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

Journal articles on the topic 'Decarboxylation reaction'

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 'Decarboxylation reaction.'

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

Zhan, Desheng, Gang Yang, Tieli Zhou, Sashirekha Nallapati, and Xiaofeng Zhang. "Decarboxylation-Driven Double Annulations: Innovative Multi-Component Reaction Pathways." Molecules 30, no. 7 (2025): 1594. https://doi.org/10.3390/molecules30071594.

Full text
Abstract:
A concerted five-component reaction strategy has been developed, featuring double [3+2] cycloadditions utilizing aspartic acid. This approach provides valuable insights into mechanistic pathways, allowing for the distinction between concerted and stepwise processes based on reaction efficiency and diastereoselectivity. Both aspartic and glutamic acids have been employed for a thorough evaluation and exploration of decarboxylation-driven double annulations. This method effectively constructs pyrrolizidine frameworks through a concerted double 1,3-dipolar cycloaddition with aspartic acid, as wel
APA, Harvard, Vancouver, ISO, and other styles
2

Zhao, Baoguo, and Wen-Wen Chen. "Decarboxylative Umpolung Synthesis of Amines from Carbonyl Compounds." Synlett 31, no. 16 (2020): 1543–50. http://dx.doi.org/10.1055/s-0040-1707157.

Full text
Abstract:
2-Azaallyl anions are valuable intermediates which have versatile applications in functionalization with various electrophiles. Decarboxylation of the imines formed from aromatic aldehydes and α,α-diphenylglycine provides an interesting and efficient way to generate delocalized 2-azaallyl anions, which display high reactivity toward different electrophiles with excellent regioselectivity at the diphenylketimino aryl carbon of the 2-azaallyl anions. The transformation produces various amines in good yields under very mild conditions. This Synpacts article highlights the recent advances on the d
APA, Harvard, Vancouver, ISO, and other styles
3

Trisrivirat, Duangthip, Narin Lawan, Pirom Chenprakhon, Daisuke Matsui, Yasuhisa Asano, and Pimchai Chaiyen. "Mechanistic insights into the dual activities of the single active site of l-lysine oxidase/monooxygenase from Pseudomonas sp. AIU 813." Journal of Biological Chemistry 295, no. 32 (2020): 11246–61. http://dx.doi.org/10.1074/jbc.ra120.014055.

Full text
Abstract:
l-Lysine oxidase/monooxygenase (l-LOX/MOG) from Pseudomonas sp. AIU 813 catalyzes the mixed bioconversion of l-amino acids, particularly l-lysine, yielding an amide and carbon dioxide by an oxidative decarboxylation (i.e. apparent monooxygenation), as well as oxidative deamination (hydrolysis of oxidized product), resulting in α-keto acid, hydrogen peroxide (H2O2), and ammonia. Here, using high-resolution MS and monitoring transient reaction kinetics with stopped-flow spectrophotometry, we identified the products from the reactions of l-lysine and l-ornithine, indicating that besides decarboxy
APA, Harvard, Vancouver, ISO, and other styles
4

Li, Jiawei, George N. Khairallah, and Richard A. J. O'Hair. "Decarboxylation versus Acetonitrile Loss in Silver Acetate and Silver Propiolate Complexes, [RCO2Ag2(CH3CN)n]+ (where R = CH3 and CH3C≡C; n = 1 and 2)." Australian Journal of Chemistry 68, no. 9 (2015): 1385. http://dx.doi.org/10.1071/ch15210.

