Journal articles on the topic 'Solvolisi'

To see the other types of publications on this topic, follow the link: Solvolisi.

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 'Solvolisi.'

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

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
2

Kašpárek, František, Kamila Vavčíková, Jiří Mollin, and Aleš Husek. "Steric effects in alkaline solvolysis of diaryl anilidophosphates." Collection of Czechoslovak Chemical Communications 55, no. 7 (1990): 1769–76. http://dx.doi.org/10.1135/cccc19901769.

Full text
Abstract:
Analysis of the products of solvolysis of aryl esters of anilidophosphoric acids in water-2-propanol mixture excludes that the reaction proceeds via E1cB mechanism. Analysis of the reaction products of solvolysis of methyl-substituted compounds made it possible to evaluate the importance of the reported steric hindrance by CH3 group. The dependence of rate constants of the solvolysis on solvent system has been established.
APA, Harvard, Vancouver, ISO, and other styles
3

D'Souza, Malcolm J., Anthony M. Darrington, and Dennis N. Kevill. "On the Importance of the Aromatic Ring Parameter in Studies of the Solvolyses of Cinnamyl and Cinnamoyl Halides." Organic Chemistry International 2010 (June 29, 2010): 1–9. http://dx.doi.org/10.1155/2010/130506.

Full text
Abstract:
In solvolysis studies using Grunwald-Winstein plots, dispersions were observed for substrates with aromatic rings at the α-carbon. Several examples for the unimolecular solvolysis of monoaryl benzylic derivatives and related diaryl- or naphthyl-substituted derivatives have now been reported, where the application of the aromatic ring parameter (I) removes this dispersion. A recent claim suggesting the presence of an appreciable nucleophilic component to the I scale has now been shown, in a review of the solvolysis of highly-hindered alkyl halides, to be unlikely to be correct. Attention is now focused on the application of the hI term for the solvolysis of compounds containing a double bond in the vicinity of any developing carbocation. Available specific rates of solvolysis (plus some new values) at 25°C of cinnamyl chloride, cinnamyl bromide, cinnamoyl chloride, p-chlorocinnamoyl chloride, and p-nitrocinnamoyl chloride are analyzed using the simple and extended (including the hI term) Grunwald-Winstein equations.
APA, Harvard, Vancouver, ISO, and other styles
4

Kumaniaev, Ivan, Elena Subbotina, Maxim V. Galkin, Pemikar Srifa, Susanna Monti, Isara Mongkolpichayarak, Duangamol Nuntasri Tungasmita, and Joseph S. M. Samec. "A combination of experimental and computational methods to study the reactions during a Lignin-First approach." Pure and Applied Chemistry 92, no. 4 (April 28, 2020): 631–39. http://dx.doi.org/10.1515/pac-2019-1002.

Full text
Abstract:
AbstractCurrent pulping technologies only valorize the cellulosic fiber giving total yields from biomass below 50 %. Catalytic fractionation enables valorization of both cellulose, lignin, and, optionally, also the hemicellulose. The process consists of two operations occurring in one pot: (1) solvolysis to separate lignin and hemicellulose from cellulose, and (2) transition metal catalyzed reactions to depolymerize lignin and to stabilized monophenolic products. In this article, new insights into the roles of the solvolysis step as well as the operation of the transition metal catalyst are given. By separating the solvolysis and transition metal catalyzed hydrogen transfer reactions in space and time by applying a flow-through set-up, we have been able to study the solvolysis and transition metal catalyzed reactions separately. Interestingly, the solvolysis generates a high amount of monophenolic compounds by pealing off the end groups from the lignin polymer and the main role of the transition metal catalyst is to stabilize these monomers by transfer hydrogenation/hydrogenolysis reactions. The experimental data from the transition metal catalyzed transfer hydrogenation/hydrogenolysis reactions was supported by molecular dynamics simulations using ReaXFF.
APA, Harvard, Vancouver, ISO, and other styles
5

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
6

Pytela, Oldřich, Stanislava Štumrová, Miroslav Ludwig, and Miroslav Večeřa. "Kinetic acidity function and solvolysis of 3-hydroxy-1,3-diphenyltriazenes." Collection of Czechoslovak Chemical Communications 51, no. 3 (1986): 564–72. http://dx.doi.org/10.1135/cccc19860564.

Full text
Abstract:
Ten 3-hydroxy-1-(X-phenyl)-3-phenyltriazines have been synthesized, and kinetics of their solvolysis have been measured in 40% (v/v) ethanol and sulphuric acid. The concept of kinetic acidity function has been generalized, its construction has been suggested, and the procedure has been applied to the solvolysis of 3-hydroxy-1,3-diphenyltriazenes. The kinetic acidity function found has been confronted with the H0 acidity function. The substituent effects have been evaluated with respect to mechanism of the acid catalyzed solvolysis.
APA, Harvard, Vancouver, ISO, and other styles
7

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
8

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
9

Conner, John K., Johanna Haider, MN Stuart Hill, Howard Maskill, and Monique Pestman. "The mechanism of solvolysis of 2-adamantyl azoxytosylate: isotopic labelling, medium effect, and attempted deoxygenation studies." Canadian Journal of Chemistry 76, no. 6 (June 1, 1998): 862–68. http://dx.doi.org/10.1139/v98-071.

