To see the other types of publications on this topic, follow the link: Solvolysis. Solution (Chemistry) Solvents.

Journal articles on the topic 'Solvolysis. Solution (Chemistry) Solvents'

Create a spot-on reference in APA, MLA, Chicago, Harvard, and other styles

Select a source type:

Consult the top 50 journal articles for your research on the topic 'Solvolysis. Solution (Chemistry) Solvents.'

Next to every source in the list of references, there is an 'Add to bibliography' button. Press on it, and we will generate automatically the bibliographic reference to the chosen work in the citation style you need: APA, MLA, Harvard, Chicago, Vancouver, etc.

You can also download the full text of the academic publication as pdf and read online its abstract whenever available in the metadata.

Browse journal articles on a wide variety of disciplines and organise your bibliography correctly.

1

Pieper, Thomas, Wolfgang Peti, and Bernhard K. Keppler. "Solvolysis of the Tumor-Inhibiting Ru(III)-Complex trans-Tetrachlorobis(Indazole)Ruthenate(III)." Metal-Based Drugs 7, no. 4 (January 1, 2000): 225–32. http://dx.doi.org/10.1155/mbd.2000.225.

Full text
Abstract:
The ruthenium(III) complex Hlnd trans-[RuCl4,(ind)2], with two trans-standing indazole (ind) ligands bound to ruthenium via nitrogen, shows remarkable activity in different tumor models in vitro and in vivo. The solvolysis of the complex trans-[RuCl4,(ind)2]- has been investigated by means of spectroscopic techniques (UV/vis, NMR)in different solvents. We investigated the indazolium as well as the sodium salt, the latter showing improved solubility in water. In aqueous acetonitrile and ethanol the solvolysis results in one main solvento complex. The hydrolysis of the complex is more complicated and depends on the pH of the solution as well as on the buffer system.
APA, Harvard, Vancouver, ISO, and other styles
2

Thiare, Diène Diègane, Abdourakhmane Khonté, Diegane Sarr, Cheikh Diop, Mame Diabou Gaye-Seye, Atanasse Coly, François Delattre, Alphonse Tine, and Jean-Jacques Aaron. "Solvolysis kinetic study and direct spectrofluorimetric analysis of the fungicide benomyl in natural waters." Macedonian Journal of Chemistry and Chemical Engineering 33, no. 2 (November 7, 2014): 237. http://dx.doi.org/10.20450/mjcce.2014.513.

Full text
Abstract:
<p>A direct spectrofluorimetric method for the quantitative analysis of benomyl in natural waters is described. Benomyl is an instable, fluorescent fungicide that mainly decomposes into carbendazim and n-butyl-isocyanate in organic and aqueous solutions. The kinetics of benomyl solvolysis reactions were investigated in organic solvents (methanol and acetonitrile) and in aqueous solvent systems, including β–cyclodextrin (β-CD), sodium dodecyl sulfate (SDS), dodecyltrimethylammonium chloride (DTAC), cetyltrimethylammonium chloride (CTAC), cetyltrimethylammonium hydroxide (CTAOH), Brij-700, Triton X-100 and water, at different pH and/or NaOH concentrations. The benomyl fluorescence signal was found to be quasi-completely stable in 10<sup>-2</sup> M NaOH aqueous solution, various alkaline (10<sup>-2</sup> M NaOH) organized media, β-CD neutral solution and Triton X-100 aqueous solutions of different pH. Based on these results, a direct spectrofluorimetric analytical method was developed for the determination of benomyl in 10<sup>-2</sup> M NaOH aqueous solution and Triton X-100 solutions (pH7 and 10<sup>-2</sup> M NaOH), with wide linear dynamic range (LDR) values of two to three orders of magnitude, very low limit of detection (LOD) and limit of quantification (LOQ) values of, respectively, 0.002-0.5 ng/mL and 0.007-2.0 ng/mL, and small relative standard deviation (RSD) values of 0.2-1.7 %, according to the medium. This direct spectrofluorimetric method was applied to the evaluation of benomyl residues in natural waters, with satisfactory recovery values (87-94%).</p>
APA, Harvard, Vancouver, ISO, and other styles
3

Hojo, Masashi. "Elucidation of specific ion association in nonaqueous solution environments." Pure and Applied Chemistry 80, no. 7 (January 1, 2008): 1539–60. http://dx.doi.org/10.1351/pac200880071539.

Full text
Abstract:
The paper reviews ion aggregation in ionic solution in solvents of low and high permittivity. Although higher ion aggregates from 1:1 type electrolytes in low-pemittivity media (εr &lt; 10) are widely accepted, only a few chemists have recognized the higher ion aggregation in the higher-permittivity media. However, we have clarified that the chemical interaction, such as coordination, can operate between simple anions and cations in nonaqueous solvents (20 &lt; εr &lt; 65) of low solvation ability. Acids (HA) and their conjugate base anions (A-) may react with each other to form homoconjugated species, such as A-(HA)2, in acetonitrile or benzonitrile (i.e., solvents possessing poor hydrogen-bonding donor and acceptor abilities). An analytical method of conductivity data for 1:1 electrolytes has been developed and successfully applied to very complicated systems, in which the ion pair (1:1), triple ions (2:1 and 1:2), and the quadrupole (2:2 association) are involved in a solution at the same time. After observing the direct reaction of some anions (e.g., Cl-) and cations (e.g., Li+) toward a certain species, we interpreted comprehensively the salt effects in chemical equilibria, based on distinct chemical interactions and not merely a vague term, "medium effect". The mechanism of salt effects on solvolysis reactions of the SN1 type in organic-aqueous mixed solvents has been elucidated. We discussed that a reaction manner similar to that in nonaqueous solution can take place even in some "aqueous" solution if the huge network of hydrogen-bonding of bulk water (the number of water, nw &gt; ~108) is destroyed due to any spatial barriers (such as ions, molecules, surface) or elevated temperature.
APA, Harvard, Vancouver, ISO, and other styles
4

Dionísio, Madalena S., Joaquim J. Moura Ramos, and Raquel M. Gonçalves. "The enthalpy and entropy of cavity formation in liquids and Corresponding States Principle." Canadian Journal of Chemistry 68, no. 11 (November 1, 1990): 1937–49. http://dx.doi.org/10.1139/v90-299.