Full text
Abstract:
Gas-phase experiments using collision-induced dissociation in an ion trap mass spectrometer have been used in combination with density functional theory (DFT) calculations (at the B3LYP/SDD6–31+G(d) level of theory) to examine the competition between decarboxylation and loss of a coordinated acetonitrile in the unimolecular fragmentation reactions of the silver acetate and silver propiolate complexes, [RCO2Ag2(CH3CN)n]+ (where R = CH3 and CH3C≡C; n = 1 and 2), introduced into the gas-phase via electrospray ionisation. When R = CH3, loss of acetonitrile is the sole reaction channel observed for
APA, Harvard, Vancouver, ISO, and other styles
5

Kluger, Ronald. "Catalyzing decarboxylation by taming carbon dioxide." Pure and Applied Chemistry 87, no. 4 (2015): 353–60. http://dx.doi.org/10.1515/pac-2014-0907.

Full text
Abstract:
AbstractDecarboxylation reactions on enzymes are consistently much faster than their nonenzymic counterparts. Examination of the potential for catalysis in the nonenzymic reactions revealed that the reaction is slowed by the failure of CO2 to be launched into solution upon C–C bond cleavage. Catalysts can facilitate the reaction by weakening the C–CO2H bond but this is not sufficient. Converting the precursor of CO2 into a precursor of bicarbonate facilitates the forward reaction as does protonation of the nascent carbanion.
APA, Harvard, Vancouver, ISO, and other styles
6

Simpson, Quillon, Robert Konrath та David W. Lupton. "Enantioselective Pd-Catalysed Deallylative γ-Lactonisation of Propargyl Carbazolone Allyl Carbonates: Mechanistic Insight into their Decarboxylative Allylation". Australian Journal of Chemistry 67, № 9 (2014): 1353. http://dx.doi.org/10.1071/ch14211.

Full text
Abstract:
Subjection of N-methyl carbazolone allyl carbonates bearing a propargyl side chain to Pd0 catalysis leads to the formation of enantioenriched γ-lactones, rather than the expected products of decarboxylative allylation. This side reaction has not been observed with the enantioselective decarboxylative allylation of related β-ketoesters, and provides evidence for a mechanism involving turnover limiting decarboxylation from the palladium carboxylate resting state. Following lactonisation, the Pd0 catalyst is regenerated by PdII reductive alkyne coupling.
APA, Harvard, Vancouver, ISO, and other styles
7

Koleda, Olesja, Janis Sadauskis, Darja Antonenko, Edvards Janis Treijs, Raivis Davis Steberis, and Edgars Suna. "Entry to 2-aminoprolines via electrochemical decarboxylative amidation of N‑acetylamino malonic acid monoesters." Beilstein Journal of Organic Chemistry 21 (March 19, 2025): 630–38. https://doi.org/10.3762/bjoc.21.50.

Full text
Abstract:
The electrochemical synthesis of 2-aminoprolines based on anodic decarboxylation–intramolecular amidation of readily available N-acetylamino malonic acid monoesters is reported. The decarboxylative amidation under Hofer–Moest reaction conditions proceeds in an undivided cell under constant current conditions in aqueous acetonitrile and provides access to N-sulfonyl, N-benzoyl, and N-Boc-protected 2-aminoproline derivatives.
APA, Harvard, Vancouver, ISO, and other styles
8

Sheik, Cody S., H. James Cleaves, Kristin Johnson-Finn, et al. "Abiotic and biotic processes that drive carboxylation and decarboxylation reactions." American Mineralogist 105, no. 5 (2020): 609–15. http://dx.doi.org/10.2138/am-2020-7166ccbyncnd.

Full text
Abstract:
Abstract Carboxylation and decarboxylation are two fundamental classes of reactions that impact the cycling of carbon in and on Earth’s crust. These reactions play important roles in both long-term (primarily abiotic) and short-term (primarily biotic) carbon cycling. Long-term cycling is important in the subsurface and at subduction zones where organic carbon is decomposed and outgassed or recycled back to the mantle. Short-term reactions are driven by biology and have the ability to rapidly convert CO2 to biomass and vice versa. For instance, carboxylation is a critical reaction in primary pr
APA, Harvard, Vancouver, ISO, and other styles
9

Syromolotov, Alexander V., Alexander A. Kimyashov, and Sergey V. Sukhorukov. "Decarboxylation 2'-dicarboxy-5-(methyl-5'-indolyl-3')- indolyl-3-acetic acid with use of salts of copper." Butlerov Communications 58, no. 4 (2019): 58–61. http://dx.doi.org/10.37952/roi-jbc-01/19-58-4-58.