Full text
Abstract:
Rates of solvolysis of 2-adamantyl azoxytosylate (1) have been measured over a range of temperatures in ethanoic acid, methanoic acid, 50:50 (v/v) trifluoroethanol:water, 80:20 (v/v) trifluoroethanol:water, 97:3 (w/w) trifluoroethanol:water, and 70:30 (v/v) ethanol:water. For comparison, rates of solvolysis of 2-adamantyl tosylate (2) have also been measured in 50:50, 80:20, and 90:10 (v/v) trifluoroethanol:water, and for both compounds, activation parameters have been determined. These and results published earlier allow a correlation of the two reactions and indicate that the m value for 2-adamantyl azoxytosylate solvolysis is only 0.46. This is one of the lowest m values for a reaction that is unambiguously an SN1 solvolysis. We have also recorded the 17O NMR spectrum of the 2-adamantyl tosylate formed from 17O-labelled 2-adamantyl azoxytosylate in deuteriochloroform, and the millimeter-wave spectrum of the nitrous oxide evolved in the hydrolysis of 18O-labelled 2-adamantyl azoxytosylate. These labelling studies have provided more detailed knowledge of the SN1 fragmentation mechanism of 1 and exclude a mechanism of reaction via rearrangement to N-nitroso,N-(2-adamantyl),O-(p-toluenesulfonyl)hydroxylamine (5). Attempted deoxygenation of 1 to give 2-adamantyl diazotosylate (8) and a subsequent fragmentation proved unsuccessful.Key words: nitrous oxide, carbenium ion, isotopic labelling, solvolysis, m value.
APA, Harvard, Vancouver, ISO, and other styles
10

Omura, Kanji. "Solvolysis of 4-Halogeno-4-Alkyl-2,6-di-tert-butylcyclohexa-2,5-dienones Induced by Positive Halogen Donors as Electrophiles." Australian Journal of Chemistry 66, no. 11 (2013): 1386. http://dx.doi.org/10.1071/ch13257.

Full text
Abstract:
Positive halogen donors such as N-iodosuccinimide (NIS) induce solvolysis of dienones 1, as model 4-halogenocyclohexa-2,5-dienones, in different hydroxylic solvents (ROH), yielding the 4-RO-cyclohexa-2,5-dienones (2). The rate of the solvolysis with NIS is highly dependent on the structure of ROH. The problem of such dependency is overcome by running the reaction in ROH diluted with MeCN, a polar aprotic solvent, in place of pure ROH; the rate of the reaction in the ROH-MeCN solvent mixture is almost independent of the structure (or the polarity) of ROH, and the reaction is completed faster or markedly faster than in neat ROH. The results suggest that the solvolysis rate is controlled by the polarity of the solvent system, although the hydrogen-bond acceptability of MeCN for dilution also accelerates the reaction. A mechanism for the solvolysis is proposed, involving electrophilic attack of a positive halogen donor at the halogen atom of 1, generating the 4-oxocyclohexa-2,5-dienyl cation intermediates (8) via the rate-limiting polar transition states.
APA, Harvard, Vancouver, ISO, and other styles
11

Bremner, JB, and KN Winzenberg. "Cyanogen-Bromide-Mediated and Methyl-Chloroformate-Mediated Synthesis of Some [3]Benzazonino[8,7,6-abc]Carbazole, [3]Benzazecino[9,8,7-abc]Carbazole, Naphth[1,8,7-def]Azonine and Naphth[1,8,7-def]Azecine Derivatives." Australian Journal of Chemistry 39, no. 1 (1986): 1. http://dx.doi.org/10.1071/ch9860001.