Full text
Abstract:
A new method to calculate the enthalpy and entropy of cavity formation in liquids is proposed. The reference cavity formation process is identified with the vaporization of liquid in the absence of order and Corresponding States Principle is used to eliminate the order contribution to vaporization enthalpy. The proposed method agrees very well with Claverie's modification of the Pierotti's method but strongly disagrees with the Sinanoglu's method, particularly in the context of entropy of cavity formation.The new method is checked by applying it to the description of solution process in alkane binary mixtures at infinite dilution and is used to study the solvent effect on the solvolysis of t-butyl chloride and bromide. Keywords: cavity formation, enthalpy and entropy of cavity formation, Corresponding States Principle, alkanes, intermolecular interactions, acentric factor, solvent effect.
APA, Harvard, Vancouver, ISO, and other styles
5

Brown, Robert Stan. "Bio-inspired approaches to accelerating metal ion-promoted reactions: enzyme-like rates for metal ion mediated phosphoryl and acyl transfer processes." Pure and Applied Chemistry 87, no. 6 (June 1, 2015): 601–14. http://dx.doi.org/10.1515/pac-2014-1008.

Full text
Abstract:
Abstract Intense efforts by many research groups for more than 50 years have been directed at biomimetic approaches to understand how enzymes achieve their remarkable rate accelerations. Nevertheless, it was noted in 2003 that, despite numerous efforts to design models for catalyzing the cleavage of such species as phosphate diesters, “none of the several models so far described approaches the enormous catalytic efficiency of natural enzymes”. The same could be said for biomimetics of other enzymes promoting acyl or phosphoryl transfer reactions, particularly those mediated by metal ions such as Zn(II). Clearly other important factors were being overlooked or awaiting discovery. In this manuscript we describe two important effects that we have implemented to accelerate metal ion catayzed phosphoryl and acyl transfer reactions. The first of these relates to a medium effect where the polarity of the solution, as measured by dielectric constant, is reduced from that of water (ε = 78) to values of 31.5 and 24.3 when the solvent is changed to methanol or ethanol. Among organic solvents these light alcohols are closest to water in terms of structure and properties as well as retaining important H-bonding properties. The second important effect involves a known but difficult to demonstrate mode of catalysis where the leaving group (LG) in a solvolysis reaction is accelerated as it becomes progressively poorer. In the cases described herein, the LG’s propensity to depart from a substrate during the course of reaction is accelerated by coordination to a metal ion in a process known as leaving group assistance, or LGA. These two effects can each impart accelerations of 109–1017 for certain metal ion catalyzed reactions relative to the corresponding solvent, or base induced reactions.
APA, Harvard, Vancouver, ISO, and other styles
6

Bolton, Judy L., and Robert A. McClelland. "Azide ion trapping and lifetime in aqueous solution of a primary carbenium ion stabilized by a 2-imidazoyl ring." Canadian Journal of Chemistry 67, no. 7 (July 1, 1989): 1139–43. http://dx.doi.org/10.1139/v89-171.

Full text
Abstract:
2-Chloromethyl-1-methylimidazole undergoes a pH-dependent aqueous hydrolysis with the neutral substrate being the reactive species, and the imidazole-protonated form (pKa = 5.7) unreactive. Addition of sodium chloride retards the hydrolysis, evidence that there is a free carbenium ion intermediate (the common ion effect). The rate constant ratio Kcl/Kw for the reactions of this cation with the added chloride and with the solvent is 7.4 M−1. Further evidence for a free cation is the observation of the 2-azidomethyl product when the hydrolysis is carried out with sodium azide present, but with no change in the rate constant. The Kaz/Kw ratio as determined by product analysis is 1.1 × 102 M−1 With the assumption that kaz represents a diffusion-controlled reaction and has a value of 7 × 109 M−1 s−1, the rate constant kw for the reaction of the cation with solvent is 6 × 107 s−1. A comparison with azide–water selectivity ratios reported for other cations shows that the imidazole-stabilized primary cation of this study is relatively long-lived. A possible explanation for this is given, in terms of the extensive resonance delocalization of the positive charge in this cation. Keywords: solvolysis, carbenium ion, heterocycle.
APA, Harvard, Vancouver, ISO, and other styles
7

Park, Kyoung Ho, Mi Hye Seong, Jin Burm Kyong, and Dennis N. Kevill. "Rate and Product Studies with 1-Adamantyl Chlorothioformate under Solvolytic Conditions." International Journal of Molecular Sciences 22, no. 14 (July 9, 2021): 7394. http://dx.doi.org/10.3390/ijms22147394.

Full text
Abstract:
A study was carried out on the solvolysis of 1-adamantyl chlorothioformate (1-AdSCOCl, 1) in hydroxylic solvents. The rate constants of the solvolysis of 1 were well correlated using the Grunwald–Winstein equation in all of the 20 solvents (R = 0.985). The solvolyses of 1 were analyzed as the following two competing reactions: the solvolysis ionization pathway through the intermediate (1-AdSCO)+ (carboxylium ion) stabilized by the loss of chloride ions due to nucleophilic solvation and the solvolysis–decomposition pathway through the intermediate 1-Ad+Cl− ion pairs (carbocation) with the loss of carbonyl sulfide. In addition, the rate constants (kexp) for the solvolysis of 1 were separated into k1-Ad+Cl− and k1-AdSCO+Cl− through a product study and applied to the Grunwald–Winstein equation to obtain the sensitivity (m-value) to change in solvent ionizing power. For binary hydroxylic solvents, the selectivities (S) for the formation of solvolysis products were very similar to those of the 1-adamantyl derivatives discussed previously. The kinetic solvent isotope effects (KSIEs), salt effects and activation parameters for the solvolyses of 1 were also determined. These observations are compared with those previously reported for the solvolyses of 1-adamantyl chloroformate (1-AdOCOCl, 2). The reasons for change in reaction channels are discussed in terms of the gas-phase stabilities of acylium ions calculated using Gaussian 03.
APA, Harvard, Vancouver, ISO, and other styles
8

D’Souza, Malcolm J., Jeremy Wirick, Osama Mahmoud, Dennis N. Kevill, and Jin Burm Kyong. "The Influence of a Terminal Chlorine Substituent on the Kinetics and the Mechanism of the Solvolyses of n-Alkyl Chloroformates in Hydroxylic Solvents." International Journal of Molecular Sciences 21, no. 12 (June 19, 2020): 4387. http://dx.doi.org/10.3390/ijms21124387.