Full text
Abstract:
In this article a decarboxylation method based on the use of quinoline with a copper salt is discussed. Decarboxylation is the elimination of CO2 from the carboxylic group of carboxylic acids or the carboxylate group of their salts. The process is used to produce a large number of organic compounds, such classes as alkanes, alkenes, alcohols, ketones, ethers and esters. Decarboxylation plays an important role in the metabolism of living organisms, namely in the decarboxylation of amino acids. From this we can conclude that the process is important to study. Decarboxylation proceeds in differen
APA, Harvard, Vancouver, ISO, and other styles
10

Richard, John P. "Enzymatic catalysis of proton transfer and decarboxylation reactions." Pure and Applied Chemistry 83, no. 8 (2011): 1555–65. http://dx.doi.org/10.1351/pac-con-11-02-05.

Full text
Abstract:
Deprotonation of carbon and decarboxylation at enzyme active sites proceed through the same carbanion intermediates as for the uncatalyzed reactions in water. The mechanism for the enzymatic reactions can be studied at the same level of detail as for nonenzymatic reactions, using the mechanistic tools developed by physical organic chemists. Triosephosphate isomerase (TIM)-catalyzed interconversion of D-glyceraldehyde 3-phosphate (GAP) and dihydroxyacetone phosphate (DHAP) is being studied as a prototype for enzyme-catalyzed proton transfer, and orotidine monophosphate decarboxylase (OMPDC)-cat
APA, Harvard, Vancouver, ISO, and other styles
11

Razaq, Iram, Keith E. Simons, and Jude A. Onwudili. "Parametric Study of Pt/C-Catalysed Hydrothermal Decarboxylation of Butyric Acid as a Potential Route for Biopropane Production." Energies 14, no. 11 (2021): 3316. http://dx.doi.org/10.3390/en14113316.

Full text
Abstract:
Sustainable fuel-range hydrocarbons can be produced via the catalytic decarboxylation of biomass-derived carboxylic acids without the need for hydrogen addition. In this present study, 5 wt% platinum on carbon (Pt/C) has been found to be an effective catalyst for hydrothermally decarboxylating butyric acid in order to produce mainly propane and carbon dioxide. However, optimisation of the reaction conditions is required to minimise secondary reactions and increase hydrocarbon selectivity towards propane. To do this, reactions using the catalyst with varying parameters such as reaction temperat
APA, Harvard, Vancouver, ISO, and other styles
12

Secara, Natalia. "Dihydroxyfumaric Acid Transformation." Chemistry Journal of Moldova 3, no. 1 (2008): 127–28. http://dx.doi.org/10.19261/cjm.2008.03(1).06.

Full text
Abstract:
This short communication presents several preliminary results obtained during kinetic investigation of dihydroxyfumaric acid decarboxylation. The reaction order towards the hydrogen ions concentration has been established, the correlation between the decarboxylation velocity and temperature has been found, the Arrhenius equation for the decarboxylation constant has been drawn.
APA, Harvard, Vancouver, ISO, and other styles
13

Melzer, E., and H. L. Schmidt. "Carbon isotope effects on the decarboxylation of carboxylic acids. Comparison of the lactate oxidase reaction and the degradation of pyruvate by H2O2." Biochemical Journal 252, no. 3 (1988): 913–15. http://dx.doi.org/10.1042/bj2520913.

