scispace - formally typeset
Search or ask a question

Showing papers on "Acetonitrile published in 1977"


Journal ArticleDOI
TL;DR: In this paper, the polarity and selectivity of several mobile phases used with reversed-phase high-pressure liquid chromatography columns were studied, including water, methanol, acetonitrile and tetrahydrofuran.

151 citations


Journal ArticleDOI
TL;DR: In this paper, the Gibbs free energies of transfer based on the bisbiphenylchromium assumption were compared with data obtained from the tetraphenylarsonium tetrahenylborate assumption in those solvents for which such data are available.

89 citations


Journal ArticleDOI
TL;DR: Raman spectroscopic studies on acetonitrile in the liquid and in dilute solutions have been used to investigate the vibrational relaxation process as discussed by the authors, which has been shown to be similar to the one described in this paper.

84 citations


Journal ArticleDOI
TL;DR: In this paper, a transient optical absorption was observed in the spectral range 500 to 1500 nm during nanosecond pulse radiolysis of liquid acetonitrile, and it was concluded that the two species are different chemical forms of a reducing entity joined by an equilibrium.
Abstract: During nanosecond pulse radiolysis of liquid acetonitrile a transient optical absorption was observed in the spectral range 500 to 1500 nm. The extreme reactivity of this species towards solutes capable of accepting electrons allowed its assignment as a reducing entity. In the purest liquid prepared this species had a half-life of some 900 ns. Variable temperature studies showed that this reducing species, having an intense maximum near 1450 nm at + 62°C, was drastically reduced in absorption intensity at –40°C, where the decay character changed and a weak maximum was observed near 550 nm. It was concluded that the two species are different chemical forms of a reducing entity joined by an equilibrium. Tentative identification of the reversible change according to (CH3CN)–+ CH3CN ⇌(CH3CN)–2 with the monomeric anion as the infrared absorbing entity was based on thermodynamic considerations and the effects of adding polar liquids. Quantitative examination of the data allowed a value of –34.9 kJ mol–1 to be extracted for the binding energy of the anion dimer. Measurements of the yields of pyrene and trans-stilbene radical anion formed after pulse radiolysis of appropriate solutions led to the evaluation of G(reducing entity)= 1.03. In addition it was found that O– radicals, formed by reaction between (CH3CN)– and N2O, reacted with acetonitrile via both addition and abstraction routes. Further, in solutions containing tetracyanobenzene (TCNB) and O2, capture of the negative entity by O2 was followed by electron transfer from O–2 to TCNB.

64 citations


Journal ArticleDOI
TL;DR: In this paper, the formation constants for the complexation of Cs/sup +/ in six solvents by three crown ethers, 18-crown-6, dibenzo-18crown 6, and dicyclohexyl-18-c Crown 6, were studied at 25/sup 0/C by cesium-133 NMR techniques.
Abstract: The formation constants for the complexation of Cs/sup +/ in six solvents by three crown ethers, 18-crown-6, dibenzo-18-crown-6, and dicyclohexyl-18-crown-6, were studied at 25/sup 0/C by cesium-133 NMR techniques. The solvents used, in the approximate order of decreasing first complexation constant for all crown ethers, were pyridine, acetone, propylene carbonate, acetonitrile, N,N-dimethylformamide, and dimethyl sulfoxide. In all solvents, 18-crown-6 formed both 1:1 and 2:1 (ligand to metal) complexes with enough difference between the complexation constants that the second constant could be determined. The formation of 2:1 complexes by the other two crown ethers was also indicated in most solvents but the second complexation constant could only be measured for the dibenzo-18-crown-6 complex in pyridine. In many cases the first complexation constant was too large to measure by the NMR technique so that only lower bounds could be established. The rate of loss of the Cs/sup +/ ion from dicyclohexyl-18-crown-6 and from monobenzo-2,2,2-cryptand in propylene carbonate was measured as a function of temperature. The activation energy for the latter case (15 kcal mol/sup -1/) is substantially larger than for the crown ether complex (8.5 kcal mol/sup -1/). Although the chemical shifts of both Cs/sup +/ and its 1:1 complex with 18-crown-6more » are strongly solvent dependent, the chemical shift of the 2:1 complex is independent of solvent, indicating that in the ''sandwich'' complex the cesium ion is effectively shielded from the solvent.« less

