chem. 2022, 5, 66, DOI: 10.1038/s42004-022-00682-8Google Scholar32https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB38XhsVSntrjK&md5=6ad1371f6ec9d67af6a5f557ae48af0bElectropolymerization without an electric power supplyIwai, Suguru; Suzuki, Taichi; Sakagami, Hiroki; Miyamoto, Kazuhiro; Chen, Zhenghao; Konishi, Mariko; Villani, Elena; Shida, Naoki; Tomita, Ikuyoshi; Inagi, ShinsukeCommunications Chemistry (2022), 5 (1), 66CODEN: CCOHCT; ISSN:2399-3669. (Nature Portfolio)Abstr.: Electrifying synthesis is now a common slogan among synthetic chemists. In addn. to the conventional two- or three-electrode systems that use batch-type cells, recent progress in org. electrochem. processes has been significant, including microflow electrochem. reactors, Li-ion battery-like technol., and bipolar electrochem. Herein we demonstrate an advanced electrosynthesis method without the application of elec. power based on the concept of streaming potential-driven bipolar electrochem. As a proof-of-concept study, the electrochem. oxidative polymn. of arom. monomers successfully yielded the corresponding polymer films on an electrode surface, which acted as an anode under the flow of electrolyte in a microchannel without an elec. power supply.
- 33Kolthoff, I. M. Acid-base equilibriums in dipolar aprotic solvents. Anal. Chem. 1974, 46, 1992– 2003, DOI: 10.1021/ac60349a005Google Scholar33https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaE2cXlsVWmsbs%253D&md5=a5ecea013728d8730d546e6cdbbd26e4Acid-base equilibriums in dipolar aprotic solventsKolthoff, I. M.Analytical Chemistry (1974), 46 (13), 1992-2003CODEN: ANCHAM; ISSN:0003-2700.A classification of org. solvents is proposed. The main emphasis is on the effect of homo- and heteroconjugate formation by H bonding on acid-base equilibria and on conductometric and potentiometric titrn. curves in dipolar aprotic solvents. The relation is discussed between resolution of acid strength and transfer activity coeffs. of ions and mols. A brief review is presented of analytical uses of protophobic solvents for the titrn. of very weak bases and of protophilic solvents for the titrn. of very weak acids. 77 Refs.
- 34Rossini, E.; Knapp, E. Proton solvation in protic and aprotic solvents. J. Comput. Chem. 2016, 37, 1082– 1091, DOI: 10.1002/jcc.24297Google Scholar34https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC28Xht1eksrk%253D&md5=c298ddbc68b3966d41021abebe4fc775Proton solvation in protic and aprotic solventsRossini, Emanuele; Knapp, Ernst-WalterJournal of Computational Chemistry (2016), 37 (12), 1082-1091CODEN: JCCHDD; ISSN:0192-8651. (John Wiley & Sons, Inc.)Protonation pattern strongly affects the properties of mol. systems. To det. protonation equil., proton solvation free energy, which is a central quantity in soln. chem., needs to be known. In this study, proton affinities (PAs), electrostatic energies of solvation, and pKA values were computed in protic and aprotic solvents. The proton solvation energy in acetonitrile (MeCN), methanol, water, and DMSO was detd. from computed and measured pKA values for a specially selected set of org. compds. pKA values were computed with high accuracy using a combination of quantum chem. and electrostatic approaches. Quantum chem. d. functional theory computations were performed evaluating PA in the gas-phase. The electrostatic contributions of solvation were computed solving the Poisson equation. The computations yield proton solvation free energies with high accuracy, which are in MeCN, MeOH, water, and DMSO -255.1, -265.9, -266.3, and -266.4 kcal/mol, resp., where the value for water is close to the consensus value of -265.9 kcal/mol. The pKA values of MeCN, MeOH, and DMSO in water correlates well with the corresponding proton solvation energies in these liqs., indicating that the solvated proton was attached to a single solvent mol. © 2016 Wiley Periodicals, Inc.
- 35Rossini, E.; Bochevarov, A. D.; Knapp, E. W. Empirical Conversion of pKa Values between Different Solvents and Interpretation of the Parameters: Application to Water, Acetonitrile, Dimethyl Sulfoxide, and Methanol. ACS Omega 2018, 3, 1653– 1662, DOI: 10.1021/acsomega.7b01895Google Scholar35https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1cXitl2lsbY%253D&md5=0fbd9b505269e3e46d02f1c3eb014eceEmpirical Conversion of pKa Values between Different Solvents and Interpretation of the Parameters: Application to Water, Acetonitrile, Dimethyl Sulfoxide, and MethanolRossini, Emanuele; Bochevarov, Art D.; Knapp, Ernst WalterACS Omega (2018), 3 (2), 1653-1662CODEN: ACSODF; ISSN:2470-1343. (American Chemical Society)An empirical conversion method (ECM) that transforms pKa values of arbitrary org. compds. from one solvent to the other is introduced. We demonstrate the method's usefulness and performance on pKa conversions involving water and org. solvents acetonitrile (MeCN), DMSO (Me2SO), and methanol (MeOH). We focus on the pKa conversion from the known ref. value in water to the other three org. solvents, although such a conversion can also be performed between any pair of the considered solvents. The ECM works with an additive parameter that is specific to a solvent and a mol. family (essentially characterized by a functional group that is titrated). We formally show that the method can be formulated with a single additive parameter, and that the extra multiplicative parameter used in other works is not required. The values of the additive parameter are detd. from known pKa data, and their interpretation is provided on the basis of physicochem. concepts. The data set of known pKa values is augmented with pKa values computed with the recently introduced electrostatic transform method, whose validity is demonstrated. For a validation of our method, we consider pKa conversions for two data sets of titratable compds. The first data set involves 81 relatively small mols. belonging to 19 different mol. families, with the pKa data available in all four considered solvents. The second data set involves 76 titratable mols. from 5 addnl. mol. families. These mols. are typically larger, and their exptl. pKa values are available only in Me2SO and water. The validation tests show that the agreement between the exptl. pKa data and the ECM predictions is generally good, with abs. errors often on the order of 0.5 pH units. The presence of a few outliers is rationalized, and obsd. trends with respect to mol. families are discussed.
- 36Borissova, M.; Gorbatšova, J.; Ebber, A.; Kaljurand, M.; Koel, M.; Vaher, M. Nonaqueous CE using contactless conductivity detection and ionic liquids as BGEs in ACN. Electrophoresis 2007, 28, 3600– 3605, DOI: 10.1002/elps.200700067Google ScholarThere is no corresponding record for this reference.
- 37Gouy, M. Sur la constitution de la charge électrique à la surface d’un électrolyte. J. Phys. Theor. Appl. 1910, 9, 457– 468, DOI: 10.1051/jphystap:019100090045700Google Scholar37https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaC3cXht1eluw%253D%253D&md5=d4237e8698c2acf4fcd7c80c970101ecConstitution of the Electric Charge at the Surface of an ElectrolyteGouy(1910), 9 (), 457-67 ISSN:.An amplification of an article previously abstracted (C. A., 4, 698).
- 38Chapman, D. L. A contribution to the theory of electrocapillarity. London Edinb. Philos. Mag. 1913, 25, 475– 481, DOI: 10.1080/14786440408634187Google ScholarThere is no corresponding record for this reference.
- 39Langmuir, I. THE ADSORPTION OF GASES ON PLANE SURFACES OF GLASS, MICA AND PLATINUM. J. Am. Chem. Soc. 1918, 40, 1361– 1403, DOI: 10.1021/ja02242a004Google Scholar39https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaC1cXht1KgsA%253D%253D&md5=3c22305e74ddd44d690168373bdd13c9The adsorption of gases on plane surfaces of glass, mica and platinumLangmuir, I.Journal of the American Chemical Society (1918), 40 (), 1361-1402CODEN: JACSAT; ISSN:0002-7863.According to L.'s hypothesis, gaseous mols. impinging on a liquid or solid surface do not in general rebound from it elastically, but are held or adsorbed on the surface by forces similar to those holding the atoms or group mols. of solid bodies. The adsorbed film should not exceed one mol. in thickness. Adsorption of permanent gases involves only secondary valence forces. In metals particularly, adsorption may be governed by primary valence forces. It is suggested that stoichiometric relations should govern the adsorption on a surface unless interfering effects caused by steric hindrance are involved. At room temp. the absorption by glass and mica was negligible, not over 1 % of the surface being covered by a single layer of mols. At lower temps. much larger quantities of gas were taken up. With Pt no absorption was observed at - 183° unless the Pt were first activated by proper heating. The adsorption of O2 was irreversible and corresponded to a monomolecular layer. CO likewise showed the same behavior. In the presence of one or the other gas adsorbed on the Pt the adsorbed and unadsorbed gases reacted immediately to form CO2.
- 40Rice, C. L.; Whitehead, R. Electrokinetic Flow in a Narrow Cylindrical Capillary. J. Phys. Chem. 1965, 69, 4017– 4024, DOI: 10.1021/j100895a062Google Scholar40https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaF28XlsFOktg%253D%253D&md5=3d4dd3365bd0d640395703e7436fc5adElectrokinetic flow in a narrow cylindrical capillaryRice, C. L.; Whitehead, R.Journal of Physical Chemistry (1965), 69 (11), 4017-24CODEN: JPCHAX; ISSN:0022-3654.An anal. study was made of electrokinetic flow in cylindrical capillaries with radii on the order of 10-5 cm. The work extends that of Burgreen and Nakache (CA 61, 15683b). Expressions are given for electroosmosis, streaming potential, and c.d. distribution. The results lead to a prediction of a max. in the electroviscous effect.
- 41Goldfarb, D. L.; Longinotti, M. P.; Corti, H. R. Electrical Conductances of Tetrabutylammonium and Decamethylferrocenium Hexafluorophosphate in Organic Solvents. J. Solution Chem. 2001, 30, 307– 322, DOI: 10.1023/A:1010334021934Google ScholarThere is no corresponding record for this reference.
- 42Hiemstra, T.; De Wit, J.; Van Riemsdijk, W. Multisite proton adsorption modeling at the solid/solution interface of (hydr)oxides: A new approach. J. Colloid Interface Sci. 1989, 133, 105– 117, DOI: 10.1016/0021-9797(89)90285-3Google Scholar42https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK3cXht1Cnuw%253D%253D&md5=3a2420d04d20dd5e805a7835084607b0Multisite proton adsorption modeling at the solid/solution interface of (hydr)oxides: a new approach. II. Application to various important (hydr)oxidesHiemstra, T.; De Wit, J. C. M.; Van Riemsdijk, W. H.Journal of Colloid and Interface Science (1989), 133 (1), 105-17CODEN: JCISA5; ISSN:0021-9797.At the solid/soln. interface of metal (hydro)oxides various types of O(H) and OH(H) groups are present, which differ in th no. of coordinating metal ions. The σ0-pH curves of metal (hydr)oxides are strongly detd. by the compn. and the relative extent of the various crystal planes of (hydr)oxides. The charging behavior is discuss for gibbsite (Al(OH)3), goethite (FeOOH), hematite (Fe2O3), rutile (TiO2), and silica (SiO2. New exptl. σ0-pH data for goethite and gibbsite are presented. Several important (hydr)oxides exhibit crystal faces which do not develop surface charge over a relatively wide pH range. An uncharged crystal face may be due to the presence of surface groups which are not reactive (inert) in the pH range under consideration, like the 001 face of gibbsite ad the 0001 face of hematite, or caused by the presence of two types of interacting charged surface groups of which the charge of one type is fully compensated by the other like at the 100 face of goethite. The charging behavior of silica and the 001 face of gibbsite is detd. by one type of reactive surface group with a large ΔpK for the consecutive protonation steps. The crystal structure imposes the presence of uncharged surface groups and this results in a quite different shape of σ0-pH curves for gibbsite and silica in comparison with the commonly obsd. σ0-pH curves for gibbsite and silica in comparison with the commonly obsd. σ0-pH curves of metal (hydr)oxides. The MUltiSIte complexation (MUSIC) model as developed by T. Hiemstra, et al. (1989) leads to a rather good prediction of σ0-pH curves for various metal (hydr)oxides using predicted affinity consts. for the various types of surface groups and Stern layer capacitance values and pair formation consts. estd. from the literature.