Full text
Abstract:
The isotope effect at C-1 on the H2O2-catalysed decarboxylation of pyruvate (used as a model reaction for the enzymic reaction) increases between pH 3 and 10 from 1.0007 +/- 0.0004 to 1.0283 +/- 0.0014 (25 degrees C). This result indicates a change in the rate-determining step from formation of the tetrahedral intermediate to decarboxylation of this intermediate. Practically no isotope fractionation at C-1 (1.0011 +/- 0.0002, pH 6.0, 25 degrees C) is found in the lactate oxidase-catalysed decarboxylation of lactate, which is indicative for the existence of an irreversible O2-dependent step pri
APA, Harvard, Vancouver, ISO, and other styles
14

Kurtay, Gülbin, Tugba Soganci, Kübra Sarikavak, Metin Ak, and Mustafa Güllü. "Synthesis and electrochemical characterization of a new benzodioxocine-fused poly(N-methylpyrrole) derivative: a joint experimental and DFT study." New Journal of Chemistry 44, no. 43 (2020): 18929–41. http://dx.doi.org/10.1039/d0nj03992f.

Full text
Abstract:
Synthesis of a new electropolymerizable monomer, XyPMe, regarding the reaction of diethyl N-methyl-3,4-dihydroxypyrrole-2,5-dicarboxylate and 1,2-bis(bromomethyl)benzene with concomitant hydrolysis and decarboxylation reactions was accomplished.
APA, Harvard, Vancouver, ISO, and other styles
15

DING, WAN-JIAN, LING-YAN NI, WEI-HAI FANG, and JIAN-GUO YU. "THEORETICAL STUDY ON THE UNIMOLECULAR REACTIONS OF GLYOXYLIC ACID." Journal of Theoretical and Computational Chemistry 04, spec01 (2005): 725–36. http://dx.doi.org/10.1142/s021963360500174x.

Full text
Abstract:
The potential energy surfaces of isomerization, decarboxylation, and decarbonylation reactions of glyoxylic acid have been characterized by the B3LYP/cc-pVTZ, B3LYP/aug-cc-pVTZ, and MP2/cc-pVTZ calculations. There is a relatively high barrier on the isomerization pathway from Tc to Tt , due to existence of the intramolecular H-bond in the Tc to Tt structures. The decarboxylation reaction proceeds mainly through a stepwise mechanism. The singlet hydroxyl carbene and carbon dioxide are formed in the first step, which is followed by isomerization of the carbene to formaldehyde in the second step.
APA, Harvard, Vancouver, ISO, and other styles
16

Guthrie, J. Peter, Sriyawathie Peiris, Margaret Simkin, and Yun Wang. "Rate constants for decarboxylation reactions calculated using no barrier theory." Canadian Journal of Chemistry 88, no. 2 (2010): 79–98. http://dx.doi.org/10.1139/v09-164.

Full text
Abstract:
No barrier theory (NBT) provides both a qualitative way of thinking about what makes a reaction fast or slow and a quantitative way of calculating the rate constant (free energy of activation) corresponding to a particular mechanism. The origin and development of this idea are reviewed and examples of its use for qualitative understanding are presented before applying it to a set of decarboxylations. From the literature, a set of best values for rate constants for decarboxylation was picked. Detailed mechanistic models were developed for reactions leading to delocalized “anions” or to localize
APA, Harvard, Vancouver, ISO, and other styles
17

Kumashiro, Masaya, Kosuke Ohsawa, and Takayuki Doi. "Photocatalyzed Oxidative Decarboxylation Forming Aminovinylcysteine Containing Peptides." Catalysts 12, no. 12 (2022): 1615. http://dx.doi.org/10.3390/catal12121615.

Full text
Abstract:
The formation of (2S,3S)-S-[(Z)-aminovinyl]-3-methyl-D-cysteine (AviMeCys) substructures was developed based on the photocatalyzed-oxidative decarboxylation of lanthionine-bearing peptides. The decarboxylative selenoetherification of the N-hydroxyphthalimide ester, generated in situ, proceeded under mild conditions at −40 °C in the presence of 1 mol% of eosin Y-Na2 as a photocatalyst and the Hantzsch ester. The following β-elimination of the corresponding N,Se-acetal was operated in a one-pot operation, led to AviMeCys substructures found in natural products in moderate to good yields. The sul
APA, Harvard, Vancouver, ISO, and other styles
18

Verma, Anand Mohan, and Nanda Kishore. "Kinetics of Decomposition Reactions of Acetic Acid Using DFT Approach." Open Chemical Engineering Journal 12, no. 1 (2018): 14–23. http://dx.doi.org/10.2174/1874123101812010014.