63 citations


Journal ArticleDOI
TL;DR: In this paper, the acceptor numbers for binary mixtures of water with acetonitrile (AN), dioxane, acetone (AC), pyridine (PY), N,N-dimethylformamide (DMF), dimethyl sulfoxide (DMSO), hexamethylphosphoramide (HMPA), methanol, ethanol andi-propyl alcohol were determined both by selective solvation and specific solvent-solvent interactions.
Abstract: Acceptor numbers have been determined for binary mixtures of water with acetonitrile (AN), dioxane, acetone (AC), pyridine (PY), N,N-dimethylformamide (DMF), dimethyl sulfoxide (DMSO), hexamethylphosphoramide (HMPA), methanol, ethanol andi-propyl alcohol. The electrophilic properties of binary aqueous-non aqueous solvent mixtures are determined both by selective solvation and specific solvent-solvent interactions. The variation of the acceptor number as a function of solvent composition is interpreted in terms of the previously determined nucleophilic and electrophilic properties of the pure components and their specific solvent structure.

54 citations


Journal ArticleDOI
TL;DR: In this paper, the authors measured the lifetimes of the triplet excited states of thionine and methylene blue in 50 v/v% aqueous acetonitrile solutions acidified with 0.01 N sulfuric or trifluoromethyl-sulfonic acid.
Abstract: — The lifetimes of the triplet excited states of thionine and methylene blue were measured in aqueous and 50 v/v% aqueous acetonitrile solutions acidified with 0.01 N sulfuric or trifluoromethyl-sulfonic acid. The rate constants for reaction of the triplet excited dyes with ferrous ions were measured in the same solutions. The triplet lifetimes in the absence of added quenchers were insensitive to a change in acid from trifluoromethylsulfonic to sulfuric or to a change in solvent from water to 50v/v% aqueous acetonitrile (τ for triplet thionine ˜7.5 μs, τ for triplet methylene blue ˜4.5 μs). In contrast, the rate constant for reaction of the triplet dyes with ferrous ions increased by nearly a factor of 10 with a change in acid from trifluoromethylsulfonic to sulfuric. In solutions containing sulfate ions this reaction rate constant increased with increasing sulfate concentration and with a change in solvent from water to 50 v/v% aqueous acetonitrile. The results are discussed in terms of the possibility of association of the positively charged reactive ions with sulfate anions. Quenching of the triplet excited dyes by ferric ions or by ground state dye molecules was shown to be negligible at the concentration used for the ferrous ion quenching study.

34 citations


Journal ArticleDOI
TL;DR: In this article, the structure of the lithium derivatives of phenyl acetonitrile and acetitrile is discussed in terms of their IR spectra and the influence of the counter ion and the solvent is discussed.

29 citations


Journal ArticleDOI
TL;DR: The reactions of 2,4- and 2,6-disubstituted phenols with two molar amounts of 1,3-benzodithiolylium tetrafluoroborate in acetonitrile at room temperature followed by treatment with triethylamine afforded the highly stable quinone methides in good yields.
Abstract: The reactions of 2,4- and 2,6-disubstituted phenols with two molar amounts of 1,3-benzodithiolylium tetrafluoroborate in acetonitrile at room temperature followed by treatment with triethylamine afforded the highly stable quinone methides, 2-(1,3-benzodithiol-2-ylidene)-3,5-cyclohexadien-1-ones and 4-(1,3-benzodithiol-2-ylidene)-2,5-cyclohexadien-1-ones, respectively, in good yields.

26 citations



Journal ArticleDOI
TL;DR: In this article, the lowest triplet of duroquinone, 3Q, with tertiary amines (triethylamine TEA, diethylaniline DEA, triphenylamine TPA) has been investigated by laser flash absorption spectroscopy in non-polar (cyclohexane or benzene) and polar (acetonitrile) solvents.
Abstract: The reaction of the lowest triplet of duroquinone, 3Q, with tertiary amines (triethylamine TEA, diethylaniline DEA, triphenylamine TPA) has been investigated by laser flash absorption spectroscopy in non-polar (cyclohexane or benzene) and polar (acetonitrile) solvents. Three pathways of 3Q deactivation involving an exciplex as intermediate were observed: (i) an electron transfer to 3Q from TPA and DEA in a polar solvent, (ii) a H atom transfer to 3Q from DEA in a non-polar solvent and from TEA in a non-polar or a polar solvent, (iii) a physical quenching of 3Q by TPA in a non-polar solvent with no photoreduction products.