- 43Behrens, S. H.; Grier, D. G. The charge of glass and silica surfaces. J. Chem. Phys. 2001, 115, 6716– 6721, DOI: 10.1063/1.1404988Google Scholar43https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3MXnt1ams7o%253D&md5=c6018787255a16a50149e21cda2f3ca0The charge of glass and silica surfacesBehrens, Sven H.; Grier, David G.Journal of Chemical Physics (2001), 115 (14), 6716-6721CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)We present a method of calcg. the elec. charge d. of glass and silica surfaces in contact with aq. electrolytes for two cases of practical relevance that are not amenable to std. techniques: surfaces of low specific area at low ionic strength and surfaces interacting strongly with a second anionic surface.
- 44Bialik, E.; Stenqvist, B.; Fang, Y.; Östlund, Å.; Furó, I.; Lindman, B.; Lund, M.; Bernin, D. Ionization of Cellobiose in Aqueous Alkali and the Mechanism of Cellulose Dissolution. J. Phys. Chem. Lett. 2016, 7, 5044– 5048, DOI: 10.1021/acs.jpclett.6b02346Google Scholar44https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC28XhvFSitrrL&md5=e6804dccfdcd71184152103737965cdeIonization of Cellobiose in Aqueous Alkali and the Mechanism of Cellulose DissolutionBialik, Erik; Stenqvist, Bjoern; Fang, Yuan; Oestlund, Aasa; Furo, Istvan; Lindman, Bjoern; Lund, Mikael; Bernin, DianaJournal of Physical Chemistry Letters (2016), 7 (24), 5044-5048CODEN: JPCLCD; ISSN:1948-7185. (American Chemical Society)Cellulose, one of the most abundant renewable resources, is insol. in most common solvents but dissolves in aq. alkali under a narrow range of conditions. To elucidate the solubilization mechanism, we performed electrophoretic NMR on cellobiose, a subunit of cellulose, showing that cellobiose acts as an acid with two dissocn. steps at pH 12 and 13.5. Chem. shift differences between cellobiose in NaOH and NaCl were estd. using 2D NMR and compared to DFT shift differences upon deprotonation. The dissocn. steps are the deprotonation of the hemiacetal OH group and the deprotonation of one of four OH groups on the nonreducing anhydroglucose unit. MD simulations reveal that aggregation is suppressed upon charging cellulose chains in soln. Our findings strongly suggest that cellulose is to a large extent charged in concd. aq. alkali, a seemingly crucial factor for solubilization. This insight, overlooked in the current literature, is important for understanding cellulose dissoln. and for synthesis of new sustainable materials.
- 45Malerz, S.; Mudryk, K.; Tomaník, L.; Stemer, D.; Hergenhahn, U.; Buttersack, T.; Trinter, F.; Seidel, R.; Quevedo, W.; Goy, C.; Wilkinson, I.; Thürmer, S.; Slavíček, P.; Winter, B. Following in Emil Fischer’s Footsteps: A Site-Selective Probe of Glucose Acid–Base Chemistry. J. Phys. Chem. A 2021, 125, 6881– 6892, DOI: 10.1021/acs.jpca.1c04695Google Scholar45https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB3MXhs1KgtrnF&md5=02664b9d409bace291887eed06258865Following in Emil Fischer's Footsteps: A Site-Selective Probe of Glucose Acid-Base ChemistryMalerz, Sebastian; Mudryk, Karen; Tomanik, Lukas; Stemer, Dominik; Hergenhahn, Uwe; Buttersack, Tillmann; Trinter, Florian; Seidel, Robert; Quevedo, Wilson; Goy, Claudia; Wilkinson, Iain; Thuermer, Stephan; Slavicek, Petr; Winter, BerndJournal of Physical Chemistry A (2021), 125 (32), 6881-6892CODEN: JPCAFH; ISSN:1089-5639. (American Chemical Society)Liq.-jet photoelectron spectroscopy was applied to det. the first acid dissocn. const. (pKa) of aq.-phase glucose while simultaneously identifying the spectroscopic signature of the resp. deprotonation site. Valence spectra from solns. at pH values below and above the first pKa reveal a change in glucose's lowest ionization energy upon the deprotonation of neutral glucose and the subsequent emergence of its anionic counterpart. Site-specific insights into the soln.-pH-dependent mol. structure changes are also shown to be accessible via C 1s photoelectron spectroscopy. The spectra reveal a considerably lower C 1s binding energy of the carbon site assocd. with the deprotonated hydroxyl group. The occurrence of photoelectron spectral fingerprints of cyclic and linear glucose prior to and upon deprotonation are also discussed. The exptl. data are interpreted with the aid of electronic structure calcns. Our findings highlight the potential of liq.-jet photoelectron spectroscopy to act as a site-selective probe of the mol. structures that underpin the acid-base chem. of polyprotic systems with relevance to environmental chem. and biochem.
- 46Uematsu, Y.; Bonthuis, D. J.; Netz, R. R. Charged Surface-Active Impurities at Nanomolar Concentration Induce Jones–Ray Effect. J. Phys. Chem. Lett. 2018, 9, 189– 193, DOI: 10.1021/acs.jpclett.7b02960Google Scholar46https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2sXitVSnurbK&md5=ff9329b83b3083f688fa5cc3ee9bd1e8Charged Surface-Active Impurities at Nanomolar Concentration Induce Jones-Ray EffectUematsu, Yuki; Bonthuis, Douwe Jan; Netz, Roland R.Journal of Physical Chemistry Letters (2018), 9 (1), 189-193CODEN: JPCLCD; ISSN:1948-7185. (American Chemical Society)The electrolyte surface tension exhibits a characteristic min. around a salt concn. of 1 mM for all ion types, known as the Jones-Ray effect. A consistent description of the exptl. surface tension of salts, bases, and acids is possible by assuming charged impurities in the water with a surface affinity typical for surfactants. Comparison with exptl. data yields an impurity concn. in the nanomolar range, well below the typical exptl. detection limit. Modeling reveals salt-screening enhanced impurity adsorption as the mechanism behind the Jones-Ray effect: for very low salt concn. added salt screens the electrostatic repulsion between impurities at the surface, which dramatically increases impurity adsorption and thereby reduces the surface tension.
- 47Plett, T.; Shi, W.; Zeng, Y.; Mann, W.; Vlassiouk, I.; Baker, L. A.; Siwy, Z. S. Rectification of nanopores in aprotic solvents–transport properties of nanopores with surface dipoles. Nanoscale 2015, 7, 19080– 19091, DOI: 10.1039/C5NR06340JGoogle ScholarThere is no corresponding record for this reference.
- 48Nasir, S.; Ali, M.; Ramirez, P.; Froehlich, K.; Cervera, J.; Mafe, S.; Ensinger, W. Ionic conduction through single-pore and multipore polymer membranes in aprotic organic electrolytes. J. Membr. Sci. 2021, 635, 119505 DOI: 10.1016/j.memsci.2021.119505Google Scholar48https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB3MXhsVSit73I&md5=46ab0d4fee27da11ca0ddc3dfdd06f01Ionic conduction through single-pore and multipore polymer membranes in aprotic organic electrolytesNasir, Saima; Ali, Mubarak; Ramirez, Patricio; Froehlich, Kristina; Cervera, Javier; Mafe, Salvador; Ensinger, WolfgangJournal of Membrane Science (2021), 635 (), 119505CODEN: JMESDO; ISSN:0376-7388. (Elsevier B.V.)We exptl. characterize the ionic conduction of single and multipore nanoporous membranes in aprotic org. electrolytes. To this end, soft-etched (SE) membranes with pore diams. in the nanometer range and track-etched (TE) membranes with pore diams. in the tens of nanometers range are investigated. In aq. conditions, the membrane ionic conduction rates follow the same trend of the bulk soln. conductivities. However, the ionic transport through the narrow SE-nanopores dramatically decreases in aprotic electrolytes due to the formation of solvated metal cations and their adsorption on the pore surface. The current-voltage recordings of single conical nanopores in aprotic electrolyte solns. with different water mole fractions reveal that the solvated metal ion (M) species [M-(solvent)4]+ formed in acetonitrile solvent are more tightly bounded to the pore walls compared with the cationic chelates obtained in propylene carbonate solvent. The basic findings reported here should be of interest for ionic/mol. nanofiltration processes in non-aq. conditions as well as for moisture sensitive and energy storage nanofluidic devices.
- 49Polster, J. W.; Souna, A. J.; Motevaselian, M. H.; Lucas, R. A.; Tran, J. D.; Siwy, Z. S.; Aluru, N. R.; Fourkas, J. T. The electrical-double layer revisited. Nat. Sci. 2022, 2, e20210099 DOI: 10.1002/ntls.20210099Google ScholarThere is no corresponding record for this reference.
- 50Souna, A. J.; Motevaselian, M. H.; Polster, J. W.; Tran, J. D.; Siwy, Z. S.; Aluru, N. R.; Fourkas, J. T. Beyond the electrical double layer model: ion-dependent effects in nanoscale solvent organization. Phys. Chem. Chem. Phys. 2024, 26, 6726– 6735, DOI: 10.1039/D3CP05712GGoogle ScholarThere is no corresponding record for this reference.
- 51Yin, X.; Zhang, S.; Dong, Y.; Liu, S.; Gu, J.; Chen, Y.; Zhang, X.; Zhang, X.; Shao, Y. Ionic Current Rectification in Organic Solutions with Quartz Nanopipettes. Anal. Chem. 2015, 87, 9070– 9077, DOI: 10.1021/acs.analchem.5b02337Google Scholar51https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2MXht1CitL7K&md5=3fe8404e459b8844d48d68429e562fb5Ionic Current Rectification in Organic Solutions with Quartz NanopipettesYin, Xiaohong; Zhang, Shudong; Dong, Yitong; Liu, Shujuan; Gu, Jing; Chen, Ye; Zhang, Xin; Zhang, Xianhao; Shao, YuanhuaAnalytical Chemistry (Washington, DC, United States) (2015), 87 (17), 9070-9077CODEN: ANCHAM; ISSN:0003-2700. (American Chemical Society)The study of behaviors of ionic current rectification (ICR) in org. solns. with quartz nanopipettes is reported. ICR can be obsd. even in org. solns. using quartz pipets with diams. varied from several to dozens of nanometers, and the direction of ICR is quite different from the ICR obsd. in aq. phase. The influences of pore size, electrolyte concn., and surface charge on the ICR were studied carefully. Water in org. solns. affects the direction and extent of ICR significantly. Mechanisms about the formation of an elec. double layer (EDL) on silica in org. solns. with different amt. of water are proposed. An improved method, which can be employed to detect trace water in org. solns., was implemented based on Au ultramicroelectrodes with cathodic differential pulse stripping voltammetry.
- 52Spångberg, D.; Hermansson, K. The solvation of Li+ and Na+ in acetonitrile from ab initio-derived many-body ion–solvent potentials. Chem. Phys. 2004, 300, 165– 176, DOI: 10.1016/j.chemphys.2004.01.011Google Scholar52https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2cXisl2lsb4%253D&md5=8962ba45e76fb9c036f074ac17334b03The solvation of Li+ and Na+ in acetonitrile from ab initio-derived many-body ion-solvent potentialsSpangberg, Daniel; Hermansson, KerstiChemical Physics (2004), 300 (1-3), 165-176CODEN: CMPHC2; ISSN:0301-0104. (Elsevier Science B.V.)Several Li+- and Na+-acetonitrile models were derived from ab initio calcns. at the counterpoise-cor. MP2/TZV++(d,p) level for distorted ion-(MeCN)n clusters with n=1, 4 and 6. Two different many-body ion-acetonitrile models were constructed: an effective three-body potential for use with the six-site effective pair model of Bohm et al., and an effective polarizable many-body model. The polarizable acetonitrile model used in the latter model is a new empirical model which was also derived in the present paper. Mainly for comparative purposes, two ion-acetonitrile pair potentials were also constructed from the ab initio cluster calcns.: one pure pair potential and one effective pair potential. Using all these potential models, MD simulations in the NPT ensemble were performed for the pure acetonitrile liq. and for Li+(MeCN) and Na+(MeCN) solns. with 1 ion in 512 solvent mols. and with a simulation time of at least 120 ps per system. Thermodn. properties, solvation-shell structure and the self-diffusion coeff. of the ions and of the solvent mols. were calcd. and compared between the different models and with exptl. data, where available. The Li+ ion is found to be four-coordinated when the new many-body potentials are used, in contrast to the six-coordinated structure obtained for the pure pair and effective pair potentials. The coordination no. of Na+ is close to six for all the models derived here, although the coordination no. becomes slightly smaller with the many-body potentials. For both ions, the solvent mols. in the first shell point their nitrogen ends towards the cation, while in the second shell the opposite orientation is the most common.