Full text
Abstract:
Object: Excessive amount of oxygen content in unprocessed bio-oil deteriorates the quality of bio-oil which cannot be used in transportation vehicles without upgrading. Acetic acid (CH3COOH) is a vital component of ‘acids’ catalogue of unprocessed bio-oil produced from thermochemical conversions of most of biomass feedstocks such as switchgrass, alfalfa, etc. In this study, the decomposition reactions of acetic acid are carried out by two reaction pathways, i.e., decarboxylation and dehydration reactions. In addition, the reaction rates of decomposition are analysed in a wide range of temperat
APA, Harvard, Vancouver, ISO, and other styles
19

Tanner, Dennis D., and Soad A. A. Osman. "Oxidative decarboxylation. On the mechanism of the potassium persulfate-promoted decarboxylation reaction." Journal of Organic Chemistry 52, no. 21 (1987): 4689–93. http://dx.doi.org/10.1021/jo00230a007.

Full text
APA, Harvard, Vancouver, ISO, and other styles
20

Zhou, Qiulan, Yaping Ma, Xuhang Ma, et al. "Synthesis of nanoporous graphenes via decarboxylation reaction." Chemical Communications 56, no. 47 (2020): 6336–39. http://dx.doi.org/10.1039/d0cc01913e.

Full text
APA, Harvard, Vancouver, ISO, and other styles
21

Song, Ming Zhi, Zai Long Zhang, Chuan Gang Fan, Da Zhi Li, and Shi Guo Zhang. "Ethylenediamine Catalyzed Decarboxylation of Oxaloacetic Acid: A DFT Investigation." Advanced Materials Research 781-784 (September 2013): 253–58. http://dx.doi.org/10.4028/www.scientific.net/amr.781-784.253.

Full text
Abstract:
The decarboxylation mechanism of oxaloacetic acid aided with ethylenediamine or without any catalyst is investigated employing Density Functional Theory (DFT). DFT calculations for both the gas phase and in water solution indicate a stepwise mechanism for each of the steps of the reactions. In the catalyzed mechanism, the dehydration of carbinolamine (IM1) is via a seven-membered ring transition structure (TS5), which is consistent with the structure proposed by Thalji, et al. The decarboxylation of the imine (IM6) is the rate determining step with an energy barrier of 16.46 kcal/mol, lower th
APA, Harvard, Vancouver, ISO, and other styles
22

Sutor-Świeży, Katarzyna, Michał Antonik, Justyna Proszek, et al. "Dehydrogenation of Betacyanins in Heated Betalain-Rich Extracts of Red Beet (Beta vulgaris L.)." International Journal of Molecular Sciences 23, no. 3 (2022): 1245. http://dx.doi.org/10.3390/ijms23031245.

Full text
Abstract:
Betacyanins are a group of water-soluble red-violet compounds containing nitrogen in their structure. These are biosynthesized in red beetroot (Beta vulgaris L.), a widely consumed vegetable that contains significant amounts of nutritious and bioactive compounds which are also found in dietary supplements. This contribution presents results of betacyanin thermal oxidation (resulting in dehydrogenation) interrelated with decarboxylation in selected acetate/phosphate buffers at pH 3–8 and at 85 °C, which may be of particular significance for formulation and performance of foods. Most of the reac
APA, Harvard, Vancouver, ISO, and other styles
23

Hardhianti, Meiga Putri Wahyu, Rochmadi, and Muhammad Mufti Azis. "Kinetic Studies of Esterification of Rosin and Pentaerythritol." Processes 10, no. 1 (2021): 39. http://dx.doi.org/10.3390/pr10010039.