Full text
Abstract:
A previous study of the effect of a 2-chloro substituent on the rates and the mechanisms of the solvolysis of ethyl chloroformate is extended to the effect of a 3-chloro substituent on the previously studied solvolysis of propyl chloroformate and to the effect of a 4-chloro substituent on the here reported rates of solvolysis of butyl chloroformate. In each comparison, the influence of the chloro substituent is shown to be nicely consistent with the proposal, largely based on the application of the extended Grunwald–Winstein equation, of an addition-elimination mechanism for solvolysis in the solvents of only modest solvent ionizing power, which changes over to an ionization mechanism for solvents of relatively high ionizing power and low nucleophilicity, such as aqueous fluoroalcohols with an appreciable fluoroalcohol content.
APA, Harvard, Vancouver, ISO, and other styles
9

D’Souza, Malcolm J., Zoon Ha Ryu, Byoung-Chun Park, and Dennis N. Kevill. "Correlation of the rates of solvolysis of acetyl chloride and α-substituted derivatives." Canadian Journal of Chemistry 86, no. 5 (May 1, 2008): 359–67. http://dx.doi.org/10.1139/v08-028.

Full text
Abstract:
Additional specific rates of solvolysis have been determined for acetyl chloride and diphenylacetyl chloride. These are combined with literature values to carry out correlation analyses, using the extended Grunwald–Winstein equation with incorporation of literature values for solvent nucleophilicity (NT) and solvent ionizing power (YCl). Parallel analysis are carried out using literature values for the specific rates of solvolysis of trimethylacetyl chloride, chloroacetyl chloride, phenylacetyl chloride, and α-methoxy-α-trifluoromethylphenylacetyl chloride (MTPAC). Chloroacetyl chloride and MTPAC react by an addition-elimination pathway, with the addition step rate-determining, over the full range of solvents. Acetyl chloride reacts over the full range of solvents by an ionization pathway, with considerable nucleophilic solvation. The other three substrates can solvolyze with the domination of either mechanism, depending on the properties of the solvent. Reports concerning the use of product selectivity values, kinetic solvent isotope effects, and computational studies as additional probes of the mechanism of solvolysis are discussed.Key words: Grunwald-Winstein equation, acyl chlorides, mechanism of solvolysis, solvent nucleophilicity.
APA, Harvard, Vancouver, ISO, and other styles
10

Kevill, Dennis N., and Malcolm J. D’Souza. "Article." Canadian Journal of Chemistry 77, no. 5-6 (June 1, 1999): 1118–22. http://dx.doi.org/10.1139/v99-083.

Full text
Abstract:
The specific rates of solvolysis of phenyl chlorothionoformate (PhOCSCl) are remarkably similar to those previously reported for phenyl chlorothioformate (PhSCOCl). When analyzed using the extended Grunwald-Winstein equation over the usual range of solvent types, these solvolyses show essentially identical divisions into the solvents favoring the addition-elimination channel and those favoring the ionization channel. The introduction of one sulfur caused a partial shift away from the addition-elimination pathway, which was dominant over the full range of solvents for phenyl chloroformate (PhOCOCl). Consistent with these results, introduction of the second sulfur within phenyl chlorodithioformate (PhSCSCl) leads to a completion of this shift, such that an extended Grunwald-Winstein treatment of the specific rates of solvolysis now shows the ionization pathway to be dominant over the full range of solvents.Key words: Grunwald-Winstein equation, solvent nucleophilicity, solvolysis, phenyl chlorothionoformate, phenyl chlorodithioformate.
APA, Harvard, Vancouver, ISO, and other styles
11

Park, Kyoung-Ho, Chan Joo Rhu, Jin Burm Kyong, and Dennis N. Kevill. "The Effect of the ortho Nitro Group in the Solvolysis of Benzyl and Benzoyl Halides." International Journal of Molecular Sciences 20, no. 16 (August 18, 2019): 4026. http://dx.doi.org/10.3390/ijms20164026.

Full text
Abstract:
A kinetic study was carried out on the solvolysis of o-nitrobenzyl bromide (o-isomer, 1) and p-nitrobenzyl bromide (p-isomer, 3), and o-nitrobenzoyl chloride (o-isomer, 2) in a wide range of solvents under various temperatures. In all of the solvents without aqueous fluoroalcohol, the reactions of 1 were solvolyzed at a similar rate to those observed for 3, and the reaction rates of 2 were about ten times slower than those of the previously studied p-nitrobenzoyl chloride (p-isomer, 4). For solvolysis in aqueous fluoroalcohol, the reactivity of 2 was kinetically more reactive than 4. The l/m values of the extended Grunwald–Winstein (G–W) equation for solvolysis of 1 and 2 in solvents without fluoroalcohol content are all significantly larger than unity while those in all the fluoroalcohol solvents are less than unity. The role of the ortho-nitro group as an intramolecular nucleophilic assistant (internal nucleophile) in the solvolytic reaction of 1 and 2 was discussed. The results are also compared with those reported earlier for o-carbomethoxybenzyl bromide (5) and o-nitrobenzyl p-toluenesulfonate (7). From the product studies and the activation parameters for solvolyses of 1 and 2 in several organic hydroxylic solvents, mechanistic conclusions are drawn.
APA, Harvard, Vancouver, ISO, and other styles
12

Abraham, Michael H., Filomena Martins, Ruben Elvas-Leitão, and Luís Moreira. "Properties of the tert-butyl halide solvolysis transition states." Physical Chemistry Chemical Physics 23, no. 5 (2021): 3311–20. http://dx.doi.org/10.1039/d0cp05099g.