- 53Amara, S.; Toulc’Hoat, J.; Timperman, L.; Biller, A.; Galiano, H.; Marcel, C.; Ledigabel, M.; Anouti, M. Comparative Study of Alkali-Cation-Based (Li+, Na+, K+) Electrolytes in Acetonitrile and Alkylcarbonates. ChemPhysChem 2019, 20, 581– 594, DOI: 10.1002/cphc.201801064Google Scholar53https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1MXisVGnsrk%253D&md5=783dc126fb8ffa9bc3806c92ab647ad8Comparative Study of Alkali-Cation-Based (Li+, Na+, K+) Electrolytes in Acetonitrile and AlkylcarbonatesAmara, Samia; Toulc'Hoat, Joel; Timperman, Laure; Biller, Agnes; Galiano, Herve; Marcel, Corinne; Ledigabel, Matthieu; Anouti, MeriemChemPhysChem (2019), 20 (4), 581-594CODEN: CPCHFT; ISSN:1439-4235. (Wiley-VCH Verlag GmbH & Co. KGaA)The development of a suitable functional electrolyte is urgently required for fast-charging and high-voltage alkali-ion (Li, Na, K) batteries as well as next-generation hybrids supercapacitors. Many recent works focused on an optimal selection of electrolytes for alkali-ion based systems and their electrochem. performance but the understanding of the fundamental aspect that explains their different behavior is rare. Herein, we report a comparative study of transport properties for LiPF6, NaPF6, KPF6 in acetonitrile (AN) and a binary mixt. of ethylene carbonate (EC), di-Me carbonate (DMC): (EC/DMC : 1/1, weigh) through conductivities, densities and viscosities measurements in wide temp. domain. By application of the Stokes-Einstein, Nernst-Einstein, and Jones Dole equations, the effective ionic solvated radius of cation (reff), the ionic dissocn. coeff. (αD) and structuring Jones Dole's parameters (A, B) for salt are calcd. and discussed according to solvent or cation nature as a function of temp. From the results, we demonstrate that better mobility of potassium can be explained by the nature of the ion-ion and ion-solvent interactions due to its polarizability. In the same time, the predominance of triple ions in the case of K+, is a disadvantage at high concn.
- 54Ding, F.; Hu, Z.; Zhong, Q.; Manfred, K.; Gattass, R. R.; Brindza, M. R.; Fourkas, J. T.; Walker, R. A.; Weeks, J. D. Interfacial Organization of Acetonitrile: Simulation and Experiment. J. Phys. Chem. C 2010, 114, 17651– 17659, DOI: 10.1021/jp104597zGoogle Scholar54https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3cXhtVGrsbvN&md5=8bcd247fc15255fce039a48a613a595fInterfacial Organization of Acetonitrile: Simulation and ExperimentDing, Feng; Hu, Zhonghan; Zhong, Qin; Manfred, Katherine; Gattass, Rafael R.; Brindza, Michael R.; Fourkas, John T.; Walker, Robert A.; Weeks, John D.Journal of Physical Chemistry C (2010), 114 (41), 17651-17659CODEN: JPCCCK; ISSN:1932-7447. (American Chemical Society)Mol. dynamics simulations and vibrational sum frequency generation (VSFG) expts. in the methyl-stretching spectral region have been used to study acetonitrile at the silica/liq., silica/vapor, and liq./vapor interfaces. Our simulations show that, at the silica/liq. interface, acetonitrile takes on a considerably different structure than in the bulk liq. The interfacial structure is reminiscent of a lipid bilayer, and this type of ordering persists for tens of Ångstroms into the bulk liq. This result has important implications for processes involving solid/acetonitrile interfaces, such as heterogeneous catalysis and chromatog. sepns. Fitting of VSFG data that have an extremely low nonresonant background contribution provides strong evidence for interfacial populations pointing in opposite directions at these interfaces, in agreement with our simulations. The picture developed from our simulations and expts. reconciles conflicting interpretations of data from previous exptl. studies of interfacial acetonitrile.
- 55Pusić, T.; Grancarić, A. M.; Soljaçić, I.; Ribitsch, V. The effect of mercerisation on the electrokinetic potential of cotton. Color. Technol. 1999, 115, 121– 124, DOI: 10.1111/j.1478-4408.1999.tb00308.xGoogle ScholarThere is no corresponding record for this reference.
- 56Sadeghi-Kiakhani, M.; Safapour, S. Salt-free reactive dyeing of the cotton fabric modified with chitosan-poly(propylene imine) dendrimer hybrid. Fibers and Polym. 2015, 16, 1075– 1081, DOI: 10.1007/s12221-015-1075-9Google ScholarThere is no corresponding record for this reference.
- 57Grancarić, A. M.; Tarbuk, A.; Hadžić, S.; Simončič, B. From Raw to Finished Cotton─Characterization by Interface Phenomena. Autex Res. J. 2023, 23, 184– 192, DOI: 10.2478/aut-2021-0055Google ScholarThere is no corresponding record for this reference.
- 58Uematsu, Y.; Bonthuis, D. J.; Netz, R. R. Nanomolar Surface-Active Charged Impurities Account for the Zeta Potential of Hydrophobic Surfaces. Langmuir 2020, 36, 3645– 3658, DOI: 10.1021/acs.langmuir.9b03795Google Scholar58https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB3cXkvFOqu7Y%253D&md5=59e6893c6592b69f5996059ab9fc5c90Nanomolar Surface-Active Charged Impurities Account for the Zeta Potential of Hydrophobic SurfacesUematsu, Yuki; Bonthuis, Douwe Jan; Netz, Roland R.Langmuir (2020), 36 (13), 3645-3658CODEN: LANGD5; ISSN:0743-7463. (American Chemical Society)The electrification of hydrophobic surfaces is an intensely debated subject in phys. chem. We theor. study the ζ potential of hydrophobic surfaces for varying pH and salt concn. by solving the Poisson-Boltzmann and Stokes equations with individual ionic adsorption affinities. Using the ionic surface affinities extd. from the exptl. measured surface tension of the air-electrolyte interface, we first show that the interfacial adsorption and repulsion of small inorg. ions such as H3O+, OH-, HCO3-, and CO32- cannot account for the ζ potential obsd. in expts. because the surface affinities of these ions are too small. Even if we take hydrodynamic slip into account, the characteristic dependence of the ζ potential on pH and salt concn. cannot be reproduced. Instead, to explain the sizable exptl. measured ζ potential of hydrophobic surfaces, we assume minute amts. of impurities in the water and include the impurities' acidic and basic reactions with water. We find good agreement between our predictions and the reported exptl. ζ potential data of various hydrophobic surfaces if we account for impurities that consist of a mixt. of weak acids (pKa = 5-7) and weak bases (pKb = 12) at a concn. of the order of 10-7 M.
- 59Sghaier, S.; Zbidi, F.; Zidi, M. Characterization of Doum Palm Fibers After Chemical Treatment. Text. Res. J. 2009, 79, 1108– 1114, DOI: 10.1177/0040517508101623Google ScholarThere is no corresponding record for this reference.
- 60Brož, Z.; Epstein, N. Electrokinetic flow through porous media composed of fine cylindrical capillaries. J. Colloid Interface Sci. 1976, 56, 605– 612, DOI: 10.1016/0021-9797(76)90127-2Google ScholarThere is no corresponding record for this reference.
- 61Levine, S.; Marriott, J.; Neale, G.; Epstein, N. Theory of electrokinetic flow in fine cylindrical capillaries at high zeta-potentials. J. Colloid Interface Sci. 1975, 52, 136– 149, DOI: 10.1016/0021-9797(75)90310-0Google ScholarThere is no corresponding record for this reference.
- 62Uematsu, Y.; Netz, R. R.; Bocquet, L.; Bonthuis, D. J. Crossover of the Power-Law Exponent for Carbon Nanotube Conductivity as a Function of Salinity. J. Phys. Chem. B 2018, 122, 2992– 2997, DOI: 10.1021/acs.jpcb.8b01975Google Scholar62https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1cXjs1Sqt7Y%253D&md5=32f80132c17eb669b616c0478bd337acCrossover of the Power-Law Exponent for Carbon Nanotube Conductivity as a Function of SalinityUematsu, Yuki; Netz, Roland R.; Bocquet, Lyderic; Bonthuis, Douwe JanJournal of Physical Chemistry B (2018), 122 (11), 2992-2997CODEN: JPCBFK; ISSN:1520-5207. (American Chemical Society)On the basis of the Poisson-Boltzmann equation in cylindrical coordinates, we calc. the cond. of a single charged nanotube filled with electrolyte. The cond. as a function of the salt concn. follows a power-law, the exponent of which has been controversially discussed in the literature. We use the co-ion-exclusion approxn. and obtain the crossover between different asymptotic power-law behaviors anal. Numerically solving the full Poisson-Boltzmann equation, we also calc. the complete diagram of exponents as a function of the salt concn. and the pH for tubes with different radii and pKa values. We apply our theory to recent exptl. results on carbon nanotubes using the pKa as a fit parameter. In good agreement with the exptl. data, the theory shows power-law behavior with the exponents 1/3 at high pH and 1/2 at low pH, with a crossover depending on salt concn., tube radius and pKa.
Cited By
This article has not yet been cited by other publications.
Article Views
Altmetric
Citations
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 62 other publications.
- 1Russel, W. B.; Saville, D. A.; Schowalter, W. R. Colloidal Dispersions; Cambridge University Press, 1989.There is no corresponding record for this reference.
- 2Uematsu, Y. Electrification of water interface. J. Phys.: Condens. Matter 2021, 33, 423001, DOI: 10.1088/1361-648X/ac15d52https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB3MXhvF2hs77O&md5=869f91e31de4b7966eb00e86a0bcf12aElectrification of water interfaceUematsu, YukiJournal of Physics: Condensed Matter (2021), 33 (42), 423001CODEN: JCOMEL; ISSN:0953-8984. (IOP Publishing Ltd.)A review. The surface charge of a water interface dets. many fundamental processes in phys. chem. and interface science, and it has been intensively studied for over a hundred years. We summarize exptl. methods to characterize the surface charge densities developed so far: electrokinetics, double-layer force measurements, potentiometric titrn., surface-sensitive nonlinear spectroscopy, and surface-sensitive mass spectrometry. Then, we elucidate phys. ion adsorption and chem. electrification as examples of electrification mechanisms. In the end, novel effects on surface electrification are discussed in detail. We believe that this clear overview of state of the art in a charged water interface will surely help the fundamental progress of physics and chem. at interfaces in the future.