Journal ArticleDOI
TL;DR: In this paper, the reaction of AgBF4 with CH3CN in acetonitrile affords the compound [(CH3CN)3Re(CO)3][BF4], which crystallizes in monoclinic space group P21/c with unit cell dimensions a
Abstract: Reaction of [n-Bu4N]2[Re4(CO)16] with AgBF4 in acetonitrile affords the compound [(CH3CN)3Re(CO)3][BF4]. The latter crystallizes in monoclinic space group P21/c with unit cell dimensions a = 11.021...

Journal ArticleDOI
TL;DR: In this article, the reduction of Ce(O-i-C3H7)4·B, where B=i-c3h7OH or pyridine, by (C2H5)3Al in the presence of cyclooctatetraene yields a new crystalline complex, η8-C8H8)Ce(Oc3c7)2Al(C2h5)2)2 (I) which can coordinate one mol of acetonitrile forming a crystalline adduct (II).

Journal ArticleDOI
TL;DR: In this article, the results of Spectroscopic data for Cu(MF6)2·5MeCN and MeCN (M = Ta or P) are discussed. But the results for MeCN are limited.

Journal ArticleDOI
TL;DR: In this paper, the nuclear magnetic resonances of alkali nuclei,7Li,23Na, and133Cs, as well as far infrared measurements are used to study alkali complexes of a bicyclic diazapolyoxa ligand, the dilactam of cryptand C222.
Abstract: Nuclear magnetic resonances of alkali nuclei,7Li,23Na, and133Cs, as well as far infrared measurements are used to study alkali complexes of a bicyclic diazapolyoxa ligand—the dilactam of cryptand C222. Measurements were carried out in pyridine, tetrahydrofuran, acetonitrile, nitromethane, dimethylformamide, and aqueous solutions. The complexing ability of the dilactam is similar to, but weaker than, that of the cryptand C222. The limiting chemical shifts of the complexed cations were solvent-dependent, indicating incomplete enclosure of the cation by the ligand. Formation constants of Li+ and Cs+ complexes were calculated from the chemical-shift dependence on the ligand/metal ion mole ratio.

Journal ArticleDOI
TL;DR: The electrochemical behavior of iron diimine complexes on a platinum working electrode in acetonitrile is described in this article, where the acceptor properties of the diimines do not depend on the presence of the aromatic rings, but on the iron-diimine chromophore.

Journal ArticleDOI
TL;DR: In this paper, the reaction of [Fe(bipy)3]2+ with cyanide ion in binary aqueous solvent mixtures containing ethylene glycol glycerol, acetonitrile, dimethyl sulphoxide (dmso), or tris (dimethylamino)phosph ine oxide (tdpo) was reported.
Abstract: Rates of reaction of [Fe(bipy)3]2+ with cyanide ion in binary aqueous solvent mixtures containing ethylene glycol glycerol, acetonitrile, dimethyl sulphoxide (dmso), or tris (dimethylamino)phosph ine oxide (tdpo), and with hydroxide ion in binary aqueous solvent mixtures containing t-butyl alcohol, dmso, or tdpo, are reported. Reactivity trends in the various solvent mixtures are discussed and, in the case of the reaction with hydroxide in aqueous dmso, compared with reactivity trends for similar inorganic and organic reactions. Where the existence of the required thermodynamic data on Gibbs free energies of transfer of single ions permits, it is shown that the effect of solvation, and thence chemical-potential, changes for cyanide and hydroxide ions is dominant in determining reactivies in their reactions with [Fe(bipy)3]2+.