Full text
Abstract:
Reaction of 7,8-dimethoxy-1,4,5,9,13c,13d-hexahydro-2H-indolo [3′,2′:4,5] indolo [1,7,6-aji] isoquinoline (4a) and 8,9-dimethoxy- 1,2,3,5,6,10,14c,14d-octahydro-indolo[3′,2′:3,4] naphtho [2,1,8-ija] quinolizine (4b) with cyanogen bromide in the presence of potassium carbonate afforded the elimination products 7,8-dimethoxy-1,2,3,4,5,9- hexahydro [3] benzazonino [8,7,6-abc]carbazole-3-carbonitrile (5a) and 8,9-dimethoxy-2,3,4,5,6,10-hexahydro-1H-[3] benzazecino [9,8,7-abc]carbazole-4-carbonitrile (5b) in 4% and 88% yield respectively; 7- (2-bromo)ethyl-1,2-dimethoxy-4,5,6,6a,7,12-hexahydroisoquino[8,1-ab ]carbazole-6-carbonitrile (6a) was also isolated in 20% yield from the former reaction. Products of analogous structure were also obtained from (4a,b) when methyl chloroformate was used in place of cyanogen bromide. Reaction of (4b) with cyanogen bromide in the presence of water and magnesium oxide afforded the solvolysis product 14d-hydroxy-8,9-dimethoxy-2,3,4,5,6,10,14d,14d-octahydro-1H-[3] benzazecino [9,8,7- abc ]carbazole-4-carbonitrile (8b) together with (5b). Similarly application of this reaction to 8,9-dimethoxy-2,3,5,6,11a,11b-hexahydro-1H-naphtho[2,1,8-ija] quinolizine (1b) afforded both the corresponding solvolysis and elimination products; (4a) and 7,8- dimethoxy-1,2,4,5,10a,10b-hexahydronaphtho[1,8,7-ghi] indolizine (1a), however, gave only solvolysis products. Solvolysis and elimination products were also isolated when (1b) and (4b) were subjected to cyanogen -bromide-mediated methanolysis reactions. Acid-catalysed elimination reactions of the alcohol solvolysis products are described along with reaction of some of the elimination products with lithium tetrahydroaluminate.
APA, Harvard, Vancouver, ISO, and other styles
12

Bolton, R., RE Burley, and NJ Williams. "Stabilities of Carbonium-Ions. IV. Steric Effects in the Solvolysis of Substituted Diphenylmethyl Chlorides." Australian Journal of Chemistry 39, no. 4 (1986): 625. http://dx.doi.org/10.1071/ch9860625.

Full text
Abstract:
The replacement of ortho-hydrogen atoms by methyl groups in diphenylmethyl chloride has three distinguishable results upon the rate of solvolysis . Firstly, the alkyl groupactivates by its electronic effect; secondly, steric interactions diminish all observed substituent effects regardless of the position of the substituent in the aryl system; and thirdly, steric acceleration of the solvolysis can be seen in the rate of reaction of bis (2,6-dimethylphenyl)methyl chloride. The ortho-methyl substituents inhibit the formation of the planar transition state necessary to allow the greatest resonance stabilization by the aryl substituents of the incipient carbocation. Greater degrees of twist are reflected both in the variations in the rates of solvolysis of the poly(orthomethyl ) diphenylmethyl chlorides and in the consistent fall in the value of p+ with successive ortho-substitution.
APA, Harvard, Vancouver, ISO, and other styles
13

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
14

Kaválek, Jaromír, Josef Jirman, Vladimír Macháček, and Vojeslav Štěrba. "The methanolysis kinetics and dissociation constants of 1-(subst. benzoyl)-3-phenylthioureas." Collection of Czechoslovak Chemical Communications 53, no. 3 (1988): 593–600. http://dx.doi.org/10.1135/cccc19880593.

Full text
Abstract:
A series of seven 1-(subst. benzoyl)-3-phenylthioureas have been prepared and their dissociation constants and solvolysis rate constants have been measured in methanol at 25 °C. The reaction constants found show that the solvolysis rate is limited by the attack of methoxide ion on the benzoyl carbonyl group of the non-dissociated substrate. The polar effect of substituents in benzoyl group is extensively transferred also by the intramolecular hydrogen bond.
APA, Harvard, Vancouver, ISO, and other styles
15

D’Souza, Malcolm J., and Dennis N. Kevill. "Mechanistic studies of the solvolysis of alkanesulfonyl and arenesulfonyl halides." Beilstein Journal of Organic Chemistry 18 (January 17, 2022): 120–32. http://dx.doi.org/10.3762/bjoc.18.13.

Full text
Abstract:
There have been several studies on the solvolysis mechanisms for alkanesulfonyl chlorides (RSO2Cl) and arenesulfonyl chlorides (ArSO2Cl). The earlier of these studies were reviewed a little over thirty years ago by Gordon, Maskill and Ruasse (Chem. Soc. Rev. 1989, 18, 123–151) in a contribution entitled “Sulfonyl Transfer Reactions”. The present review will emphasize more recent contributions and, in particular, the application of the extended Grunwald–Winstein equation and kinetic solvent isotope effects to the solvolysis reactions. There is also an appreciable number of reports concerning the corresponding anhydrides with the chloride leaving group replaced by the appropriate sulfonate leaving group, concerning sulfamoyl chlorides (ZZ'NSO2Cl) with Z and Z' being alkyl or aryl and concerning the solvolysis of chlorosulfate esters (alkoxy- or aryloxysulfonyl chlorides), with the structures ROSO2Cl or ArOSO2Cl. The solvolyses of these additional types of sulfur(VI) substrates will be the topics of a future review.
APA, Harvard, Vancouver, ISO, and other styles
16

SANO, YOSHIHIRO. "Solvolysis Pulping." Sen'i Gakkaishi 43, no. 8 (1987): P306—P311. http://dx.doi.org/10.2115/fiber.43.8_p306.