Full text
APA, Harvard, Vancouver, ISO, and other styles
13

Khalifa, M. A., A. M. Ismail, M. El-Batouti, and A. El-Hawaty. "kinetics of Aquation of Dichloro Tetrapyridine Ruthenium(Ii) Complex in Binary Aqueous Solvents." Journal of Chemical Research 2003, no. 1 (January 2003): 42–45. http://dx.doi.org/10.3184/030823403103172896.

Full text
Abstract:
First-order solvolysis rates of trans-dichloro tetrapyridine ruthenium(II) have been measured UV spectrophoto metrically over a wide range of solvent compositions in temperature ranges(40–55°C) in water–2-propanol and water– t-butanol mixtures. The rate of solvolysis is faster in the former than in the latter. Plots of log (rate constant) versus the reciprocal of relative permitivity of the co-solvent gave a non-linear relation for both co-solvents, this non-linearity is derived from a large differential effect of solvent structure between the initial and transition states. Δ S# of activation correlates well with the extrema in physical properties of the mixtures which are related to changes in solvent structure. Linear plots of Δ H# versus Δ S# were obtained and the isokinetic temperature indicates that the reaction is entropy controlled.
APA, Harvard, Vancouver, ISO, and other styles
14

Ishikawa, Ryuta, Shunya Ueno, Yumi Hamatake, Yoji Horii, Yuji Miyazaki, Motohiro Nakano, Takeshi Noda, Mikoto Uematsu, and Satoshi Kawata. "Versatile coordination architectures of products generated by the in situ reaction of a doubly bis(2-pyridyl)pyrazolate bridged dinuclear copper(ii) complex with tetracyanoethylene." CrystEngComm 21, no. 12 (2019): 1886–94. http://dx.doi.org/10.1039/c9ce00036d.

Full text
APA, Harvard, Vancouver, ISO, and other styles
15

Kevill, Dennis N., and Veena Upadhyay. "Solvolysis-decomposition ofN-1-adamantyl-N-p-tolylcarbamoyl chloride in hydroxylic solvents." Journal of Physical Organic Chemistry 10, no. 8 (August 1997): 600–606. http://dx.doi.org/10.1002/(sici)1099-1395(199708)10:8<600::aid-poc928>3.0.co;2-q.

Full text
APA, Harvard, Vancouver, ISO, and other styles
16

Kevill, Dennis N., and Zoon Ha Ryu. "Rate and Product Studies in the Solvolyses of Two Arenesulfonic Anhydrides." Journal of Chemical Research 2007, no. 6 (June 2007): 365–69. http://dx.doi.org/10.3184/030823407x218084.

Full text
Abstract:
The specific rates of solvolysis of benzenesulfonic anhydride (1) and p-toluenesulfonic anhydride (2) have been measured conductometrically at −10°C in 34 solvents for 1 and 33 solvents for 2. Studies at higher temperatures have allowed extrapolated values in additional solvents to be calculated. All of the values, for 35 solvents for 1 and for 37 solvents for 2, have been used in an extended Grunwald–Winstein equation treatment using NT and YOTs values. Activation parameters in several solvents and kinetic solvent isotope effects (MeOH/MeOD) have been determined for both substrates. Product selectivity values ( S) have been determined for binary mixtures of water with ethanol, methanol, or 2, 2, 2-trifluoroethanol. The results from the kinetic and product studies are compared to those previously reported for methanesulfonic anhydride (3). An SN2 mechanism is proposed for the solvolytic displacement reactions of the three substrates in all of the solvents used in the investigation.
APA, Harvard, Vancouver, ISO, and other styles
17

Kevill, Dennis N., Jong Chul Kim, and Jin Burm Kyong. "Correlation of the Rates of Solvolysis of Methyl Chloroformate with Solvent Properties." Journal of Chemical Research 23, no. 2 (February 1999): 150–51. http://dx.doi.org/10.1177/174751989902300242.

Full text
Abstract:
The specific rates of solvolyis of methyl chloroformate are very well correlated by the extended Grunwald–Winstein equation over a wide range of solvents; the pathway is believed to be predominantly addition–elimination, except that a positive deviation for solvolysis in 90% 1,1,1,3,3,3-hexafluoropropan-2-ol suggests an 80% contribution from an ionisation mechanism.
APA, Harvard, Vancouver, ISO, and other styles
18

Wiering, P. G., H. Steinberg, and Th J. de Boer. "Solvolysis of 7-exo-bicyclo[4.1.0]heptyl triflate in protic solvents: (Preliminary communication)." Recueil des Travaux Chimiques des Pays-Bas 96, no. 4 (September 2, 2010): 119–20. http://dx.doi.org/10.1002/recl.19770960408.

Full text
APA, Harvard, Vancouver, ISO, and other styles
19

Mayr, Herbert, and Armin R. Ofial. "How to predict changes in solvolysis mechanisms." Pure and Applied Chemistry 81, no. 4 (January 1, 2009): 667–83. http://dx.doi.org/10.1351/pac-con-08-08-26.

Full text
Abstract:
Stopped-flow and laser flash techniques have been employed to investigate the individual steps of the solvolysis reactions of benzhydryl (diarylmethyl) halides and carboxylates. In this way, absolute rate constants for the ionization (k1), recombination of the carbocation with the leaving group (k-1), and subsequent reaction with the solvent (kSolvOH) have been determined. As the stabilization of the carbocations increases, the mechanism changes from (a) SN1 reactions with irreversible ionization through (b) SN1 reactions with common-ion return and (c) SN2C+ reactions, where the intermediate carbocations accumulate, to (d) the formation of persistent carbocations which do not undergo subsequent reactions under the selected solvolysis conditions. The correlation equation log k = s(N + E), where the carbocations are characterized by the electrophilicity parameter E, and leaving groups and solvents are characterized by the nucleophile-specific parameters s and N can be employed to predict the changes of mechanism.
APA, Harvard, Vancouver, ISO, and other styles
20

Cherton, Jean-Claude, Marc Bazinet, Marie-Madeleine Bolze, Marc Lanson, and Paul-Louis Desbene. "Réactivité du nucléophile azoture vis-à-vis de cations hétérocycliques aromatiques. VIII. Réarrangement de β-tétrazolo-trans-benzalacétophénones." Canadian Journal of Chemistry 63, no. 10 (October 1, 1985): 2601–7. http://dx.doi.org/10.1139/v85-432.