Journal ArticleDOI
TL;DR: In this paper, the binding energies of ion-molecule complexes in the gas phase were measured and compared with single ion energies of solvation, showing that differences in ion solvation in solution are reflected in the binding energy of ion molecules in aprotic solvents.
Abstract: Measurement of the gas phase ion equilibria between ions M and solvent molecules Sl provide binding energies of the complexes M±(S1)n(for n= 1 to ∼6). Comparison of these data with single ion energies of solvation shows that differences in ion solvation in solution are reflected in the binding energies of ion-molecule complexes in the gas phase. The weaker solvation of negative ions (relative to positive ions) observed in liquid aprotic solvents is reflected in the binding energies of negative ion aprotic molecule complexes, a weaker binding being found for the first and subsequent few aprotic molecules. An analysis of the bonding in Cl–(CH3CN) and K+(CH3CN) shows that the weaker bonding to Cl– is due to the very diffuse distribution of the positive pole of the dipole in acetonitrile. In effect the dipole can not come close to the negative ion. Analysis shows a similar picture also for acetone. Experimental results for the bonding between Cl–HR are given for a variety of compounds HR. These show that for RH = protic compounds, like oxygen acids, the hydrogen bond in Cl–HR increases with the acidity of HR. For aprotic compounds, i.e., carbon acids, no relationship between the bond in the complex and the acidity of HR is found. An examination of the solvation of substituted phenoxide ions by protic and aprotic solvents shows that solvation by protic solvents is adversely affected by charge dispersal in the ion, while aprotic solvents are much less sensitive to charge dispersal. The reasons for this important difference in behaviour are examined.

Journal ArticleDOI
TL;DR: In this paper, the authors measured near and vacuum uv absorption spectra for aliphatic amine-acetic acid systems in the gaseous state and in solution at room temperature.

Journal ArticleDOI
TL;DR: In this article, the electrochemical bromination of benzene, toluene and p-xylene has been investigated in acetonitrile; even if the mechanism was found to be somewhat similar to that postulated in a previous paper for acetic acid as the solvent, the yields of the brominated products in this case are near to 100% because the solvent is not brominating in these conditions.

Journal ArticleDOI
TL;DR: In this paper, the authors measured 25 ns laser flashes (λ = 347.1 nm) to elucidate the mechanism of the reaction of triplet excited benzophenone (BP) and derivatives with unsaturated compounds whose triplet energy ET is lower than that of BP compounds.

Journal ArticleDOI
TL;DR: In this paper, the authors proposed a method for the synthesis of neutral adducts from titanium, zirconium, or hafnium in the presence of a solution of chlorine or bromine in acetonitrile (L) in good yield.
Abstract: Electrochemical oxidation of titanium, zirconium, or hafnium(IV) in the presence of a solution of chlorine or bromine (X) in acetonitrile (L) leads to direct synthesis of MX4L2 species in good yield. These compounds are easily transformed into other neutral adducts. On addition of tetraalkylammonium salts to the solution phase, the products are the salts (R4N)MCl5 or (R4N)2MBr6, except that with titanium Et4NTiBr4 was also formed under some conditions. The advantages of this method are discussed, and a possible reaction mechanism proposed.

Journal ArticleDOI
TL;DR: In this article, the authors compared the reactivity of solvated electron reactions in ammonia with some inorganic ions and organic neutral molecules with corresponding data in water, and concluded that in general the reactivities of the electron in ammonia are appreciably lower than in water.
Abstract: New and literature data of solvated electron reactions in ammonia with some inorganic ions and organic neutral molecules are compared with corresponding data in water. In ammonia only a few reactions with aromatic molecules are diffusion controlled and therefore faster than in water (k ≈ 1 × 1011 and 1 × 1010 M−1 s−1,respectively). After correcting for the electrostatic contribution to the rate constant of the other reactions it is concluded that in general the reactivity of the solvated electron in ammonia is appreciably lower, than in water. For the slow reactions of ammoniated electrons with acetonitrile and dimethylsulfoxide we find activation energies of 7 to 9 kcal/mol and activation volumes of −40 to −60 ml/mol. In these reactions it is suggested that the rate determining step is associated with the collapse of the large electron cavity in liquid ammonia.