Full text
APA, Harvard, Vancouver, ISO, and other styles
17

Kooduvalli, Komal, John Unser, Soydan Ozcan, and Uday K. Vaidya. "Embodied Energy in Pyrolysis and Solvolysis Approaches to Recycling for Carbon Fiber-Epoxy Reinforced Composite Waste Streams." Recycling 7, no. 1 (February 14, 2022): 6. http://dx.doi.org/10.3390/recycling7010006.

Full text
Abstract:
Carbon fiber composites are increasingly used in aerospace, motorcycles, sporting, and high-performance vehicles, and their end of life recycling is of growing interest. This study deals with the life cycle assessment (LCA) of carbon fiber reinforced plastics (CFRP) waste streams. The embodied energy (EE) of recycling CFRP via two viable methods—i.e., pyrolysis and solvolysis—is studied. Both pyrolysis and solvolysis were studied for EE with different variants. Alongside fiber recovery from CFRP, the pyrolysis process calculations consider energy recovery from syngas and oil produced within the system. For pyrolysis, electric furnace and natural gas were primarily considered. For solvolysis, different solvent scenarios were considered, including (a) deionized water, (b) water and potassium hydroxide, (c) acetone and water, and (d) water with acetic acid and potassium hydroxide. Energy reduction from one generation to the next has also been highlighted. The EE for recycling CFRP is quantified and discussed for these scenarios in this paper.
APA, Harvard, Vancouver, ISO, and other styles
18

Prasad, Pritesh, Angela A. Salim, Shamsunnahar Khushi, Zeinab G. Khalil, Michelle Quezada, and Robert J. Capon. "Solvolysis Artifacts: Leucettazoles as Cryptic Macrocyclic Alkaloid Dimers from a Southern Australian Marine Sponge, Leucetta sp." Marine Drugs 17, no. 2 (February 9, 2019): 106. http://dx.doi.org/10.3390/md17020106.

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

Fu, Lianshe, Rute A. Sá Ferreira, Sonia S. Nobre, Luís D. Carlos, and João Rocha. "Organically Modified Silica-Based Xerogels Derived from 3-Aminopropyltrimethoxysilane and 3-Isocyanatepropyltriethoxysilane through Carboxylic Acid Solvolysis." Materials Science Forum 514-516 (May 2006): 108–12. http://dx.doi.org/10.4028/www.scientific.net/msf.514-516.108.

Full text
Abstract:
Organically-modified silica xerogels from 3-aminopropyltrimethoxysilane (APTES) and 3-isocyanatepropyltriethoxysilane (ICPTES) have been synthesized through carboxylic acid (formic acid, acetic acid and valeric acid) solvolysis. The resulting hybrid materials have been characterized by powder X-ray diffraction, mid-infrared spectroscopy, 29Si and 13C nuclear magnetic resonance, and photoluminescence spectroscopy. The results show that urea cross-links have been formed in these hybrids. The luminescence features depend on the selected carboxylic acids. For example, comparatively to the hybrids derived from formic and acetic acid solvolysis, valeric acid shows a red-shift of the emission features.
APA, Harvard, Vancouver, ISO, and other styles
20

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
21

Lee, Choi Chuck, and Charles Y. Fiakpui. "Some isotopic scrambling studies with singly or doubly labeled triphenylvinyl bromide." Canadian Journal of Chemistry 63, no. 3 (March 1, 1985): 681–84. http://dx.doi.org/10.1139/v85-112.

Full text
Abstract:
The solvolysis of triphenyl[2-14C]vinyl bromide (1-Br-2-14C) in 70% HOAc – 30% H2O or in 2,2,2-trifluoroethanol (TFE) carried out in the presence of an excess of p-CH3C6H4SNa gave the triphenylvinyl p-tolyl thioether (1-STol) with a greatly decreased extent of scrambling of the 14C-label from C-2 to C-1 when compared with analogous reactions without the presence of the p-toluenethiolate anion. For example, the 1-STol product from the reaction of 1-Br-2-14C in 70% HOAc – p-CH3C6H4SNa showed only 0.5–0.9% scrambling, while previous studies on the reaction of 1-Br-2-14C in 70% HOAc – NaOAc gave 14.7 ± 0.7% scrambling. Previous work on the solvolysis of 1,2-diphenyl-2-[2H5]phenyl[2-13C]vinyl bromide (1-Br-2-13C-2-Ph-d5 in 70% HOAc, as well as the present results from the solvolysis of 1-Br-2-13C-2-Ph-d5 in TFE – 2,6-lutidine, gave products derived from all 4 possible isotopomeric triphenylvinyi cations 3, 4, 5, and 6 arising from successive 1,2-phenyl shifts. On the other hand, solvolysis of 1-Br-2-13C-2-Ph-d5 in 70% HOAc or TFE containing an excess of p-CH3C6H4SNa gave 1-STol in which the isotopically rearranged product was derived only from ion 5 arising from a 1,2-shift of the unlabeled phenyl group with no detectable amount of production derived from a 1,2-shift of the perdeuterophenyl group. These results are interpreted by the rapid trapping of the triphenylvinyi cation by the highly nucleophilic p-toluenethiolate anion.
APA, Harvard, Vancouver, ISO, and other styles
22

Browne, EJ. "Synthesis of 1H-[1]Benzothieno[3,2-d]azonine and [1]Benzothieno[3,2-d]azecine Derivatives." Australian Journal of Chemistry 38, no. 5 (1985): 765. http://dx.doi.org/10.1071/ch9850765.