- 3Siria, A.; Poncharal, P.; Biance, A.-L.; Fulcrand, R.; Blase, X.; Purcell, S. T.; Bocquet, L. Giant osmotic energy conversion measured in a single transmembrane boron nitride nanotube. Nature 2013, 494, 455– 458, DOI: 10.1038/nature118763https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3sXjtFSrt7Y%253D&md5=6b675bc4d36dfb57cb309bad50613859Giant osmotic energy conversion measured in a single transmembrane boron nitride nanotubeSiria, Alessandro; Poncharal, Philippe; Biance, Anne-Laure; Fulcrand, Remy; Blase, Xavier; Purcell, Stephen T.; Bocquet, LydericNature (London, United Kingdom) (2013), 494 (7438), 455-458CODEN: NATUAS; ISSN:0028-0836. (Nature Publishing Group)New models of fluid transport are expected to emerge from the confinement of liqs. at the nanoscale, with potential applications in ultrafiltration, desalination and energy conversion. Nevertheless, advancing our fundamental understanding of fluid transport on the smallest scales requires mass and ion dynamics to be ultimately characterized across an individual channel to avoid averaging over many pores. A major challenge for nanofluidics thus lies in building distinct and well-controlled nanochannels, amenable to the systematic exploration of their properties. Here we describe the fabrication and use of a hierarchical nanofluidic device made of a boron nitride nanotube that pierces an ultrathin membrane and connects two fluid reservoirs. Such a transmembrane geometry allows the detailed study of fluidic transport through a single nanotube under diverse forces, including elec. fields, pressure drops and chem. gradients. Using this device, we discover very large, osmotically induced elec. currents generated by salinity gradients, exceeding by two orders of magnitude their pressure-driven counterpart. We show that this result originates in the anomalously high surface charge carried by the nanotube's internal surface in water at large pH, which we independently quantify in conductance measurements. The nano-assembly route using nanostructures as building blocks opens the way to studying fluid, ionic and mol. transport on the nanoscale, and may lead to biomimetic functionalities. Our results furthermore suggest that boron nitride nanotubes could be used as membranes for osmotic power harvesting under salinity gradients.
- 4Gortner, R. A. Electrokinetics XXIII. Electrokinetics as a tool for the study of the molecular structure of organic compounds. Trans. Faraday Soc. 1940, 35, 63, DOI: 10.1039/TF9403500063There is no corresponding record for this reference.
- 5Westermann-Clark, G.; Christoforou, C. Note on nonaqueous electrokinetic transport in charged porous media. J. Membr. Sci. 1984, 20, 325– 338, DOI: 10.1016/S0376-7388(00)82009-85https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaL2cXlvFyktrw%253D&md5=8117a095c3950f93c6a60a515684abfeNote on nonaqueous electrokinetic transport in charged porous mediaWestermann-Clark, G. B.; Christoforou, C. C.Journal of Membrane Science (1984), 20 (3), 325-39CODEN: JMESDO; ISSN:0376-7388.A model for electrokinetic transport in charged capillaries is compared with expts. by using nonaq. LiCl solns. The electrokinetic parameters considered are the pore fluid cond. and the concn. potential. Methanol/water mixts. were the solvents and track-etched mica membranes with a well-characterized pore structure were the porous medium. The electrolyte concns. used were such that the Debye lengths of solns. in pores ranged from much smaller to much larger than the radii of the pores. The space-charge model is capable of describing qual. the trend of the electrokinetic data, but as expected, at higher concns. the model fails, probably because the assumption that ion-ion interactions are negligible no longer holds. The pore fluid cond. depends strongly on the dielec. const. of the solvent; the abs. value of the pore wall charge tends to decrease with the lowering of the solvent dielec. const., and the wall charge tends to increase with the concn. of the chloride ion.
- 6Lucas, R. A.; Lin, C.-Y.; Siwy, Z. S. Electrokinetic Phenomena in Organic Solvents. J. Phys. Chem. B 2019, 123, 6123– 6131, DOI: 10.1021/acs.jpcb.9b049696https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1MXhtF2ks7fM&md5=e06b885a013b890be5cabf431cf8877aElectrokinetic Phenomena in Organic SolventsLucas, Rachel A.; Lin, Chih-Yuan; Siwy, Zuzanna S.Journal of Physical Chemistry B (2019), 123 (28), 6123-6131CODEN: JPCBFK; ISSN:1520-5207. (American Chemical Society)Solid/liq. interfaces play a key role in sepn. processes, energy storage devices, and transport in nanoscale systems. Nanopores and mesopores with well-defined geometry and chem. characteristics have been a valuable tool to unravel electrochem. properties of interfaces, but the majority of studies have been focused on aq. solns. Here, we present expts. and numerical modeling aimed at characterizing effective surface charge of polymer pores in mixts. of water and alcs. as well as in propylene carbonate and acetone. The charge properties of pore walls are probed through anal. of current-voltage curves recorded in the presence of salt concn. gradients. The presence and direction of electro-osmotic flow lead to asym. current-voltage curves, with rectification characteristics detd. by the polarity of surface charge. The results suggest that the effective surface charge of the pore walls depends not only on the type of solvent but also on the concn. of the electrolyte and voltage. We identified conditions at which polymer pores that are neg. charged in aq. solns. become pos. charged in propylene carbonate and acetone. The findings are of importance for nonaq. sepns., fundamental knowledge on solid/liq. interfaces in org. media, and prepn. of porous devices with tunable surface charge characteristics.
- 7Russell, W. S.; Siwy, Z. Enhanced electro-osmosis in propylene carbonate salt solutions. J. Chem. Phys. 2021, 154, 134707, DOI: 10.1063/5.00444027https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB3MXotlGnsL0%253D&md5=5f659f27bbe30ab75e421d54494131a6Enhanced electro-osmosis in propylene carbonate salt solutionsRussell, Wilfred Shelby; Siwy, ZuzannaJournal of Chemical Physics (2021), 154 (13), 134707CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)Properties of solid-liq. interfaces and surface charge characteristics mediate ionic and mol. transport through porous systems, affecting many processes such as sepns. Herein, the authors report expts. designed to probe the electrochem. properties of solid-liq. interfaces using a model system of a single polyethylene terephthalate (PET) pore in contact with aq. and propylene carbonate solns. of LiClO4. First, the existence and polarity of surface charges were inferred from current-voltage curves recorded when a pore was placed in contact with a LiClO4 concn. gradient. Second, the electroosmotic transport of uncharged polystyrene particles through the PET pore provided information on the polarity and the magnitude of the pore walls' zeta potential. The authors' expts. show that the PET pores become effectively pos. charged when in contact with LiClO4 solns. in propylene carbonate, even though in aq. LiClO4, the same pores are neg. charged. Addnl., the electroosmotic velocity of the particles revealed a significantly higher magnitude of the pos. zeta potential of the pores in propylene carbonate compared to the magnitude of the neg. zeta potential in H2O. The presented methods of probing the properties of solid-liq. interfaces are expected to be applicable to a wide variety of solid and liq. systems. (c) 2021 American Institute of Physics.
- 8Suzuki, Y.; Mizuhata, M. Predictive Zeta Potential Measurement Method Applicable to Nonaqueous Solvents in High-concentration Dispersion Systems for the System of LiClO4–Propylene Carbonate Solution and LiCoO2 Powder Sheet. Electrochemistry 2022, 90, 103001– 103001, DOI: 10.5796/electrochemistry.22-66050There is no corresponding record for this reference.
- 9Schwer, C.; Kenndler, E. Electrophoresis in fused-silica capillaries: the influence of organic solvents on the electroosmotic velocity and the zeta potential. Anal. Chem. 1991, 63, 1801– 1807, DOI: 10.1021/ac00017a0269https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK3MXltVSitro%253D&md5=6b5df657d236b8d895b4a924bba8eb6cElectrophoresis in fused-silica capillaries: the influence of organic solvents on the electroosmotic velocity and the ζ potentialSchwer, Christine; Kenndler, ErnstAnalytical Chemistry (1991), 63 (17), 1801-7CODEN: ANCHAM; ISSN:0003-2700.The influence of pH and solvent compn. on the electroosmotic flow in fused-silica capillaries was studied. Binary mixts. of H2O and a protic (MeOH, EtOH, iso-PrOH) or an aprotic dipolar solvent (MeCN, Me2CO, DMSO) were used for the buffer electrolyte. The dependence of the electroosmotic flow on the pH was detd. in the pH range from 3 to 11 for pure aq. soln. and for solns. contg. 50 vol.% org. solvent. Adding org. solvents to the electrolyte, the inflection points of the resulting curves, corresponding to the pK values of the surface silanol groups, are shifted to higher values. At high pH, increasing the fraction of org. cosolvent generally decreases the electroosmotic velocity. Zeta potentials were calcd. from the Smoluchowski equation by using electroosmotic velocity data and values for the ratio of the viscosity coeff. and dielec. const. for binary mixts., taken from the authors own measurements and from the literature. With the exception of Me2CO-H2O mixts., the ζ potentials obtained show a similar trend to that of the electroosmotic velocity; they decrease with increasing content of org. solvent. This trend is explained by changes of the dielec. properties of the elec. double layer and of the charge generation on the fused-silica surface.
- 10Wright, P. B.; Lister, A. S.; Dorsey, J. G. Behavior and Use of Nonaqueous Media without Supporting Electrolyte in Capillary Electrophoresis and Capillary Electrochromatography. Anal. Chem. 1997, 69, 3251– 3259, DOI: 10.1021/ac961318610https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK2sXkslKmtLc%253D&md5=12548b0d583d659dcf0685819dfc1d56Behavior and Use of Nonaqueous Media without Supporting Electrolyte in Capillary Electrophoresis and Capillary ElectrochromatographyWright, Paul B.; Lister, Ashley S.; Dorsey, John G.Analytical Chemistry (1997), 69 (16), 3251-3259CODEN: ANCHAM; ISSN:0003-2700. (American Chemical Society)Five nonaq. solvents (acetonitrile, methanol, DMF, DMSO, formamide) and deionized water were studied for their ability to support electroosmotic flow (EOF) without electrolytic additives. In general, flow is equal to or greater than flow with typical CE buffer systems. The magnitude of EOF was detd. for each solvent by open tubular capillary electrophoresis (CE) and related to viscosity (η), dielec. const. (ε), and the ratio of dielec. const. to viscosity (ε/η). Zeta potentials (ζ) were derived indirectly from flow data and tabulated. Comparisons of flow behavior and ζ were made between pure solvents and conventional CE buffers, and questions of equilibration and reproducibility were addressed. Similar expts. were performed using hydroorg. mobile phases (ACN/water, MeOH/water) across the complete compositional range (100% water-100% org.), with flow characteristics and ζ reported for each mobile phase system. Packed capillary columns (5-μm ODS) were evaluated for flow and retention stability under capillary electrochromatog. (CEC) conditions. A sepn. of 11 polycyclic arom. hydrocarbons was performed in under 13 min by CEC with an ACN/water mobile phase. Reduced plate heights (h) were calcd. between 2.5 and 3.0 for solutes with capacity factors (k') up to 4.5 for the most retained solute.
- 11Tjo̷rnelund, J.; Hansen, S. H. The effect of water on separations in non-aqueous capillary electrophoresis systems. Chromatographia 1997, 44, 5– 9, DOI: 10.1007/BF02466508There is no corresponding record for this reference.
- 12Valkó, I. E.; Sirén, H.; Riekkola, M.-L. Characteristics of electroosmotic flow in capillary electrophoresis in water and in organic solvents without added ionic species. J. Microcolumn Sep. 1999, 11, 199– 208, DOI: 10.1002/(SICI)1520-667X(1999)11:3<199::AID-MCS5>3.0.CO;2-D12https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK1MXhtlGntLo%253D&md5=5565670086434868d01b1c6931468c0fCharacteristics of electroosmotic flow in capillary electrophoresis in water and in organic solvents without added ionic speciesValko, Istvan E.; Siren, Heli; Riekkola, Marja-LiisaJournal of Microcolumn Separations (1999), 11 (3), 199-208CODEN: JMSEEJ; ISSN:1040-7685. (John Wiley & Sons, Inc.)Although proper buffering is essential for repeatable sepns. in capillary electrophoresis, electroosmotic flow can be developed in many solvents without the addn. of buffer components. In this study we detected electroosmotic flow in fused silica capillary in the following solvents without the addn. of ionic species: water, deuterium oxide, acetonitrile, acetone, 2-butanone, formamide, N-methylformamide, N,N-dimethylformamide, methanol, ethanol, 1-propanol, DMSO, Et acetate, THF, and morpholine. The electroosmotic flow was in the range of 3.4 X 10-9-1.8 X 10-7 m2/V s. About 100 times weaker electroosmosis was detected in glacial acetic acid. We also detected electroosmotic flow in N-methylacetamide, which is solid at room temp. (melting range 28-30°C). No electroosmotic flow was found in inert solvents such as n-hexane and chloroform. The zeta potential of the fused silica capillary was calcd. for the solvents where electroosmotic flow was found. The zeta potential of the capillary wall was also studied in water-methanol mixts., where it has a max. at about 40% methanol concn.