Full text
Abstract:
β-Tetrazolyl-trans-benzalacetophenones (T) isomerize upon heating into the corresponding azido-azomethines (A). These non-isolable products undergo various transformations depending on the reaction conditions. With organic bases such as triphenylphosphine or pyridine, their interception can occur. In aromatic solvents, a rearrangement involving loss of nitrogen leads to five-membered diazoted heterocycles:-N-benzoyl imidazoles and pyrazoles, via intermediate diazirines and oxazepines. Protic solvents have been found to facilitate the isomerisation of tetrazoles into azidoazomethines. Solvolysis of the azidooxazine and in some cases of the oxazinium species resulting from the equilibrium and reaction:[Formula: see text]are then observed.
APA, Harvard, Vancouver, ISO, and other styles
21

Mollin, Jiří, Zdeněk Pavelek, and František Kašpárek. "Ratio of lyate ions activities in two-components protogenic solvents." Collection of Czechoslovak Chemical Communications 52, no. 5 (1987): 1115–30. http://dx.doi.org/10.1135/cccc19871115.

Full text
Abstract:
Kinetic method for the evaluation of lyate ions activities in two-component protogenic solvents has been discussed. Spectral methods for the measurement of equilibria between alkoxide and hydroxide ions have been compared. A correlation between dissociation constants of water and of weak organic acids in solvent mixtures has been found. By means of this correlation, ratios of lyate ions activities in water-ethanol and water-2-propanol mixtures have been calculated. These activity ratios, together with the known composition of reaction products, have been used to evaluate the selectivity of the neutral and alkaline solvolysis of phthalic anhydride and benzyl chloride in solvent mixtures quoted above.
APA, Harvard, Vancouver, ISO, and other styles
22

Fischer, Alfred, George N. Henderson, and Trevor A. Smyth. "Reactions of the 1-hydroxy-1,4-dimethylcyclohexadienyl cation, an intermediate in the solvolysis of 1,4-dimethyl-4-nitrocyclohexa-2,5-dien-1-ol." Canadian Journal of Chemistry 64, no. 6 (June 1, 1986): 1093–101. http://dx.doi.org/10.1139/v86-184.

Full text
Abstract:
Solvolysis of 1,4-dimethyl-4-nitrocyclohexa-2,5-dien-1-ol in mixed aqueous organic solvents gives the diastereomers of 1,4-dimethylcyclohexa-2,5-diene-1,4-diol, 1,4-dimethylcyclohexa-3,5-diene-1,2-diol, 2-nitro-p-xylene, 2,4-dimethylphenol (all derived from the title cation, itself formed by ionization of the nitro group as nitrite), and 2,5-dimethylphenol. In aqueous methanol the diastereomers of 4-methoxy-1,4-dimethylcyclohexa-2,5-dienol are also obtained. Significant yields of 2,5-dimethylphenol are only obtained on the acid-catalysed further reaction of the dienediol (or the methoxydienol) and involve the intermediate formation of 1,4-dimethylcyclohexa-3,5-diene-1,2-diol. In the absence of added base the acid released in the solvolysis catalyses this reaction and leads to the aromatization of the dienes.
APA, Harvard, Vancouver, ISO, and other styles
23

Fabregat, Azael, Jaume Giralt, and Francesc Giralt. "Solvolysis of a Catalan lignite with solvents of low hydrogen donor capacity." Fuel 66, no. 6 (June 1987): 835–39. http://dx.doi.org/10.1016/0016-2361(87)90133-5.

Full text
APA, Harvard, Vancouver, ISO, and other styles
24

D’Souza, Malcolm J., Matthew J. McAneny, Dennis N. Kevill, Jin Burm Kyong, and Song Hee Choi. "Kinetic evaluation of the solvolysis of isobutyl chloro- and chlorothioformate esters." Beilstein Journal of Organic Chemistry 7 (April 29, 2011): 543–52. http://dx.doi.org/10.3762/bjoc.7.62.

Full text
Abstract:
The specific rates of solvolysis of isobutyl chloroformate (1) are reported at 40.0 °C and those for isobutyl chlorothioformate (2) are reported at 25.0 °C, in a variety of pure and binary aqueous organic mixtures with wide ranging nucleophilicity and ionizing power. For 1, we also report the first-order rate constants determined at different temperatures in pure ethanol (EtOH), methanol (MeOH), 80% EtOH, and in both 97% and 70% 2,2,2-trifluoroethanol (TFE). The enthalpy (ΔH≠) and entropy (ΔS≠) of activation values obtained from Arrhenius plots for 1 in these five solvents are reported. The specific rates of solvolysis were analyzed using the extended Grunwald–Winstein equation. Results obtained from correlation analysis using this linear free energy relationship (LFER) reinforce our previous suggestion that side-by-side addition–elimination and ionization mechanisms operate, and the relative importance is dependent on the type of chloro- or chlorothioformate substrate and the solvent.
APA, Harvard, Vancouver, ISO, and other styles
25

Salmar, Siim, Jaak Järv, Tiina Tenno, and Ants Tuulmets. "Role of water in determining organic reactivity in aqueous binary solvents." Open Chemistry 10, no. 5 (October 1, 2012): 1600–1608. http://dx.doi.org/10.2478/s11532-012-0080-8.

Full text
Abstract:
AbstractKinetic data for organic reactions in various binary water-organic solvent mixtures were collected and quantitatively analysed in terms of linear-free-energy relationships by using tert-butyl chloride (2-chloro-2-methylpropane) solvolysis as the reference system. Linear similarity plots for these kinetic data were determined for solvent systems ranging from pure water mixtures up to considerable amount of cosolvent, and 161 similarity coefficients were calculated from slopes of these plots. The existence of these linear plots demonstrated that the solvent effects are of some common nature in all analysed reaction mixtures independent of the reaction type and the cosolvent used. Therefore it was concluded that the observed effects could be connected to the specific solvating properties of water, which govern reactivity even in significant dilution of water by an organic cosolvent. This conclusion was supported by the linear interrelationship between the slopes of similarity plots of different reactions, and hydrophobicity parameters log P of the reacting compounds. The relative solvent effects observed in binary water-organic solvent mixtures were for the first time directly related to the structure of reacting compounds.
APA, Harvard, Vancouver, ISO, and other styles
26

McManus, Samuel P., and Ahmad Safavy. "Solvolysis in mixed solvents with complementary electrophilic and nucleophilic properties. Hexafluoro-2-propanol and 1,3-propanedithiol." Journal of Organic Chemistry 51, no. 18 (September 1986): 3532–35. http://dx.doi.org/10.1021/jo00368a026.