Full text
Abstract:
Derivatives of two new [1] benzothieno medium-ring heterocyclic systems have been prepared by ring degradation using cyanogen bromide-induced solvolysis of tetracyclic precursors. Reaction of a hexahydro -[1] benzothieno [3,2-g] indolizine (4a) and a hexahydro-2H-[1] benzo-thieno [2,3-a] quinolizine (4b) with cyanogen bromide and magnesium oxide under solvolytic conditions yielded the hexahydro-1H-[1] benzothieno [3,2-d] azonines (5a) and (6a) and the octahydro -[1] benzothieno [3,2-d] azecines (5b) and (6b), respectively. Functional group interconversions of these medium-ring systems are described, including oxidations to the cyclic ketones (7) and (9). The 11b-phenyl derivative (13) of (4a) reacted under similar conditions to give both solvolysis (14) and (16) and elimination (15) medium-ring products, the ratios depending on the solvent. By contrast the analogous 9a-phenylthienoindolizine derivative (17) under these conditions gave only the medium-ring elimination product (18) in aqueous medium, and only the equivalent solvolysis product in methanol. Both the thieno and [1] benzothienoazonine elimination products (18) and (15) appear to be mixtures of E and Z isomers. The [1] benzothieno [3,2-g] indolizine bases (4a) and (13) are the first reported examples of this ring system.
APA, Harvard, Vancouver, ISO, and other styles
23

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
24

Nielsen, J. B., A. Jensen, C. B. Schandel, C. Felby, and A. D. Jensen. "Solvent consumption in non-catalytic alcohol solvolysis of biorefinery lignin." Sustainable Energy Fuels 1, no. 9 (2017): 2006–15. http://dx.doi.org/10.1039/c7se00381a.

Full text
APA, Harvard, Vancouver, ISO, and other styles
25

Figueirêdo, M. B., H. J. Heeres, and P. J. Deuss. "Ozone mediated depolymerization and solvolysis of technical lignins under ambient conditions in ethanol." Sustainable Energy & Fuels 4, no. 1 (2020): 265–76. http://dx.doi.org/10.1039/c9se00740g.

Full text
APA, Harvard, Vancouver, ISO, and other styles
26

D'Souza, Malcolm J., Anthony M. Darrington, and Dennis N. Kevill. "A Study of Solvent Effects in the Solvolysis of Propargyl Chloroformate." ISRN Organic Chemistry 2011 (April 11, 2011): 1–6. http://dx.doi.org/10.5402/2011/767141.

Full text
Abstract:
The specific rates of solvolysis of propargyl chloroformate (1) are analyzed in 22 solvents of widely varying nucleophilicity and ionizing power values at 25.0°C using the extended Grunwald-Winstein equation. Sensitivities to solvent nucleophilicity (l) of 1.37 and to solvent ionizing power (m) of 0.47 suggest a bimolecular process with the formation of a tetrahedral intermediate. A plot of the rates of solvolysis of 1 against those previously reported for phenyl chloroformate (2) results in a correlation coefficient (R) of 0.996, a slope of 0.86, and an F-test value of 2161. The unequivocal correlation between these two substrates attests that the stepwise association-dissociation (AN + DN) mechanism previously proposed for 2 is also operative in 1.
APA, Harvard, Vancouver, ISO, and other styles
27

Fowless, AD, GA Lawrance, DR Stranks, TR Sullivan, and N. Vanderhoek. "The Effect of Pressure on the Anation of Diaquabis(ethane-1,2-diamine)cobalt(III) by Selenate and Sulfate, and on the Isomerization of the trans-Aquabis(ethane-1,2-diamine)(hydrogenselenito)cobalt(III) Ion." Australian Journal of Chemistry 41, no. 9 (1988): 1263. http://dx.doi.org/10.1071/ch9881263.