- 13Riekkola, M.-L.; Jussila, M.; Porras, S. P.; Valkó, I. E. Non-aqueous capillary electrophoresis. J. Chromatogr. A 2000, 892, 155– 170, DOI: 10.1016/S0021-9673(00)00108-4There is no corresponding record for this reference.
- 14Geiser, L.; Mirgaldi, M.; Veuthey, J.-L. Determination of electroosmotic flow in nonaqueous capillary electrophoresis. J. Chromatogr. A 2005, 1068, 75– 81, DOI: 10.1016/j.chroma.2005.02.001There is no corresponding record for this reference.
- 15Kitahara, A. Zeta potential in nonaqueous media and its effect on dispersion stability. Prog. Org. Coat. 1973, 2, 81– 98, DOI: 10.1016/0300-9440(73)80001-315https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaE2cXktFKqtrs%253D&md5=0e7f651e29ac8a76fc64d605281265a4Zeta potential in nonaqueous media and its effect on dispersion stabilityKitahara, AyaoProgress in Organic Coatings (1973), 2 (2), 81-98CODEN: POGCAT; ISSN:0300-9440.A review is given with 38 refs. The properties and stability of org. coatings in relation to zeta potential are also discussed.
- 16Fowkes, F. Dispersions of ceramic powders in organic media. Adv. Ceram.: Ceram. Powder Sci. 1987, 21, 411– 421There is no corresponding record for this reference.
- 17Van Der Hoeven, P.; Lyklema, J. Electrostatic stabilization in non-aqueous media. Adv. Colloid Interface Sci. 1992, 42, 205– 277, DOI: 10.1016/0001-8686(92)80024-R17https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK38XmsFyksbo%253D&md5=5c29bb91e393773edc986c1ba9396877Electrostatic stabilization in non-aqueous mediaVan der Hoeven, Ph. C.; Lyklema, J.Advances in Colloid and Interface Science (1992), 42 (), 205-77CODEN: ACISB9; ISSN:0001-8686.A review on electrostatic stabilization of non-aq. dispersions with particular emphasis on the roles of dielec. const. and ionic strength. The liq. dielec. const. controls the degree of ionization of the stabilizing electrolyte and 3 regimes (semipolar, low-polar, apolar) can be distinguished. When attraction between suspended salt particles in nonionic liqs. is weak, electrostatic stabilization is achieved easily. 107 Refs.
- 18Siffert, B.; Jada, A.; Eleli-Letsango, J. Stability Calculations of TiO2 Nonaqueous Suspensions: Thickness of the Electrical Double Layer. J. Colloid Interface Sci. 1994, 167, 281– 286, DOI: 10.1006/jcis.1994.1362There is no corresponding record for this reference.
- 19Hughes, D. F. K.; Robb, I. D.; Dowding, P. J. Stability of Copper Phthalocyanine Dispersions in Organic Media. Langmuir 1999, 15, 5227– 5231, DOI: 10.1021/la981389v19https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK1MXkt1aqsrw%253D&md5=647f09165ba8da13704defab831405adStability of Copper Phthalocyanine Dispersions in Organic MediaHughes, D. F. K.; Robb, Ian D.; Dowding, Peter J.Langmuir (1999), 15 (16), 5227-5231CODEN: LANGD5; ISSN:0743-7463. (American Chemical Society)It was found that quite small surface potentials (a few millivolts) were sufficient to stabilize the dil. dispersions of the particles in solvents of low dielec. const. (possibly because this would be a const. charge system) though insufficient to stabilize particles in solvents with dielec. consts. above about 20. Addn. of org. acids enhanced the particles' stability, probably by a proton-exchange mechanism, and addn. of an org.-sol. electrolyte, tetrabutylammonium bromide, caused all systems to flocculate. The presence of a polymer at concns. above about 1% caused flocculation to occur, probably by a depletion mechanism. The attractive depletion energy increased with polymer concn., resulting in the sedimentation vol. increasing with polymer concn.
- 20Kosmulski, M. Zeta potentials in nonaqueous media: how to measure and control them. Colloids Surf., A 1999, 159, 277– 281, DOI: 10.1016/S0927-7757(99)00273-320https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK1MXnsFertbY%253D&md5=30b927e714a0130eb179580f2eb4616bZeta potentials in nonaqueous media. How to measure and control themKosmulski, M.Colloids and Surfaces, A: Physicochemical and Engineering Aspects (1999), 159 (2-3), 277-281CODEN: CPEAEH; ISSN:0927-7757. (Elsevier Science B.V.)A review, with 20 refs. The effect of traces of water in nonaq. solvents on the ζ-potential is overrated, and the role of impurities in presumably pure org. solvents is often overlooked. Oxides can serve as potentiometric sensors to det. the concn. and nature of impurities. Even in very inert solvents the counter-charge in soln. must exist. Small ions are stabilized by homoconjugation and by amphiphilic nonionic compds. The electrokinetic phenomena of the second kind are more common in nonaq. solvents and they lead to very high abs. values of mobility at high field strengths.
- 21Kosmulski, M.; Eriksson, P.; Brancewicz, C.; Rosenholm, J. B. Zeta potentials of monodispersed, spherical silica particles in mixed solvents as a function of cesium chloride concentration. Colloids Surf., A 2000, 162, 37– 48, DOI: 10.1016/S0927-7757(99)00027-821https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3cXhtFeltL4%253D&md5=de877254cdeaede02f8f4c7684abf0d1Zeta potentials of monodispersed, spherical silica particles in mixed solvents as a function of cesium chloride concentrationKosmulski, Marek; Eriksson, Patrik; Brancewicz, Chris; Rosenholm, Jarl B.Colloids and Surfaces, A: Physicochemical and Engineering Aspects (2000), 162 (1-3), 37-48CODEN: CPEAEH; ISSN:0927-7757. (Elsevier Science B.V.)Zeta potentials of silica are neg. in most mixed solvents (99% org. co-solvent, 1% water), but in t-butanol and pyridine they are pos. When CsCl is added to the suspension as a probe substance, two types of behavior are obsd. In solvents of ε >25, the zeta potential asymptotically approaches zero, while in solvents of lower dielec. const., the sign of zeta potential is reversed to pos. at a sufficiently high CsCl concn., crev. To evaluate the nature of interaction leading to different solvent responses, a range of frequently used solvent scales relating to specific interactions between the components of the system were compared. A moderate correlation was found between crev for silica and one solvent property alone, and those previously found for anatase, resp. Using linear combinations of 2 solvent scales improved the correlation, but it was still not satisfactory. However, considering sep. the contributions from both solvent components, a nearly statistically significant correlation was found. The results are evaluated considering the particular behavior of each solvent group, and the specific influence of the material properties of silica and of possible impurities present in the system.
- 22Zhukov, A. Integrated investigations of the electrosurface properties of nonaqueous disperse and capillary systems. Adv. Colloid Interface Sci. 2007, 134–135, 330– 345, DOI: 10.1016/j.cis.2007.04.014There is no corresponding record for this reference.
- 23Hamieh, T.; Toufaily, J.; Alloul, H. Physicochemical Properties of the Dispersion of Titanium Dioxide in Organic Media by Using Zetametry Technique. J. Dispers. Sci. Technol. 2008, 29, 1181– 1188, DOI: 10.1080/01932690701856626There is no corresponding record for this reference.
- 24Kosmulski, M.; Prochniak, P.; Rosenholm, J. B. Control of the Zeta Potential in Semiconcentrated Dispersions of Titania in Polar Organic Solvents. J. Phys. Chem. C 2009, 113, 12806– 12810, DOI: 10.1021/jp903845e24https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1MXnsVWgtbg%253D&md5=796d406d0c2347c33f35537538c4ce80Control of the Zeta Potential in Semiconcentrated Dispersions of Titania in Polar Organic SolventsKosmulski, Marek; Prochniak, Piotr; Rosenholm, Jarl B.Journal of Physical Chemistry C (2009), 113 (29), 12806-12810CODEN: JPCCCK; ISSN:1932-7447. (American Chemical Society)The electrokinetic potential of com. titania powder (chiefly anatase) in its 1-10% dispersions in water, lower aliph. alcs., and DMSO is pos., and rather insensitive to the nature of the solvent. The sign of the electrokinetic potential can be reversed to neg. by addn. of phosphoric acid, or of combination of phosphoric acid with triethylamine or with alkali hydroxide. The crit. concn. of surface-active compds., which induces a sign reversal, depends on the nature of the solvent, and it is higher in org. solvents than in water. The crit. surface concn. of surface-active compds. for given org. solvent is rather insensitive to the solid-to-liq. ratio, and typically it is in the range of a few micromoles per square meter.
- 25Cihlar, J.; Drdlik, D.; Cihlarova, Z.; Hadraba, H. Effect of acids and bases on electrophoretic deposition of alumina and zirconia particles in 2-propanol. J. Eur. Ceram. Soc. 2013, 33, 1885– 1892, DOI: 10.1016/j.jeurceramsoc.2013.02.017There is no corresponding record for this reference.
- 26Lyklema, J. Principles of interactions in non-aqueous electrolyte solutions. Curr. Opin. Colloid Interface Sci. 2013, 18, 116– 128, DOI: 10.1016/j.cocis.2013.02.00226https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3sXivFylsrY%253D&md5=296ea4b32938cacda3d79800931810b7Principles of interactions in non-aqueous electrolyte solutionsLyklema, JohannesCurrent Opinion in Colloid & Interface Science (2013), 18 (2), 116-128CODEN: COCSFL; ISSN:1359-0294. (Elsevier Ltd.)In this paper a review is presented on the mol. interactions in non-aq. media of low dielec. permittivity. Qual. and quant. distinctions with aq. solns. are emphasized. The reviewed themes include dispersion forces, dissocn. and assocn. equil., discrimination between electrostatic and non-electrostatic interactions, ionic specificity, cond., electrokinetics and colloid interaction. Distinctions between so-called primitive and non-primitive interpretations and between individual and collective behavior are discussed; in these respects the colloid stability phenomena behave differently from the corresponding ones in aq. solvents.
- 27Mazzini, V.; Craig, V. S. Specific-ion effects in non-aqueous systems. Curr. Opin. Colloid Interface Sci. 2016, 23, 82– 93, DOI: 10.1016/j.cocis.2016.06.00927https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC28XhtFCitbfK&md5=2aacad84502b94640a7f3075129d85ecSpecific-ion effects in non-aqueous systemsMazzini, Virginia; Craig, Vincent S. J.Current Opinion in Colloid & Interface Science (2016), 23 (), 82-93CODEN: COCSFL; ISSN:1359-0294. (Elsevier Ltd.)A review. It is widely acknowledged that specific-ion effects are ubiquitous in aq. systems and undoubtedly are essential to the fundamental processes of life, although a deep fundamental understanding of specific-ion effects remains an important challenge. Specific-ion effects in non-aq. solvents are known but have attracted far less attention, yet knowledge of specific-ion effects in non-aq. systems is likely to provide important information for guiding, evaluating and testing our theories of specific-ion effects. Here, the literature on specific-ion effects in non-aq. solvents is surveyed with a view to detg. if the Hofmeister series or lyotropic series so universally obsd. in aq. systems is widely evident in non-aq. systems. Particular attention has been applied to expts. on non-aq. systems that are known to exhibit Hofmeister series in aq. systems with the aim of detg. if a consistent ion ordering in the strength of specific-ion effects is obsd. in other solvents. We find that specific-ion effects are ubiquitous in non-aq. solvents, that both Hofmeister and lyotropic series are widely obsd., although not necessarily for the same class of expt. Moreover, we find that Hofmeister and lyotropic series are obsd. in non-aq. solvents even for expts. in which these series are not obsd. for water. Addnl., series reversal is seen for a given expt. when the solvent is changed. All this poses significant challenges for our understanding of specific-ion effects in aq. and non-aq. systems and also provides guideposts for future investigations.