Full text
APA, Harvard, Vancouver, ISO, and other styles
27

Manege, Ludovick Christian, Tadaharu Ueda, and Masashi Hojo. "Concentrated Salt Effects on the Solvolysis Reaction Rates in Methanol–Water Solution." Bulletin of the Chemical Society of Japan 71, no. 3 (March 1998): 589–96. http://dx.doi.org/10.1246/bcsj.71.589.

Full text
APA, Harvard, Vancouver, ISO, and other styles
28

Kevill, Dennis N., Byoung-Chun Park, and Jin Burm Kyong. "Evidence for general base catalysis by protic solvents in a kinetic study of alcoholyses and hydrolyses of 1-(phenoxycarbonyl)pyridinium ions under both solvolytic and non-solvolytic conditions." Collection of Czechoslovak Chemical Communications 74, no. 1 (2009): 43–55. http://dx.doi.org/10.1135/cccc2008164.

Full text
Abstract:
The kinetics of nucleophilic substitution reactions of 1-(phenoxycarbonyl)pyridinium ions, prepared with the essentially non-nucleophilic/non-basic fluoroborate as the counterion, have been studied using up to 1.60 M methanol in acetonitrile as solvent and under solvolytic conditions in 2,2,2-trifluoroethan-1-ol (TFE) and its mixtures with water. Under the non- solvolytic conditions, the parent and three pyridine-ring-substituted derivatives were studied. Both second-order (first-order in methanol) and third-order (second-order in methanol) kinetic contributions were observed. In the solvolysis studies, since solvent ionizing power values were almost constant over the range of aqueous TFE studied, a Grunwald–Winstein equation treatment of the specific rates of solvolysis for the parent and the 4-methoxy derivative could be carried out in terms of variations in solvent nucleophilicity, and an appreciable sensitivity to changes in solvent nucleophilicity was found.
APA, Harvard, Vancouver, ISO, and other styles
29

Stein, Allan R. "β-Deuterium kinetic isotope effects for identity processes: bromide ion substitution at 1-bromo-1-arylethanes and 2-bromooctane." Canadian Journal of Chemistry 72, no. 8 (August 1, 1994): 1789–96. http://dx.doi.org/10.1139/v94-227.

Full text
Abstract:
While deuterium kinetic isotope effects for solvolyses have been extensively studied, other nucleophilic substitutions have received less attention, and identity processes, that is, substitutions where the nucleophile and leaving group are the same, have rarely been examined. Identity reactions must pass through a truly symmetrical stage, a transition state or an intermediate, so that data will be of interest to both theoretical and experimental chemists. Values of kH/kD have been determined by polarimetry for bromide exchange– racemization at ArCHBrCH3/CD3 (Ar = C6H5,4-Br- and 4-Me-C6H4, and 3,4,-dimethyl-C6H3) in acetone, acetonitrile, and nitromethane. Observed values are analogous to values seen in solvolyses. They range from 1.01 to 1.35 and, in some cases, increase markedly as the concentration of Bu4NBr decreases. Solvolyses are either first order or pseudo first order whereas plotting observed racemization rate versus [Bu4NBr] allows separation of first- and second-order components; those species giving more stable carbocations in the more dipolar solvents, the systems showing kH/kD variation with Br− concentration, alone show an appreciable first-order component. The second-order kH/kD ratio averages 1.062 ± −0.018 at temperatures ranging from 25 to 50 °C for all substrates in the three solvents, very analogous to the values seen for racemization of 1,1,1-d3-2-bromooctane or solvolysis of ethyl substrates but considerably lower than the typical solvolysis values of 1.15–1.25 for secondary, and 1.35–1.5 for tertiary substrates. The first-order kH/kD values obtained are higher, 1.1–1.5. These and other results are discussed.
APA, Harvard, Vancouver, ISO, and other styles
30

Guiton, Theresa A., and Carlo G. Pantano. "Solution/gelation of arsenic trisulfide in amine solvents." Chemistry of Materials 1, no. 5 (September 1989): 558–63. http://dx.doi.org/10.1021/cm00005a018.

Full text
APA, Harvard, Vancouver, ISO, and other styles
31

Xin, Hong-Xing, Qi Liu, Hong Yan, and Xiu-Qing Song. "Stability and decarbonylation of 1,3-dialkyl-2-formylimidazolium perchlorate in solution." Canadian Journal of Chemistry 91, no. 6 (June 2013): 442–47. http://dx.doi.org/10.1139/cjc-2012-0497.

Full text
Abstract:
The stability of 1,3-dialkyl-2-formylimidazolium perchlorate 1 in solution was studied in detail and found to be related to its structure and the solvent character and temperature. 1 was stable in common solvents at room temperature and unstable in protic solvents under reflux. In protic solvents, such as H2O, MeOH, EtOH, and AcOH, 1 decarbonylated into 1,3-dialkylimidazole perchlorates 2, which was confirmed by 1H NMR, 13C NMR, HRMS, and X-ray spectroscopy. The decarbonylation of 1 was proposed to occur via its hemiacetal formed by the addition of solvents based on the tracking NMR spectra of 1 in deuterated reagents.
APA, Harvard, Vancouver, ISO, and other styles
32

Smith, Peter James, and Jyotsna Pradhan. "Solvolysis of 2-substituted-9-(otrho-substituted phenylmethyl)fluoren-9-yltrimethylammonium ions in various solvents. The effect of steric crowding on alkene formation." Canadian Journal of Chemistry 64, no. 6 (June 1, 1986): 1060–71. http://dx.doi.org/10.1139/v86-178.