Full text
Abstract:
The reversible nucleophilic substitution of cis-[Co(en)2(OH2)2]3+ (en = ethane-1,2-diamine) by sulfate and selenate at pH 2.0 was studied at elevated pressure, and activation volumes (ΔV‡) for anation and solvolysis were determined. The ΔV‡ values for anation by sulfate and selenate were +8.3�0.5 and +8.5�0.4 cm3 mol-1 respectively, whereas solvolysis of sulfato and selenato complexes cis -[Co(en)2(OH2)(XO4)]+ gave ΔV‡ values of -2.2�0.4 and -3.3�0.7 cm3 mol-1 respectively. The trans → cis [Co(en)2(OH2)(OSeO2H)]2+ isomerization studied at pH 1.0 and elevated pressure yielded ΔV‡ + 7.2�0.4 cm3 mol-1. All reactions can be interpreted in terms of dissociative mechanisms.
APA, Harvard, Vancouver, ISO, and other styles
28

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
29

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
30

Brandner, David G., Jacob S. Kruger, Nicholas E. Thornburg, Gregory G. Facas, Jacob K. Kenny, Reagan J. Dreiling, Ana Rita C. Morais, et al. "Correction: Flow-through solvolysis enables production of native-like lignin from biomass." Green Chemistry 23, no. 24 (2021): 10168–70. http://dx.doi.org/10.1039/d1gc90119b.

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

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
32

Hirayama, Yusaku, Kyohei Kanomata, Mayumi Hatakeyama, and Takuya Kitaoka. "Chitosan nanofiber-catalyzed highly selective Knoevenagel condensation in aqueous methanol." RSC Advances 10, no. 45 (2020): 26771–76. http://dx.doi.org/10.1039/d0ra02757j.

Full text
Abstract:
Chitosan nanofibers bearing abundant and accessible amines exposed on the solid surface catalyze a highly selective Knoevenagel condensation in green solvent, which completely avoids the formation of solvolysis byproducts.
APA, Harvard, Vancouver, ISO, and other styles
33

Qin, Qi, Youwei Xie, and Paul E. Floreancig. "Diarylmethane synthesis through Re2O7-catalyzed bimolecular dehydrative Friedel–Crafts reactions." Chemical Science 9, no. 45 (2018): 8528–34. http://dx.doi.org/10.1039/c8sc03570a.

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

Protsenko, A. E., and V. V. Petrov. "Recycling of the polymer composite fillers in amino alcohol medium." Journal of Physics: Conference Series 2353, no. 1 (October 1, 2022): 012009. http://dx.doi.org/10.1088/1742-6596/2353/1/012009.

Full text
Abstract:
Abstract The paper presents data from an experimental study of the possibility of recycling polymer composites on the example of fiberglass and different thermoset binders. The goal is to remove the polymer matrix and obtain recovered fiberglass that is not inferior in properties to the virgin material. In this paper, considered the solvolysis method of the composite in the amino alcohols. As the medium used methyldiethanolamine and 3-aminopropanol-1. The dependence of the influence of the solvolysis time and the concentration of alkali metal hydroxide in the reaction medium on the strength of recovered glass fibers has been established. The solvolysis at a temperature of 180 °C and a catalyst concentration of 5 % for 6 hours was accepted as optimal regime. As a result of the research, a regime was obtained that allows recovering fibers with a tensile strength of 92 % of virgin fibers. The recovered fabrics were studied by thermal analysis and scanning electron microscopy. GFRP samples were obtained from the recovered fabrics by the VaRTM method. The flexural strength of composites made from secondary filler is 8.5% lower than the same material based on virgin fabric. Products of the polymer matrix degradation and fibrous fillers are released during recycling. These products could be used in chemical technology to produce, for example, binders components and secondary fillers for polymer composites.
APA, Harvard, Vancouver, ISO, and other styles
35

Čapek, Karel, Jindra Čapková, Jiří Jarý, Yurii A. Knirel, and Alexander S. Shashkov. "Preparation of methyl 2,4-diacetamido-2,4,6-trideoxy-α-D-ido-, α-D-talo-, α-D-altro-, and α-D-mannopyranoside." Collection of Czechoslovak Chemical Communications 52, no. 9 (1987): 2248–59. http://dx.doi.org/10.1135/cccc19872248.

Full text
Abstract:
Title compounds were prepared from methyl 2-O-acetyl-3,4-anhydro-6-deoxy-α-D-galactopyranoside (I). Solvolysis of I afforded methyl 3-O-acetyl-6-deoxy-α-D-gulopyranoside (II) from which in turn was prepared its 2,4-di-O-methanesulfonyl derivative III. The reaction of III with sodium azide undergoes under participation of the O-acetyl group in the position 3 and produces methyl 2,4-diazido-2,4,6-trideoxy-α-D-idopyranoside (V) that after hydrogenation and acetylation yields methyl 2,4-diacetamido-2,4,6-trideoxy-α-D-idopyranoside (VIII). Methyl 2,4-diacetamido-2,4,6-trideoxy-α-D-talopyranoside (XI) was prepared by its mesylation and by solvolysis of the product. Azidolysis of I, followed by hydrogenation and acetylation gave methyl 4-acetamido-4,6-dideoxy-α-D-glucopyranoside (XII). Its subsequent reaction with diethyl azodicarboxylate-triphenylphosphine mixture led to methyl 4-acetamido-2,3-anhydro-4,6-dideoxy-α-D-allopyranoside (XV). Mixture of methyl 4-acetamido-2-azido-2,4,6-trideoxy-α-D-altropyranoside (XVII) and its gluco- isomer XVIII is formed in the reaction of XV with sodium azide. Hydrogenation of XVII yields amino derivative XIX that in turn gives methyl 2,4-diacetamido-2,4,6-trideoxy-α-D-altropyranoside (XXI) by acetylation. Methyl 2,4-diacetamido-2,4,6-trideoxy-α-D-mannopyranoside (XXII) was prepared by its mesylation and solvolysis. All structures were confirmed by 1H and 13C NMR spectra.
APA, Harvard, Vancouver, ISO, and other styles
36