Full text
Abstract:
Esterification of rosin with pentaerythritol produces rosin pentaerythritol ester (RPE) which is widely used in paint, coating, and pressure-sensitive and hot-melt adhesive industries. Although RPE has excellent valuable applications and has been industrially produced, studies on the reaction kinetics have not been widely reported in the present literature. This work proposed a kinetic study of RPE synthesis by including a series of consecutive reactions forming mono-, di-, tri-, and tetra-ester with decarboxylation of rosin as a side reaction in the kinetics model. For esterification, the rea
APA, Harvard, Vancouver, ISO, and other styles
24

Yeasmin, Humaira. "Plasmon Induced Decarboxylation of Aromatic Carboxylic Acids on Bipolar Electrodeposited Au- and Ag-Film Monitored by SERS." ECS Meeting Abstracts MA2023-01, no. 51 (2023): 2760. http://dx.doi.org/10.1149/ma2023-01512760mtgabs.

Full text
Abstract:
Here, we describe the decarboxylation of 4-mercaptobenzoic acid and its isomers using SERS as an in-situ probe of chemical reactions near metal nanostructures to understand the reaction path as well as confirm new species of reacting systems. To avoid the shortcomings of conventional techniques of nanostructure synthesis, such as hydrothermal processes1, chemical reduction , electrochemical synthesis2 which involve the use of surfactants, reducing agents or toxic organic3 , 4 solvents, that limit their applications in terms of the economic and environmental considerations, we have designed and
APA, Harvard, Vancouver, ISO, and other styles
25

Hoffmann, Artur, and Peter Dimroth. "Stereochemistry of the methylmalonyl-CoA decarboxylation reaction." FEBS Letters 220, no. 1 (1987): 121–25. http://dx.doi.org/10.1016/0014-5793(87)80888-8.

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

Dickstein, Joshua S., Carol A. Mulrooney, Erin M. O'Brien, Barbara J. Morgan, and Marisa C. Kozlowski. "Development of a Catalytic Aromatic Decarboxylation Reaction." Organic Letters 9, no. 13 (2007): 2441–44. http://dx.doi.org/10.1021/ol070749f.

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

Patil, Gaurav, Hanna Michlits, Paul G. Furtmüller, and Stefan Hofbauer. "Reactivity of Coproheme Decarboxylase with Monovinyl, Monopropionate Deuteroheme." Biomolecules 13, no. 6 (2023): 946. http://dx.doi.org/10.3390/biom13060946.

Full text
Abstract:
Coproheme decarboxylases (ChdCs) are terminal enzymes of the coproporphyrin-dependent heme biosynthetic pathway. In this reaction, two propionate groups are cleaved from the redox-active iron-containing substrate, coproheme, to form vinyl groups of the heme b product. The two decarboxylation reactions proceed sequentially, and a redox-active three-propionate porphyrin, called monovinyl, monopropionate deuteroheme (MMD), is transiently formed as an intermediate. While the reaction mechanism for the first part of the redox reaction, which is initiated by hydrogen peroxide, has been elucidated in
APA, Harvard, Vancouver, ISO, and other styles
28

Cabanes, J., A. Sanchez-Ferrer, R. Bru, and F. García-Carmona. "Chemical and enzymic oxidation by tyrosinase of 3,4-dihydroxymandelate." Biochemical Journal 256, no. 2 (1988): 681–84. http://dx.doi.org/10.1042/bj2560681.