Full text
Abstract:
The solvolytic reaction of several 9-(ortho-substituted phenylmethyl)fluoren-9-yltrimethylammonium salts has been investigated in several different solvents. Substitution and elimination products were found for the reactions in all the solvents studied, with the exceptions that reaction in both tert-butyl alcohol and chloroform led exclusively to the alkene product. The observed rate constants for alkene formation and the percent alkene were measured and it was found that the di-ortho compounds reacted at a faster rate but produced less alkene than the reaction of the corresponding mono-ortho salts. Hydrogen–deuterium isotope effects were also determined for the various reactions. The results are discussed in terms of the reaction proceeding by way of the E1 mechanism, where steric acceleration promotes the loss of the bulky ammonium leaving group to give the carbocation intermediate.
APA, Harvard, Vancouver, ISO, and other styles
33

Sikora, Antonin, and Frank E. Karasz. "Solution-phase equilibria for block copolymers in selective solvents." Macromolecules 26, no. 1 (January 1993): 177–81. http://dx.doi.org/10.1021/ma00053a027.

Full text
APA, Harvard, Vancouver, ISO, and other styles
34

Luan, Yanan, Jing Li, Musika Kaliwanda, Na Wang, Kui Chen, Xin Li, Weiyi Su, and Hongxun Hao. "Solution Thermodynamics of Benzotriazole in Different Pure Solvents." Journal of Chemical & Engineering Data 63, no. 5 (March 2, 2018): 1546–55. http://dx.doi.org/10.1021/acs.jced.7b01085.

Full text
APA, Harvard, Vancouver, ISO, and other styles
35

Kebarle, P., G. Caldwell, T. Magnera, and J. Sunner. "Ions-gas phase and solution-dipolar aprotic solvents." Pure and Applied Chemistry 57, no. 2 (January 1, 1985): 339–46. http://dx.doi.org/10.1351/pac198557020339.

Full text
APA, Harvard, Vancouver, ISO, and other styles
36

Tanaka, Yoshiaki, and Masakuni Matsuoka. "Selection of solvents for organic crystal growth from solution." Journal of Crystal Growth 99, no. 1-4 (January 1990): 1130–33. http://dx.doi.org/10.1016/s0022-0248(08)80094-2.

Full text
APA, Harvard, Vancouver, ISO, and other styles
37

Burrows, Hugh D., Maria da Graça Miguel, Rui P. C. Pereira, Nuno M. B. Proença, Sónia M. C. Cardoso, Carlos F. G. C. Geraldes, M. Helena Gil, and Wyn Brown. "Solution behaviour of lead(II) carboxylates in organic solvents." Colloids and Surfaces A: Physicochemical and Engineering Aspects 250, no. 1-3 (December 2004): 459–65. http://dx.doi.org/10.1016/j.colsurfa.2004.06.039.

Full text
APA, Harvard, Vancouver, ISO, and other styles
38

De, Tapas K., and Amarnath Maitra. "Solution behaviour of Aerosol OT in non-polar solvents." Advances in Colloid and Interface Science 59 (August 1995): 95–193. http://dx.doi.org/10.1016/0001-8686(95)80005-n.

Full text
APA, Harvard, Vancouver, ISO, and other styles
39

Herschlag, Daniel, and William P. Jencks. "Evidence that metaphosphate monoanion is not an intermediate in solvolysis reactions in aqueous solution." Journal of the American Chemical Society 111, no. 19 (September 1989): 7579–86. http://dx.doi.org/10.1021/ja00201a047.

Full text
APA, Harvard, Vancouver, ISO, and other styles
40

Figeys, Daniel, Maegorzata Koschmidder, and Robert L. Benoit. "Enthalpies of solution of naphthalene, N,N-dimethyl-1-naphthylamine, and 1,8-bis(dimethylamino)naphthalene in 16 organic solvents." Canadian Journal of Chemistry 70, no. 6 (June 1, 1992): 1586–89. http://dx.doi.org/10.1139/v92-195.

Full text
Abstract:
The enthalpies of solution of naphthalene, N,N-dimethyl-1-naphthylamine, and 1,8-bis(dimethylamino)naphthalene (proton sponge) were determined at 298.15 K in 16 organic solvents (n-hexane, cyclohexane, carbon tetrachloride, chloroform, 1,2-dichloroethane, benzene, chlorobenzene, dimethyl sulfoxide, N,N-dimethylformamide, ethyl acetate, 1,4-dioxane, anisole, nitrobenzene, benzonitrile, methanol, ethanol). Additional determinations were made with benzene. Useful linear relationships are observed between the molar enthalpies of solution of the four compounds in the solvents. The molar enthalpies of solution were correlated with the solvatochromic parameter of the solvents. The presence of N(CH3)2 groups on naphthalene does not significantly contribute to the solute–solvent interactions.
APA, Harvard, Vancouver, ISO, and other styles
41

Černý, Antonín, Jan Beneš, Jaroslav Vachek, Milan Pešák, Josef Stuchlík, Milan Stuchlík, Petr Sedmera, Miroslav Flieger, and Jindřich Vokoun. "Solvolysis of some 1-(8α-ergolinyl)-3,3-diethylureas and their salts." Collection of Czechoslovak Chemical Communications 52, no. 5 (1987): 1331–39. http://dx.doi.org/10.1135/cccc19871331.

Full text
Abstract:
Nine salts of 1-(8α-ergolinyl)-3,3-diethylurea (II) were prepared and their solubility in water and the stability of the aqueous solutions at 60 and 100 °C were studied. The main product of hydrolysis is 6-methyl-8α-aminoergoline IV. The urethan VII is formed in the ethanolic solution. Both decomposition products are also formed under long-term storage at +5 °C. The course of hydrolysis of N-propyl homologue III is similar. The decomposition of 9,10-didehydro derivative I is much slower under the conditions used.
APA, Harvard, Vancouver, ISO, and other styles
42

McManus, Samuel P., Terry Crutcher, R. Wendel Naumann, Kerry L. Tate, Steven E. Zutaut, Alan R. Katritzky, and Dennis N. Kevill. "Selectivity in the solvolysis in binary solvents of 1-adamantyl derivatives bearing leaving groups that depart as neutral molecules." Journal of Organic Chemistry 53, no. 18 (September 1988): 4401–3. http://dx.doi.org/10.1021/jo00253a041.