- 28Rosenholm, J. B. Evaluation of particle charging in non-aqueous suspensions. Adv. Colloid Interface Sci. 2018, 259, 21– 43, DOI: 10.1016/j.cis.2018.06.00428https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1cXhtlWht7vN&md5=e3f9922bedc0c19a5d2d175e4a7b98b4Evaluation of particle charging in non-aqueous suspensionsRosenholm, Jarl B.Advances in Colloid and Interface Science (2018), 259 (), 21-43CODEN: ACISB9; ISSN:0001-8686. (Elsevier B.V.)Factors influencing the sign and size of effective surface (zeta) potential in suspensions of very low dielec. consts. are evaluated. For non-aq. suspensions it was found that Gutmann's donor no. (DN = neg. Lewis type molar acid-base adduct formation enthalpy) was successfully related to zeta potential changes, similarly as pH is optimal for aq. suspensions. Neg. molar proton dissocn. enthalpy (Br.vphi.nsted type HD no.), neg. hydrogen bond enthalpy (HB no.), logarithmic hydrogen bond equil. const. (molar Gibbs free energy), std. redn. potential of solvated protons (Eo(H+L/H2)), electrolytic dissocn. potential of water (Eo(H2O/H2, O2)) and electron exchange Fermi potentials could equally well be related to zeta potential changes. All these properties were linearly dependent on each other. When evaluating factors contributing to attractive and repulsive interaction energies, it is found that in order for the models to be relevant the extension of diffuse charging has to be much larger than the distance to repulsive barrier ensuring suspension stability. At this limit and at high surface potentials, the repulsive energy grows exceptionally large being in the range of lattice energy of each solid. The models fail when surface potential is low and the extension of diffuse charging is much smaller than the distance to repulsive barrier.
- 29Barthel, J.; Gores, H.-J.; Schmeer, G.; Wachter, R. Topics in Current Chemistry; Springer: Berlin Heidelberg, 1983; pp 33– 144.There is no corresponding record for this reference.
- 30Izutsu, K. Electrochemistry in Nonaqueous Solutions; Wiley-VCH Verlag GmbH & Co. KGaA, 2002.There is no corresponding record for this reference.
- 31Iwasaki, H.; Kimura, Y.; Uematsu, Y. Ubiquitous Preferential Water Adsorption to Electrodes in Water/1-Propanol Mixtures Detected by Electrochemical Impedance Spectroscopy. J. Phys. Chem. C 2023, 127, 23382– 23389, DOI: 10.1021/acs.jpcc.3c05320There is no corresponding record for this reference.
- 32Iwai, S.; Suzuki, T.; Sakagami, H.; Miyamoto, K.; Chen, Z.; Konishi, M.; Villani, E.; Shida, N.; Tomita, I.; Inagi, S. Electropolymerization without an electric power supply. Commun. Chem. 2022, 5, 66, DOI: 10.1038/s42004-022-00682-832https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB38XhsVSntrjK&md5=6ad1371f6ec9d67af6a5f557ae48af0bElectropolymerization without an electric power supplyIwai, Suguru; Suzuki, Taichi; Sakagami, Hiroki; Miyamoto, Kazuhiro; Chen, Zhenghao; Konishi, Mariko; Villani, Elena; Shida, Naoki; Tomita, Ikuyoshi; Inagi, ShinsukeCommunications Chemistry (2022), 5 (1), 66CODEN: CCOHCT; ISSN:2399-3669. (Nature Portfolio)Abstr.: Electrifying synthesis is now a common slogan among synthetic chemists. In addn. to the conventional two- or three-electrode systems that use batch-type cells, recent progress in org. electrochem. processes has been significant, including microflow electrochem. reactors, Li-ion battery-like technol., and bipolar electrochem. Herein we demonstrate an advanced electrosynthesis method without the application of elec. power based on the concept of streaming potential-driven bipolar electrochem. As a proof-of-concept study, the electrochem. oxidative polymn. of arom. monomers successfully yielded the corresponding polymer films on an electrode surface, which acted as an anode under the flow of electrolyte in a microchannel without an elec. power supply.
- 33Kolthoff, I. M. Acid-base equilibriums in dipolar aprotic solvents. Anal. Chem. 1974, 46, 1992– 2003, DOI: 10.1021/ac60349a00533https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaE2cXlsVWmsbs%253D&md5=a5ecea013728d8730d546e6cdbbd26e4Acid-base equilibriums in dipolar aprotic solventsKolthoff, I. M.Analytical Chemistry (1974), 46 (13), 1992-2003CODEN: ANCHAM; ISSN:0003-2700.A classification of org. solvents is proposed. The main emphasis is on the effect of homo- and heteroconjugate formation by H bonding on acid-base equilibria and on conductometric and potentiometric titrn. curves in dipolar aprotic solvents. The relation is discussed between resolution of acid strength and transfer activity coeffs. of ions and mols. A brief review is presented of analytical uses of protophobic solvents for the titrn. of very weak bases and of protophilic solvents for the titrn. of very weak acids. 77 Refs.
- 34Rossini, E.; Knapp, E. Proton solvation in protic and aprotic solvents. J. Comput. Chem. 2016, 37, 1082– 1091, DOI: 10.1002/jcc.2429734https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC28Xht1eksrk%253D&md5=c298ddbc68b3966d41021abebe4fc775Proton solvation in protic and aprotic solventsRossini, Emanuele; Knapp, Ernst-WalterJournal of Computational Chemistry (2016), 37 (12), 1082-1091CODEN: JCCHDD; ISSN:0192-8651. (John Wiley & Sons, Inc.)Protonation pattern strongly affects the properties of mol. systems. To det. protonation equil., proton solvation free energy, which is a central quantity in soln. chem., needs to be known. In this study, proton affinities (PAs), electrostatic energies of solvation, and pKA values were computed in protic and aprotic solvents. The proton solvation energy in acetonitrile (MeCN), methanol, water, and DMSO was detd. from computed and measured pKA values for a specially selected set of org. compds. pKA values were computed with high accuracy using a combination of quantum chem. and electrostatic approaches. Quantum chem. d. functional theory computations were performed evaluating PA in the gas-phase. The electrostatic contributions of solvation were computed solving the Poisson equation. The computations yield proton solvation free energies with high accuracy, which are in MeCN, MeOH, water, and DMSO -255.1, -265.9, -266.3, and -266.4 kcal/mol, resp., where the value for water is close to the consensus value of -265.9 kcal/mol. The pKA values of MeCN, MeOH, and DMSO in water correlates well with the corresponding proton solvation energies in these liqs., indicating that the solvated proton was attached to a single solvent mol. © 2016 Wiley Periodicals, Inc.
- 35Rossini, E.; Bochevarov, A. D.; Knapp, E. W. Empirical Conversion of pKa Values between Different Solvents and Interpretation of the Parameters: Application to Water, Acetonitrile, Dimethyl Sulfoxide, and Methanol. ACS Omega 2018, 3, 1653– 1662, DOI: 10.1021/acsomega.7b0189535https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1cXitl2lsbY%253D&md5=0fbd9b505269e3e46d02f1c3eb014eceEmpirical Conversion of pKa Values between Different Solvents and Interpretation of the Parameters: Application to Water, Acetonitrile, Dimethyl Sulfoxide, and MethanolRossini, Emanuele; Bochevarov, Art D.; Knapp, Ernst WalterACS Omega (2018), 3 (2), 1653-1662CODEN: ACSODF; ISSN:2470-1343. (American Chemical Society)An empirical conversion method (ECM) that transforms pKa values of arbitrary org. compds. from one solvent to the other is introduced. We demonstrate the method's usefulness and performance on pKa conversions involving water and org. solvents acetonitrile (MeCN), DMSO (Me2SO), and methanol (MeOH). We focus on the pKa conversion from the known ref. value in water to the other three org. solvents, although such a conversion can also be performed between any pair of the considered solvents. The ECM works with an additive parameter that is specific to a solvent and a mol. family (essentially characterized by a functional group that is titrated). We formally show that the method can be formulated with a single additive parameter, and that the extra multiplicative parameter used in other works is not required. The values of the additive parameter are detd. from known pKa data, and their interpretation is provided on the basis of physicochem. concepts. The data set of known pKa values is augmented with pKa values computed with the recently introduced electrostatic transform method, whose validity is demonstrated. For a validation of our method, we consider pKa conversions for two data sets of titratable compds. The first data set involves 81 relatively small mols. belonging to 19 different mol. families, with the pKa data available in all four considered solvents. The second data set involves 76 titratable mols. from 5 addnl. mol. families. These mols. are typically larger, and their exptl. pKa values are available only in Me2SO and water. The validation tests show that the agreement between the exptl. pKa data and the ECM predictions is generally good, with abs. errors often on the order of 0.5 pH units. The presence of a few outliers is rationalized, and obsd. trends with respect to mol. families are discussed.
- 36Borissova, M.; Gorbatšova, J.; Ebber, A.; Kaljurand, M.; Koel, M.; Vaher, M. Nonaqueous CE using contactless conductivity detection and ionic liquids as BGEs in ACN. Electrophoresis 2007, 28, 3600– 3605, DOI: 10.1002/elps.200700067There is no corresponding record for this reference.
- 37Gouy, M. Sur la constitution de la charge électrique à la surface d’un électrolyte. J. Phys. Theor. Appl. 1910, 9, 457– 468, DOI: 10.1051/jphystap:01910009004570037https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaC3cXht1eluw%253D%253D&md5=d4237e8698c2acf4fcd7c80c970101ecConstitution of the Electric Charge at the Surface of an ElectrolyteGouy(1910), 9 (), 457-67 ISSN:.An amplification of an article previously abstracted (C. A., 4, 698).
- 38Chapman, D. L. A contribution to the theory of electrocapillarity. London Edinb. Philos. Mag. 1913, 25, 475– 481, DOI: 10.1080/14786440408634187There is no corresponding record for this reference.
- 39Langmuir, I. THE ADSORPTION OF GASES ON PLANE SURFACES OF GLASS, MICA AND PLATINUM. J. Am. Chem. Soc. 1918, 40, 1361– 1403, DOI: 10.1021/ja02242a00439https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaC1cXht1KgsA%253D%253D&md5=3c22305e74ddd44d690168373bdd13c9The adsorption of gases on plane surfaces of glass, mica and platinumLangmuir, I.Journal of the American Chemical Society (1918), 40 (), 1361-1402CODEN: JACSAT; ISSN:0002-7863.According to L.'s hypothesis, gaseous mols. impinging on a liquid or solid surface do not in general rebound from it elastically, but are held or adsorbed on the surface by forces similar to those holding the atoms or group mols. of solid bodies. The adsorbed film should not exceed one mol. in thickness. Adsorption of permanent gases involves only secondary valence forces. In metals particularly, adsorption may be governed by primary valence forces. It is suggested that stoichiometric relations should govern the adsorption on a surface unless interfering effects caused by steric hindrance are involved. At room temp. the absorption by glass and mica was negligible, not over 1 % of the surface being covered by a single layer of mols. At lower temps. much larger quantities of gas were taken up. With Pt no absorption was observed at - 183° unless the Pt were first activated by proper heating. The adsorption of O2 was irreversible and corresponded to a monomolecular layer. CO likewise showed the same behavior. In the presence of one or the other gas adsorbed on the Pt the adsorbed and unadsorbed gases reacted immediately to form CO2.
- 40Rice, C. L.; Whitehead, R. Electrokinetic Flow in a Narrow Cylindrical Capillary. J. Phys. Chem. 1965, 69, 4017– 4024, DOI: 10.1021/j100895a06240https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaF28XlsFOktg%253D%253D&md5=3d4dd3365bd0d640395703e7436fc5adElectrokinetic flow in a narrow cylindrical capillaryRice, C. L.; Whitehead, R.Journal of Physical Chemistry (1965), 69 (11), 4017-24CODEN: JPCHAX; ISSN:0022-3654.An anal. study was made of electrokinetic flow in cylindrical capillaries with radii on the order of 10-5 cm. The work extends that of Burgreen and Nakache (CA 61, 15683b). Expressions are given for electroosmosis, streaming potential, and c.d. distribution. The results lead to a prediction of a max. in the electroviscous effect.