ADACHI, Yoshio, and Hideharu HIROSUE. "Solvolysis of Coal by Hydrogenated Solvent VI. Solvolysis of Canadian Coals." Journal of the Fuel Society of Japan 70, no. 10 (1991): 970–77. http://dx.doi.org/10.3775/jie.70.10_970.

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

Hieta, Kaoru, Makoto Wakai, Miyuki Kojima, Keiji Miyazaki, Yuzo Iwanaga, Takao Matsushita, and Hiroshi Tsuchiya. "Research and developement of solvolysis pulping. (II). Washing of solvolysis pulp." JAPAN TAPPI JOURNAL 43, no. 10 (1989): 1031–39. http://dx.doi.org/10.2524/jtappij.43.1031.

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

Brandner, David G., Jacob S. Kruger, Nicholas E. Thornburg, Gregory G. Facas, Jacob K. Kenny, Reagan J. Dreiling, Ana Rita C. Morais, et al. "Flow-through solvolysis enables production of native-like lignin from biomass." Green Chemistry 23, no. 15 (2021): 5437–41. http://dx.doi.org/10.1039/d1gc01591e.

Full text
Abstract:
Flow-through solvolysis offers an opportunity to limit lignin condensation reactions that prevent isolation of native lignin in biomass processing, thus allowing for the study of intrinsic lignin properties and steady-state lignin depolymerization.
APA, Harvard, Vancouver, ISO, and other styles
39

Kudavalli, Jaya S., and Rory A. More O'Ferrall. "β-Hydroxy carbocation intermediates in solvolyses of di- and tetra-hydronaphthalene substrates." Beilstein Journal of Organic Chemistry 6 (November 3, 2010): 1035–42. http://dx.doi.org/10.3762/bjoc.6.118.

Full text
Abstract:
Solvolysis of trichloroacetate esters of 2-methoxy-1,2-dihydro-1-naphthols shows a remarkably large difference in rates between the cis and trans isomers, k cis /k trans = 1800 in aqueous acetonitrile. This mirrors the behaviour of the acid-catalysed dehydration of cis- and trans-naphthalene-1,2-dihydrodiols to form 2-naphthol, for which k cis /k trans = 440, but contrasts with that for solvolysis of tetrahydronaphthalene substrates, 1-chloro-2-hydroxy-1,2,3,4-tetrahydronaphthalenes, for which k cis /k trans = 0.5. Evidence is presented showing that the trans isomer of the dihydro substrates reacts unusually slowly rather than the cis isomer unusually rapidly. Comparison of rates of solvolysis of 1-chloro-1,2,3,4-tetrahydronaphthalene and the corresponding (cis) substrate with a 2-hydroxy group indicates that a β-OH slows the reaction by nearly 2000-fold, which represents a typical inductive effect characteristic also of cis-dihydrodiol substrates. The slow reaction of the trans-dihydrodiol substrate is consistent with initial formation of a β-hydroxynaphthalenium carbocation with a conformation in which a C–OH occupies an axial position β to the carbocation centre preventing stabilisation of the carbocation by C–H hyperconjugation, which would occur in the conformation initially formed from the cis isomer. It is suggested that C–H hyperconjugation is particularly pronounced for a β-hydroxynaphthalenium ion intermediate because the stability of its no-bond resonance structure reflects the presence of an aromatic naphthol structure.
APA, Harvard, Vancouver, ISO, and other styles
40

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
41

Ferber, PH, GE Gream, and TI Stoneman. "The 9-Decalyl and Related cations VII. Solvolysis of 3-(Cyclohex-1′-enyloxy)propyl p-Nitrobenzenesulfonate." Australian Journal of Chemistry 38, no. 5 (1985): 699. http://dx.doi.org/10.1071/ch9850699.