Full text
APA, Harvard, Vancouver, ISO, and other styles
43

Alhalaweh, Amjad, Anders Sokolowski, Naír Rodríguez-Hornedo, and Sitaram P. Velaga. "Solubility Behavior and Solution Chemistry of Indomethacin Cocrystals in Organic Solvents." Crystal Growth & Design 11, no. 9 (September 7, 2011): 3923–29. http://dx.doi.org/10.1021/cg200517r.

Full text
APA, Harvard, Vancouver, ISO, and other styles
44

Liu, Xinghuan, Junjie Kang, Yiqing Wang, Wei Li, Heling Guo, Liang Xu, Xuhong Guo, Feng Zhou, and Xin Jia. "Amine-Triggered Dopamine Polymerization: From Aqueous Solution to Organic Solvents." Macromolecular Rapid Communications 39, no. 12 (May 11, 2018): 1800160. http://dx.doi.org/10.1002/marc.201800160.

Full text
APA, Harvard, Vancouver, ISO, and other styles
45

Aragón, Diana M., Jaiver E. Rosas, and Fleming Martínez. "Solution thermodynamics of naproxen in some volatile organic solvents." Physics and Chemistry of Liquids 48, no. 4 (August 2010): 437–49. http://dx.doi.org/10.1080/00319100902894249.

Full text
APA, Harvard, Vancouver, ISO, and other styles
46

Tkadlecová, Marcela, Jaroslav Havlíček, and Vladimír Dohnal. "Association between halothane and oxygenated solvents by 1H NMR spectroscopy." Canadian Journal of Chemistry 73, no. 9 (September 1, 1995): 1406–11. http://dx.doi.org/10.1139/v95-175.

Full text
Abstract:
Using 1H NMR spectroscopy the complex-formation equilibria between halothane (2-bromo-2-chloro-1,1,1-trifluoroethane) and methyl tert-butyl ether or tetrahydrofuran in various inert solvents (hexane, heptane, decane, cyclohexane) were measured as a function of temperature. For two different association models (ideal solution and athermal solution), assuming only the formation of a 1:1 H-bonded complex, the equilibrium constants and the standard enthalpies of the complex-formation reaction were calculated. The ideal solution model provides values of the equilibrium constant that differ for different inert solvents. The athermal solution model makes this false solvent effect much smaller. For the low halothane concentration used, its dimerization was neglected. This assumption was verified experimentally. Keywords: 1H NMR, association, complex formation, halothane.
APA, Harvard, Vancouver, ISO, and other styles
47

Rakipov, Ilnaz T., Artem A. Petrov, Aydar A. Akhmadiyarov, Artashes A. Khachatrian, Timur A. Mukhametzyanov, and Boris N. Solomonov. "Thermochemistry of Solution, Solvation, and Hydrogen Bonding of Cyclic Amides in Proton Acceptor and Donor Solvents. Amide Cycle Size Effect." Molecules 26, no. 5 (March 5, 2021): 1411. http://dx.doi.org/10.3390/molecules26051411.

Full text
Abstract:
In the present work, the thermochemistry of solution, solvation, and hydrogen bonding of cyclic amides in proton acceptor (B) and proton donor (RXH) solvents were studied. The infinite dilution solution enthalpies of δ-valerolactam, N-methylvalerolactam, ε-caprolactam, and N-methylcaprolactam were measured at 298.15 K. The solvation enthalpies of cyclic amides were calculated based on the measured solution enthalpies and sublimation/vaporization enthalpies from literature. The enthalpies of hydrogen bonding between cyclic amides and proton acceptor and donor solvents were then calculated as a difference between the total solvation enthalpy and the non-specific contribution. The latter was estimated via two different approaches in proton donor and proton accepting solvents. The effect of the cycle size on the strength of hydrogen bonding of the cyclic amides in solution is discussed.
APA, Harvard, Vancouver, ISO, and other styles
48

Zechner, J., N. Getoff, and St Stoyanov. "Irreversible Anaerobic Photoreactions of Phenylazopyrazolone-Dyes in Solution." Zeitschrift für Naturforschung A 40, no. 1 (January 1, 1985): 37–42. http://dx.doi.org/10.1515/zna-1985-0108.

Full text
Abstract:
Irreversible degradation reactions of some l-phenyl-3-methyl-4-arylazopyrazole-5-ones in deoxygenated solutions were studied by flash and steady state photolysis. In addition to a reversible photoisomerization a reduction of the substrates under investigation to amino compounds was also observed. Degradation quantum yields were found to be lowest in hydrocarbon solvents. They show a pronounced wavelength dependence in all used solvents. Possible reaction mechanisms are discussed.
APA, Harvard, Vancouver, ISO, and other styles
49

Antonietti, Markus, Stephan Foerster, Mathias Zisenis, and Juergen Conrad. "Solution Viscosity of Polyelectrolyte-Surfactant Complexes: Polyelectrolyte Behavior in Nonaqueous Solvents." Macromolecules 28, no. 7 (March 1995): 2270–75. http://dx.doi.org/10.1021/ma00111a022.

Full text
APA, Harvard, Vancouver, ISO, and other styles
50

Grigoras, Anca Giorgiana, and Niculae Olaru. "Solubility Behaviour of Cellulose Acetate Butyrate in Mixture of Solvents." Revista de Chimie 68, no. 6 (July 15, 2017): 1281–84. http://dx.doi.org/10.37358/rc.17.6.5657.

Full text
Abstract:
In order to identify the proper solvent mixture for future processing from solution of cellulose derivatives, the laser light scattering method was used. The multiangle laser light scattering parameters of macromolecular chains in solution (weight-average molecular weigh , radius of gyration Rg and second virial coefficient A2) were correlated with the values of dielectric constant of solvent mixture (2-methoxyethanol + N,N-dimethylformamide). The experimental results revealed that the protic solvent favored the extended conformation of the macromolecules in solution, necessary condition to obtaine nanofibers by electrospinning.
APA, Harvard, Vancouver, ISO, and other styles
We offer discounts on all premium plans for authors whose works are included in thematic literature selections. Contact us to get a unique promo code!

To the bibliography