- 41Goldfarb, D. L.; Longinotti, M. P.; Corti, H. R. Electrical Conductances of Tetrabutylammonium and Decamethylferrocenium Hexafluorophosphate in Organic Solvents. J. Solution Chem. 2001, 30, 307– 322, DOI: 10.1023/A:1010334021934There is no corresponding record for this reference.
- 42Hiemstra, T.; De Wit, J.; Van Riemsdijk, W. Multisite proton adsorption modeling at the solid/solution interface of (hydr)oxides: A new approach. J. Colloid Interface Sci. 1989, 133, 105– 117, DOI: 10.1016/0021-9797(89)90285-342https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK3cXht1Cnuw%253D%253D&md5=3a2420d04d20dd5e805a7835084607b0Multisite proton adsorption modeling at the solid/solution interface of (hydr)oxides: a new approach. II. Application to various important (hydr)oxidesHiemstra, T.; De Wit, J. C. M.; Van Riemsdijk, W. H.Journal of Colloid and Interface Science (1989), 133 (1), 105-17CODEN: JCISA5; ISSN:0021-9797.At the solid/soln. interface of metal (hydro)oxides various types of O(H) and OH(H) groups are present, which differ in th no. of coordinating metal ions. The σ0-pH curves of metal (hydr)oxides are strongly detd. by the compn. and the relative extent of the various crystal planes of (hydr)oxides. The charging behavior is discuss for gibbsite (Al(OH)3), goethite (FeOOH), hematite (Fe2O3), rutile (TiO2), and silica (SiO2. New exptl. σ0-pH data for goethite and gibbsite are presented. Several important (hydr)oxides exhibit crystal faces which do not develop surface charge over a relatively wide pH range. An uncharged crystal face may be due to the presence of surface groups which are not reactive (inert) in the pH range under consideration, like the 001 face of gibbsite ad the 0001 face of hematite, or caused by the presence of two types of interacting charged surface groups of which the charge of one type is fully compensated by the other like at the 100 face of goethite. The charging behavior of silica and the 001 face of gibbsite is detd. by one type of reactive surface group with a large ΔpK for the consecutive protonation steps. The crystal structure imposes the presence of uncharged surface groups and this results in a quite different shape of σ0-pH curves for gibbsite and silica in comparison with the commonly obsd. σ0-pH curves for gibbsite and silica in comparison with the commonly obsd. σ0-pH curves of metal (hydr)oxides. The MUltiSIte complexation (MUSIC) model as developed by T. Hiemstra, et al. (1989) leads to a rather good prediction of σ0-pH curves for various metal (hydr)oxides using predicted affinity consts. for the various types of surface groups and Stern layer capacitance values and pair formation consts. estd. from the literature.
- 43Behrens, S. H.; Grier, D. G. The charge of glass and silica surfaces. J. Chem. Phys. 2001, 115, 6716– 6721, DOI: 10.1063/1.140498843https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3MXnt1ams7o%253D&md5=c6018787255a16a50149e21cda2f3ca0The charge of glass and silica surfacesBehrens, Sven H.; Grier, David G.Journal of Chemical Physics (2001), 115 (14), 6716-6721CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)We present a method of calcg. the elec. charge d. of glass and silica surfaces in contact with aq. electrolytes for two cases of practical relevance that are not amenable to std. techniques: surfaces of low specific area at low ionic strength and surfaces interacting strongly with a second anionic surface.
- 44Bialik, E.; Stenqvist, B.; Fang, Y.; Östlund, Å.; Furó, I.; Lindman, B.; Lund, M.; Bernin, D. Ionization of Cellobiose in Aqueous Alkali and the Mechanism of Cellulose Dissolution. J. Phys. Chem. Lett. 2016, 7, 5044– 5048, DOI: 10.1021/acs.jpclett.6b0234644https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC28XhvFSitrrL&md5=e6804dccfdcd71184152103737965cdeIonization of Cellobiose in Aqueous Alkali and the Mechanism of Cellulose DissolutionBialik, Erik; Stenqvist, Bjoern; Fang, Yuan; Oestlund, Aasa; Furo, Istvan; Lindman, Bjoern; Lund, Mikael; Bernin, DianaJournal of Physical Chemistry Letters (2016), 7 (24), 5044-5048CODEN: JPCLCD; ISSN:1948-7185. (American Chemical Society)Cellulose, one of the most abundant renewable resources, is insol. in most common solvents but dissolves in aq. alkali under a narrow range of conditions. To elucidate the solubilization mechanism, we performed electrophoretic NMR on cellobiose, a subunit of cellulose, showing that cellobiose acts as an acid with two dissocn. steps at pH 12 and 13.5. Chem. shift differences between cellobiose in NaOH and NaCl were estd. using 2D NMR and compared to DFT shift differences upon deprotonation. The dissocn. steps are the deprotonation of the hemiacetal OH group and the deprotonation of one of four OH groups on the nonreducing anhydroglucose unit. MD simulations reveal that aggregation is suppressed upon charging cellulose chains in soln. Our findings strongly suggest that cellulose is to a large extent charged in concd. aq. alkali, a seemingly crucial factor for solubilization. This insight, overlooked in the current literature, is important for understanding cellulose dissoln. and for synthesis of new sustainable materials.
- 45Malerz, S.; Mudryk, K.; Tomaník, L.; Stemer, D.; Hergenhahn, U.; Buttersack, T.; Trinter, F.; Seidel, R.; Quevedo, W.; Goy, C.; Wilkinson, I.; Thürmer, S.; Slavíček, P.; Winter, B. Following in Emil Fischer’s Footsteps: A Site-Selective Probe of Glucose Acid–Base Chemistry. J. Phys. Chem. A 2021, 125, 6881– 6892, DOI: 10.1021/acs.jpca.1c0469545https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB3MXhs1KgtrnF&md5=02664b9d409bace291887eed06258865Following in Emil Fischer's Footsteps: A Site-Selective Probe of Glucose Acid-Base ChemistryMalerz, Sebastian; Mudryk, Karen; Tomanik, Lukas; Stemer, Dominik; Hergenhahn, Uwe; Buttersack, Tillmann; Trinter, Florian; Seidel, Robert; Quevedo, Wilson; Goy, Claudia; Wilkinson, Iain; Thuermer, Stephan; Slavicek, Petr; Winter, BerndJournal of Physical Chemistry A (2021), 125 (32), 6881-6892CODEN: JPCAFH; ISSN:1089-5639. (American Chemical Society)Liq.-jet photoelectron spectroscopy was applied to det. the first acid dissocn. const. (pKa) of aq.-phase glucose while simultaneously identifying the spectroscopic signature of the resp. deprotonation site. Valence spectra from solns. at pH values below and above the first pKa reveal a change in glucose's lowest ionization energy upon the deprotonation of neutral glucose and the subsequent emergence of its anionic counterpart. Site-specific insights into the soln.-pH-dependent mol. structure changes are also shown to be accessible via C 1s photoelectron spectroscopy. The spectra reveal a considerably lower C 1s binding energy of the carbon site assocd. with the deprotonated hydroxyl group. The occurrence of photoelectron spectral fingerprints of cyclic and linear glucose prior to and upon deprotonation are also discussed. The exptl. data are interpreted with the aid of electronic structure calcns. Our findings highlight the potential of liq.-jet photoelectron spectroscopy to act as a site-selective probe of the mol. structures that underpin the acid-base chem. of polyprotic systems with relevance to environmental chem. and biochem.
- 46Uematsu, Y.; Bonthuis, D. J.; Netz, R. R. Charged Surface-Active Impurities at Nanomolar Concentration Induce Jones–Ray Effect. J. Phys. Chem. Lett. 2018, 9, 189– 193, DOI: 10.1021/acs.jpclett.7b0296046https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2sXitVSnurbK&md5=ff9329b83b3083f688fa5cc3ee9bd1e8Charged Surface-Active Impurities at Nanomolar Concentration Induce Jones-Ray EffectUematsu, Yuki; Bonthuis, Douwe Jan; Netz, Roland R.Journal of Physical Chemistry Letters (2018), 9 (1), 189-193CODEN: JPCLCD; ISSN:1948-7185. (American Chemical Society)The electrolyte surface tension exhibits a characteristic min. around a salt concn. of 1 mM for all ion types, known as the Jones-Ray effect. A consistent description of the exptl. surface tension of salts, bases, and acids is possible by assuming charged impurities in the water with a surface affinity typical for surfactants. Comparison with exptl. data yields an impurity concn. in the nanomolar range, well below the typical exptl. detection limit. Modeling reveals salt-screening enhanced impurity adsorption as the mechanism behind the Jones-Ray effect: for very low salt concn. added salt screens the electrostatic repulsion between impurities at the surface, which dramatically increases impurity adsorption and thereby reduces the surface tension.
- 47Plett, T.; Shi, W.; Zeng, Y.; Mann, W.; Vlassiouk, I.; Baker, L. A.; Siwy, Z. S. Rectification of nanopores in aprotic solvents–transport properties of nanopores with surface dipoles. Nanoscale 2015, 7, 19080– 19091, DOI: 10.1039/C5NR06340JThere is no corresponding record for this reference.
- 48Nasir, S.; Ali, M.; Ramirez, P.; Froehlich, K.; Cervera, J.; Mafe, S.; Ensinger, W. Ionic conduction through single-pore and multipore polymer membranes in aprotic organic electrolytes. J. Membr. Sci. 2021, 635, 119505 DOI: 10.1016/j.memsci.2021.11950548https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB3MXhsVSit73I&md5=46ab0d4fee27da11ca0ddc3dfdd06f01Ionic conduction through single-pore and multipore polymer membranes in aprotic organic electrolytesNasir, Saima; Ali, Mubarak; Ramirez, Patricio; Froehlich, Kristina; Cervera, Javier; Mafe, Salvador; Ensinger, WolfgangJournal of Membrane Science (2021), 635 (), 119505CODEN: JMESDO; ISSN:0376-7388. (Elsevier B.V.)We exptl. characterize the ionic conduction of single and multipore nanoporous membranes in aprotic org. electrolytes. To this end, soft-etched (SE) membranes with pore diams. in the nanometer range and track-etched (TE) membranes with pore diams. in the tens of nanometers range are investigated. In aq. conditions, the membrane ionic conduction rates follow the same trend of the bulk soln. conductivities. However, the ionic transport through the narrow SE-nanopores dramatically decreases in aprotic electrolytes due to the formation of solvated metal cations and their adsorption on the pore surface. The current-voltage recordings of single conical nanopores in aprotic electrolyte solns. with different water mole fractions reveal that the solvated metal ion (M) species [M-(solvent)4]+ formed in acetonitrile solvent are more tightly bounded to the pore walls compared with the cationic chelates obtained in propylene carbonate solvent. The basic findings reported here should be of interest for ionic/mol. nanofiltration processes in non-aq. conditions as well as for moisture sensitive and energy storage nanofluidic devices.
- 49Polster, J. W.; Souna, A. J.; Motevaselian, M. H.; Lucas, R. A.; Tran, J. D.; Siwy, Z. S.; Aluru, N. R.; Fourkas, J. T. The electrical-double layer revisited. Nat. Sci. 2022, 2, e20210099 DOI: 10.1002/ntls.20210099There is no corresponding record for this reference.
- 50Souna, A. J.; Motevaselian, M. H.; Polster, J. W.; Tran, J. D.; Siwy, Z. S.; Aluru, N. R.; Fourkas, J. T. Beyond the electrical double layer model: ion-dependent effects in nanoscale solvent organization. Phys. Chem. Chem. Phys. 2024, 26, 6726– 6735, DOI: 10.1039/D3CP05712GThere is no corresponding record for this reference.