Full text
Abstract:
Tyrosinase usually catalyses the conversion of monophenols into o-diphenols and the oxidation of diphenols to the corresponding o-quinones. Sugumaran [(1986) Biochemistry 25, 4489-4492] has previously proposed an unusual oxidative decarboxylation of 3,4-dihydroxymandelate catalysed by tyrosinase. Our determination of the intermediates involved in the reaction demonstrated that 3,4-dihydroxybenzaldehyde is not the first intermediate appearing in the medium during the enzymic reaction. Re-examination of this new activity of tyrosinase has demonstrated that the product of the enzyme action is the
APA, Harvard, Vancouver, ISO, and other styles
29

Fan, Chuangang, and Mingzhi Song. "Mechanistic Insights into Protonated Diamines-catalyzed Decarboxylation of Oxaloacetate." Letters in Organic Chemistry 16, no. 3 (2019): 202–8. http://dx.doi.org/10.2174/1570178615666181003133432.

Full text
Abstract:
The chemical mechanisms of protonated diamines-catalyzed decarboxylation of oxaloacetic acid anions in water solutions have been studied by using density functional theory. The calculated results show that the activated Gibbs free energy of the decarboxylation step is the highest in the whole diamine-catalytic processes for OA2-, and protonated ethylenediamine (ENH+) is the best catalyst of the five diamines, which is consistent with the study of Thalji et al. However, for OA-, different with OA2-, the dehydration step is the rate-determining one except 1,3-diaminopropane, and protonated 1,4-
APA, Harvard, Vancouver, ISO, and other styles
30

Aguer, Jean-Pierre, Frédéric Blachère, Pierre Boule, Sandrine Garaudee, and Chantal Guillard. "Photolysis of dicamba (3,6-dichloro-2-methoxybenzoic acid) in aqueous solution and dispersed on solid supports." International Journal of Photoenergy 2, no. 2 (2000): 81–86. http://dx.doi.org/10.1155/s1110662x00000118.

Full text
Abstract:
Dicamba (3,6-dichloro-2-methoxybenzoic acid) was exposed to UV light in aqueous solution. It was also irradiated in the solid phase without support or adsorbed on laponite (a synthetic clay) and ferric oxide.Two main photoproducts (1) and (2) were identified in irradiated aqueous solutions. Both involve the substitution of chlorine by OH. The unexpected product (2) is formed through a kinetic reaction of primary product; it results from an oxidation and it is not formed in the absence of oxygen.A huge number of intermediate products were simultaneously formed when dicamba is irradiated in the
APA, Harvard, Vancouver, ISO, and other styles
31

Kuwano, Ryoichi, Yusuke Makida, and Yasutaka Matsumoto. "Palladium-Catalyzed Decarboxylation of Benzyl Fluorobenzoates." Synlett 28, no. 19 (2017): 2573–76. http://dx.doi.org/10.1055/s-0036-1588572.

Full text
Abstract:
The decarboxylation of benzyl fluorobenzoates has been developed by using the palladium catalyst prepared in situ from Pd(η3-allyl)Cp and bulky monophosphine ligand XPhos. The catalytic reaction afforded a range of fluorinated diarylmethanes in good yields with broad functional-group compatibility. The substrates were readily synthesized by condensation of the corresponding benzoic acid with benzyl alcohol. Therefore, the transformation is formally regarded as a cross-coupling reaction between fluorine-containing benzoic acids and benzyl alcohols.
APA, Harvard, Vancouver, ISO, and other styles
32

O’Hair, Richard A. J. "Gas-phase studies of metal catalyzed decarboxylative cross-coupling reactions of esters." Pure and Applied Chemistry 87, no. 4 (2015): 391–404. http://dx.doi.org/10.1515/pac-2014-1108.