Full text
Abstract:
The solvolysis of 3-(cyclohex-1′-enyloxy) propyl p- nitrobenzenesulfonate (5) in ethanol buffered separately with sodium ethoxide and triethylamine and 2,2,2-trifluoroethanol buffered with triethylamine has been investigated. Kinetic determinations and product studies have been carried out. In ethanol buffered with sodium ethoxide , π-bond participation in the above ester occurs to the extent of 30%; this is raised to 84% when triethylamine is used as the buffering agent. With buffered trifluoroethanol as solvent, π-bond participation in the ester is complete; kunsat/ksat = 920 and a quantitative yield of cyclized products is obtained. Kinetic evidence indicates a lack of significant involvement of a lone pair on oxygen (enol ether system) in the solvolysis of the sulfonate ester; in trifluoroethanol , the compound solvolyses only 1.15 times more rapidly than does 4-(cyclohex-1′-enyl)butyl p-nitrobenzenesulfonate (2), its carbon analogue.
APA, Harvard, Vancouver, ISO, and other styles
42

Huang, Xicai, and Andrew J. Bennet. "Kinetic evidence for the importance of solvent-separated ion pairs during the solvolyses of adamantylideneadamantyl derivatives." Canadian Journal of Chemistry 82, no. 9 (September 1, 2004): 1336–40. http://dx.doi.org/10.1139/v04-101.

Full text
Abstract:
The aqueous ethanolysis reactions of adamantylideneadamantyl tosylate, -bromide, and -iodide (1-OTs, 1-Br and 1-I) were monitored as a function of ionic strength. Special salt effects are observed during the solvolyses of both homoallylic halides, but not in the case of the tosylate 1-OTs. The measured α-secondary deuterium kinetic isotope effects for the solvolysis of 1-Br in 80:20 and 60:40 v/v ethanol–water mixtures at 25 °C are 1.110 ± 0.018 and 1.146 ± 0.009, respectively. The above results are consistent with the homoallylic halides reacting via a virtual transition state in which both formation and dissociation of a solvent-separated ion pair are partially rate-determining. While the corresponding transition state for adamantylideneadamantyl tosylate involves formation of the solvent-separated ion pair.Key words: salt effects, kinetic isotope effect, internal return, solvolysis, ion pairs.
APA, Harvard, Vancouver, ISO, and other styles
43

Wang, Guoqiang, Chuanjun Wang, Hao Zhang, Youle Liu, and Jing Xu. "Facile preparation of Cu–Fe oxide nanoplates for ammonia borane decomposition and tandem nitroarene hydrogenation." RSC Advances 11, no. 48 (2021): 29920–24. http://dx.doi.org/10.1039/d1ra04175d.

Full text
Abstract:
CuFe oxide nanoplates were developed which displayed good activity and stability toward AB solvolysis. The material transformed to Cu/Fe2O3 while catalyzing AB reduction and in a tandem system reduced nitro compounds to amines in substantial yield.
APA, Harvard, Vancouver, ISO, and other styles
44

Sano, Yoshihiro, Takashi Sasaya, and Akira Sakakibara. "Solvolysis pulping of hardwoods." JAPAN TAPPI JOURNAL 42, no. 5 (1988): 487–96. http://dx.doi.org/10.2524/jtappij.42.487.

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

Na, Younghwa. "Solvolysis Study of Cycliciminomitomycins." CHEMICAL & PHARMACEUTICAL BULLETIN 55, no. 3 (2007): 482–87. http://dx.doi.org/10.1248/cpb.55.482.

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

Mayer, Roland, Peter Schönfeld, Horst Viola, and Jürgen Fabian. "Zur Solvolyse von Thiocarbonsäurechloriden." Zeitschrift für Chemie 15, no. 11 (September 1, 2010): 443–45. http://dx.doi.org/10.1002/zfch.19750151110.

Full text
APA, Harvard, Vancouver, ISO, and other styles
47

Al-Talib, Mahmoud, and Hasan Tashtoush. "Solvolysis ofN-(Chlorosulfinyl)diarylketimines." Liebigs Annalen der Chemie 1990, no. 6 (June 12, 1990): 611–12. http://dx.doi.org/10.1002/jlac.1990199001114.

Full text
APA, Harvard, Vancouver, ISO, and other styles
48

Nateghi, Bahareh, Ishtvan Boldog, Konstiantyn V. Domasevitch, and Christoph Janiak. "More versatility than thought: large {Zr26} oxocarboxylate cluster by corner-sharing of standard octahedral subunits." CrystEngComm 20, no. 35 (2018): 5132–36. http://dx.doi.org/10.1039/c8ce01064a.

Full text
APA, Harvard, Vancouver, ISO, and other styles
49

Nakamura, Tetsuji, Hiroshi Tsuchiya, and Takeo Nagasawa. "Research and development of solvolysis pulping. (V). Process engineering for solvolysis pulping." JAPAN TAPPI JOURNAL 44, no. 2 (1990): 235–41. http://dx.doi.org/10.2524/jtappij.44.235.

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

Takagi, Hitoshi, Shogo Kachi, Haruhiko Kawabata, and Peter Sandström. "Research and development of solvolysis pulping. Part 4. Bleaching of solvolysis pulp." JAPAN TAPPI JOURNAL 43, no. 12 (1989): 1290–97. http://dx.doi.org/10.2524/jtappij.43.1290.

Full text
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