Journal ArticleDOI
TL;DR: In this article, the complex formation between sulfur dioxide and iodide ions in acetonitrile (AN) has been studied by vapour pressure measurements, and the enthalpy changes ΔH0 for 1:1 association reactions between SO2 an...
Abstract: Complex formation between sulfur dioxide and iodide ions in acetonitrile (AN) has been studied by vapour pressure measurements. The enthalpy changes ΔH0 for 1:1 association reactions between SO2 an...


Journal ArticleDOI
TL;DR: RuCl2CO(MeCN)C8H12 as mentioned in this paper was shown by X-ray diffraction analysis to have octahedral stereochemistry with a trans-OC-Ru-MeCN arrangement.
Abstract: Reaction of [RuCl2CO(C8H12)]2 with acetonitrile gives [RuCl2CO(MeCN)C8H12](Ia), shown by X-ray diffraction analysis to have octahedral stereochemistry with a trans-OC–Ru–MeCN arrangement. Crystals are monoclinic, space group P21/c, with a= 12.933(2), b= 7.641 (8), c= 13.435(3)A, β= 104.36(2)° and contain one molecule per asymmetric unit. The structure was solved by use of 1 086 film data, and refined to R 0.09. The related complexes [RuCl2CO(L)C8H12](L = acrylonitrile and cyclopropyl cyanide) have also been synthesised.

Journal ArticleDOI
TL;DR: In this paper, the authors studied hydrogen evolution reactions from acidic MeCN solutions at Pt and Au electrodes and showed that the reaction is irreversible with a Tafel slope of 2 RT/F after diffusion polarization and pseudo-ohmic drop.

Journal ArticleDOI
TL;DR: In this paper, the rate law d[MoIV]/dt=k2[MoVI][PPh3], with k2(25 °C) = 1.1 ± 0.3 dm3 mol−1 s−1, describes the kinetic data obtained, with activation parameters at 25 °C of ΔH‡= 8.4 ± 1.5 kcal mol −1 and ΔS ǫ = −30 ± 1
Abstract: The kinetics of the oxygen-transfer reaction from [MoO2(S2CNEt2)2] to PPh3 in acetonitrile solution have been monitored using stopped-flow techniques at temperatures between 15 and 45 °C. The rate law d[MoIV]/dt=k2[MoVI][PPh3], with k2(25 °C)= 1.1 ± 0.3 dm3 mol–1 s–1, describes the kinetic data obtained, with activation parameters at 25 °C of ΔH‡= 8.4 ± 0.5 kcal mol–1 and ΔS‡=–30 ± 1.6 cal K–1 mol–1.

Journal ArticleDOI
TL;DR: In this article, a cyclic voltammetry and controlled potential electrolysis at a glassy-carbon anode in acetonitrile was carried out to investigate anodic oxidation of benzenesulfon-p-anisidide (I), n-methyl-3-benzenes-ulfonamido-6-methoxyphenyl) pyridinium perchlorate (VIII).
Abstract: Anodic oxidation of benzenesulfon-p-anisidide (I), benzenesulfon-p-toluidide (II), and N-methylbenzenesulfon-p-anisidide (III) were investigated by cyclic voltammetry and controlled potential electrolysis at a glassy-carbon anode in acetonitrile. I showed a single anodic wave in the absence of pyridine, and three anodic waves in the presence of pyridine. On electrolysis of I in the absence of pyridine, p-benzoquinone and benzenesulfonamide were formed. On electrolysis of I in the presence of pyridine, N, N'-dibenzenesulfonyl-N-(4'-methoxyphenyl)-5-methoxy-o-phenylenediamine (IV) and 1-(2-benzenesulfonamido-5-methoxyphenyl) pyridinium perchlorate (V) were obtained. Anodic oxidation of II in acetonitrile containing excess of pyridine gave N, N'-dibenzenesulfonyl-N-(p-tolyl)-5-methyl-o-phenylenediamine (VI) and N, N'-dibenzenesulfonyl-5, 5'-dimethyl-2, 2'-biphenyldiamine (VII). The formation of VI and VII was interpreted in terms of one-electron transfer followed by dimerizations of the free radical of II. On electrolysis of III in the presence of pyridine, 1-(N-methyl-3-benzenesulfonamido-6-methoxyphenyl) pyridinium perchlorate (VIII) was formed as the main product. The position of pyridination, which was ortho to the methoxy-group, was explained on the basis of steric factors.