- 51Yin, X.; Zhang, S.; Dong, Y.; Liu, S.; Gu, J.; Chen, Y.; Zhang, X.; Zhang, X.; Shao, Y. Ionic Current Rectification in Organic Solutions with Quartz Nanopipettes. Anal. Chem. 2015, 87, 9070– 9077, DOI: 10.1021/acs.analchem.5b0233751https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2MXht1CitL7K&md5=3fe8404e459b8844d48d68429e562fb5Ionic Current Rectification in Organic Solutions with Quartz NanopipettesYin, Xiaohong; Zhang, Shudong; Dong, Yitong; Liu, Shujuan; Gu, Jing; Chen, Ye; Zhang, Xin; Zhang, Xianhao; Shao, YuanhuaAnalytical Chemistry (Washington, DC, United States) (2015), 87 (17), 9070-9077CODEN: ANCHAM; ISSN:0003-2700. (American Chemical Society)The study of behaviors of ionic current rectification (ICR) in org. solns. with quartz nanopipettes is reported. ICR can be obsd. even in org. solns. using quartz pipets with diams. varied from several to dozens of nanometers, and the direction of ICR is quite different from the ICR obsd. in aq. phase. The influences of pore size, electrolyte concn., and surface charge on the ICR were studied carefully. Water in org. solns. affects the direction and extent of ICR significantly. Mechanisms about the formation of an elec. double layer (EDL) on silica in org. solns. with different amt. of water are proposed. An improved method, which can be employed to detect trace water in org. solns., was implemented based on Au ultramicroelectrodes with cathodic differential pulse stripping voltammetry.
- 52Spångberg, D.; Hermansson, K. The solvation of Li+ and Na+ in acetonitrile from ab initio-derived many-body ion–solvent potentials. Chem. Phys. 2004, 300, 165– 176, DOI: 10.1016/j.chemphys.2004.01.01152https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2cXisl2lsb4%253D&md5=8962ba45e76fb9c036f074ac17334b03The solvation of Li+ and Na+ in acetonitrile from ab initio-derived many-body ion-solvent potentialsSpangberg, Daniel; Hermansson, KerstiChemical Physics (2004), 300 (1-3), 165-176CODEN: CMPHC2; ISSN:0301-0104. (Elsevier Science B.V.)Several Li+- and Na+-acetonitrile models were derived from ab initio calcns. at the counterpoise-cor. MP2/TZV++(d,p) level for distorted ion-(MeCN)n clusters with n=1, 4 and 6. Two different many-body ion-acetonitrile models were constructed: an effective three-body potential for use with the six-site effective pair model of Bohm et al., and an effective polarizable many-body model. The polarizable acetonitrile model used in the latter model is a new empirical model which was also derived in the present paper. Mainly for comparative purposes, two ion-acetonitrile pair potentials were also constructed from the ab initio cluster calcns.: one pure pair potential and one effective pair potential. Using all these potential models, MD simulations in the NPT ensemble were performed for the pure acetonitrile liq. and for Li+(MeCN) and Na+(MeCN) solns. with 1 ion in 512 solvent mols. and with a simulation time of at least 120 ps per system. Thermodn. properties, solvation-shell structure and the self-diffusion coeff. of the ions and of the solvent mols. were calcd. and compared between the different models and with exptl. data, where available. The Li+ ion is found to be four-coordinated when the new many-body potentials are used, in contrast to the six-coordinated structure obtained for the pure pair and effective pair potentials. The coordination no. of Na+ is close to six for all the models derived here, although the coordination no. becomes slightly smaller with the many-body potentials. For both ions, the solvent mols. in the first shell point their nitrogen ends towards the cation, while in the second shell the opposite orientation is the most common.
- 53Amara, S.; Toulc’Hoat, J.; Timperman, L.; Biller, A.; Galiano, H.; Marcel, C.; Ledigabel, M.; Anouti, M. Comparative Study of Alkali-Cation-Based (Li+, Na+, K+) Electrolytes in Acetonitrile and Alkylcarbonates. ChemPhysChem 2019, 20, 581– 594, DOI: 10.1002/cphc.20180106453https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1MXisVGnsrk%253D&md5=783dc126fb8ffa9bc3806c92ab647ad8Comparative Study of Alkali-Cation-Based (Li+, Na+, K+) Electrolytes in Acetonitrile and AlkylcarbonatesAmara, Samia; Toulc'Hoat, Joel; Timperman, Laure; Biller, Agnes; Galiano, Herve; Marcel, Corinne; Ledigabel, Matthieu; Anouti, MeriemChemPhysChem (2019), 20 (4), 581-594CODEN: CPCHFT; ISSN:1439-4235. (Wiley-VCH Verlag GmbH & Co. KGaA)The development of a suitable functional electrolyte is urgently required for fast-charging and high-voltage alkali-ion (Li, Na, K) batteries as well as next-generation hybrids supercapacitors. Many recent works focused on an optimal selection of electrolytes for alkali-ion based systems and their electrochem. performance but the understanding of the fundamental aspect that explains their different behavior is rare. Herein, we report a comparative study of transport properties for LiPF6, NaPF6, KPF6 in acetonitrile (AN) and a binary mixt. of ethylene carbonate (EC), di-Me carbonate (DMC): (EC/DMC : 1/1, weigh) through conductivities, densities and viscosities measurements in wide temp. domain. By application of the Stokes-Einstein, Nernst-Einstein, and Jones Dole equations, the effective ionic solvated radius of cation (reff), the ionic dissocn. coeff. (αD) and structuring Jones Dole's parameters (A, B) for salt are calcd. and discussed according to solvent or cation nature as a function of temp. From the results, we demonstrate that better mobility of potassium can be explained by the nature of the ion-ion and ion-solvent interactions due to its polarizability. In the same time, the predominance of triple ions in the case of K+, is a disadvantage at high concn.
- 54Ding, F.; Hu, Z.; Zhong, Q.; Manfred, K.; Gattass, R. R.; Brindza, M. R.; Fourkas, J. T.; Walker, R. A.; Weeks, J. D. Interfacial Organization of Acetonitrile: Simulation and Experiment. J. Phys. Chem. C 2010, 114, 17651– 17659, DOI: 10.1021/jp104597z54https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3cXhtVGrsbvN&md5=8bcd247fc15255fce039a48a613a595fInterfacial Organization of Acetonitrile: Simulation and ExperimentDing, Feng; Hu, Zhonghan; Zhong, Qin; Manfred, Katherine; Gattass, Rafael R.; Brindza, Michael R.; Fourkas, John T.; Walker, Robert A.; Weeks, John D.Journal of Physical Chemistry C (2010), 114 (41), 17651-17659CODEN: JPCCCK; ISSN:1932-7447. (American Chemical Society)Mol. dynamics simulations and vibrational sum frequency generation (VSFG) expts. in the methyl-stretching spectral region have been used to study acetonitrile at the silica/liq., silica/vapor, and liq./vapor interfaces. Our simulations show that, at the silica/liq. interface, acetonitrile takes on a considerably different structure than in the bulk liq. The interfacial structure is reminiscent of a lipid bilayer, and this type of ordering persists for tens of Ångstroms into the bulk liq. This result has important implications for processes involving solid/acetonitrile interfaces, such as heterogeneous catalysis and chromatog. sepns. Fitting of VSFG data that have an extremely low nonresonant background contribution provides strong evidence for interfacial populations pointing in opposite directions at these interfaces, in agreement with our simulations. The picture developed from our simulations and expts. reconciles conflicting interpretations of data from previous exptl. studies of interfacial acetonitrile.
- 55Pusić, T.; Grancarić, A. M.; Soljaçić, I.; Ribitsch, V. The effect of mercerisation on the electrokinetic potential of cotton. Color. Technol. 1999, 115, 121– 124, DOI: 10.1111/j.1478-4408.1999.tb00308.xThere is no corresponding record for this reference.
- 56Sadeghi-Kiakhani, M.; Safapour, S. Salt-free reactive dyeing of the cotton fabric modified with chitosan-poly(propylene imine) dendrimer hybrid. Fibers and Polym. 2015, 16, 1075– 1081, DOI: 10.1007/s12221-015-1075-9There is no corresponding record for this reference.
- 57Grancarić, A. M.; Tarbuk, A.; Hadžić, S.; Simončič, B. From Raw to Finished Cotton─Characterization by Interface Phenomena. Autex Res. J. 2023, 23, 184– 192, DOI: 10.2478/aut-2021-0055There is no corresponding record for this reference.
- 58Uematsu, Y.; Bonthuis, D. J.; Netz, R. R. Nanomolar Surface-Active Charged Impurities Account for the Zeta Potential of Hydrophobic Surfaces. Langmuir 2020, 36, 3645– 3658, DOI: 10.1021/acs.langmuir.9b0379558https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB3cXkvFOqu7Y%253D&md5=59e6893c6592b69f5996059ab9fc5c90Nanomolar Surface-Active Charged Impurities Account for the Zeta Potential of Hydrophobic SurfacesUematsu, Yuki; Bonthuis, Douwe Jan; Netz, Roland R.Langmuir (2020), 36 (13), 3645-3658CODEN: LANGD5; ISSN:0743-7463. (American Chemical Society)The electrification of hydrophobic surfaces is an intensely debated subject in phys. chem. We theor. study the ζ potential of hydrophobic surfaces for varying pH and salt concn. by solving the Poisson-Boltzmann and Stokes equations with individual ionic adsorption affinities. Using the ionic surface affinities extd. from the exptl. measured surface tension of the air-electrolyte interface, we first show that the interfacial adsorption and repulsion of small inorg. ions such as H3O+, OH-, HCO3-, and CO32- cannot account for the ζ potential obsd. in expts. because the surface affinities of these ions are too small. Even if we take hydrodynamic slip into account, the characteristic dependence of the ζ potential on pH and salt concn. cannot be reproduced. Instead, to explain the sizable exptl. measured ζ potential of hydrophobic surfaces, we assume minute amts. of impurities in the water and include the impurities' acidic and basic reactions with water. We find good agreement between our predictions and the reported exptl. ζ potential data of various hydrophobic surfaces if we account for impurities that consist of a mixt. of weak acids (pKa = 5-7) and weak bases (pKb = 12) at a concn. of the order of 10-7 M.
- 59Sghaier, S.; Zbidi, F.; Zidi, M. Characterization of Doum Palm Fibers After Chemical Treatment. Text. Res. J. 2009, 79, 1108– 1114, DOI: 10.1177/0040517508101623There is no corresponding record for this reference.
- 60Brož, Z.; Epstein, N. Electrokinetic flow through porous media composed of fine cylindrical capillaries. J. Colloid Interface Sci. 1976, 56, 605– 612, DOI: 10.1016/0021-9797(76)90127-2There is no corresponding record for this reference.
- 61Levine, S.; Marriott, J.; Neale, G.; Epstein, N. Theory of electrokinetic flow in fine cylindrical capillaries at high zeta-potentials. J. Colloid Interface Sci. 1975, 52, 136– 149, DOI: 10.1016/0021-9797(75)90310-0There is no corresponding record for this reference.
- 62Uematsu, Y.; Netz, R. R.; Bocquet, L.; Bonthuis, D. J. Crossover of the Power-Law Exponent for Carbon Nanotube Conductivity as a Function of Salinity. J. Phys. Chem. B 2018, 122, 2992– 2997, DOI: 10.1021/acs.jpcb.8b0197562https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1cXjs1Sqt7Y%253D&md5=32f80132c17eb669b616c0478bd337acCrossover of the Power-Law Exponent for Carbon Nanotube Conductivity as a Function of SalinityUematsu, Yuki; Netz, Roland R.; Bocquet, Lyderic; Bonthuis, Douwe JanJournal of Physical Chemistry B (2018), 122 (11), 2992-2997CODEN: JPCBFK; ISSN:1520-5207. (American Chemical Society)On the basis of the Poisson-Boltzmann equation in cylindrical coordinates, we calc. the cond. of a single charged nanotube filled with electrolyte. The cond. as a function of the salt concn. follows a power-law, the exponent of which has been controversially discussed in the literature. We use the co-ion-exclusion approxn. and obtain the crossover between different asymptotic power-law behaviors anal. Numerically solving the full Poisson-Boltzmann equation, we also calc. the complete diagram of exponents as a function of the salt concn. and the pH for tubes with different radii and pKa values. We apply our theory to recent exptl. results on carbon nanotubes using the pKa as a fit parameter. In good agreement with the exptl. data, the theory shows power-law behavior with the exponents 1/3 at high pH and 1/2 at low pH, with a crossover depending on salt concn., tube radius and pKa.