Full text
Abstract:
AbstractMetal-catalyzed decarboxylative coupling reactions of esters offer new opportunities for formation of C–C bonds with CO2as the only coproduct. Here I provide an overview of: key solution phase literature; thermochemical considerations for decarboxylation of esters and thermolysis of esters in the absence of a metal catalyst. Results from my laboratory on the use of multistage ion trap mass spectrometry experiments and DFT calculations to probe the gas-phase metal catalyzed decarboxylative cross-coupling reactions of allyl acetate and related esters are then reviewed. These studies have
APA, Harvard, Vancouver, ISO, and other styles
33

Bian, Junjie, Yue Wang, Qi Zhang, Xudong Fang, Lijuan Feng, and Chunhu Li. "Fatty acid decarboxylation reaction kinetics and pathway of co-conversion with amino acid on supported iron oxide catalysts." RSC Adv. 7, no. 75 (2017): 47279–87. http://dx.doi.org/10.1039/c7ra08507a.

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

Turhanen, Petri A., Janne Weisell, and Jouko J. Vepsäläinen. "Preparation of mixed trialkyl alkylcarbonate derivatives of etidronic acid via an unusual route." Beilstein Journal of Organic Chemistry 8 (November 20, 2012): 2019–24. http://dx.doi.org/10.3762/bjoc.8.228.

Full text
Abstract:
A method to prepare four (3a–d) trialkyl alkylcarbonate esters of etidronate from P,P'-dimethyl etidronate and alkyl chloroformate was developed by utilizing unexpected demethylation and decarboxylation reactions. The reaction with the sterically more hindered isobutyl chloroformate at a lower temperature (90 °C) produced the P,P'-diester (2) as a stable intermediate product. A possible reaction mechanism is discussed to explain these methyl substitutions. These unusual reactions also clarify why it is difficult to prepare alkylcarbonate prodrugs from bisphosphonates. The compounds prepared we
APA, Harvard, Vancouver, ISO, and other styles
35

Gupta, Shelaka, Md Imteyaz Alam, Tuhin Suvra Khan, Nishant Sinha, and M. Ali Haider. "On the mechanism of retro-Diels–Alder reaction of partially saturated 2-pyrones to produce biorenewable chemicals." RSC Advances 6, no. 65 (2016): 60433–45. http://dx.doi.org/10.1039/c6ra11697c.

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

Perera-Solis, Diego D., Vladimir L. Zholobenko, Andrew Whiting, and Hugh Christopher Greenwell. "Heterogeneous ketonic decarboxylation of dodecanoic acid: studying reaction parameters." RSC Advances 11, no. 56 (2021): 35575–84. http://dx.doi.org/10.1039/d1ra06871g.

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

Mola, Antonia Di, Antonio Macchia, Laura Palombi, and Antonio Massa. "Methyl 2-(1-methyl-3-oxoisoindolin-1-yl)acetate." Molbank 2020, no. 2 (2020): M1131. http://dx.doi.org/10.3390/m1131.

Full text
Abstract:
In this work, we report a facile access to a 3,3-disubstituted isoindolinone by means of Krapcho decarboxylation reaction of the respective substituted dimethyl malonate derivative. Good isolated yields were obtained under mild reaction conditions.
APA, Harvard, Vancouver, ISO, and other styles
38

Yu, Jiyao, Runyu Mao, Qingyu Wang та Jie Wu. "Synthesis of β-keto sulfones via a multicomponent reaction through sulfonylation and decarboxylation". Organic Chemistry Frontiers 4, № 4 (2017): 617–21. http://dx.doi.org/10.1039/c7qo00026j.

Full text
Abstract:
A copper(i)-catalyzed synthesis of β-keto sulfones through reaction of aryldiazonium tetrafluoroborates, 3-arylpropiolic acids, sulfur dioxide, and water is realized. This reaction proceeds through a tandem radical process involving sulfonylation and decarboxylation.
APA, Harvard, Vancouver, ISO, and other styles
39

Papineau, Dominic. "Chemically oscillating reactions in the formation of botryoidal malachite." American Mineralogist 105, no. 4 (2020): 447–54. http://dx.doi.org/10.2138/am-2020-7029.

Full text
Abstract:
Abstract The origin of banding patterns in malachite [Cu2CO3(OH)2] is an enduring problem in geology. While the bright green, vivid colors of this mineral have been attributed to the presence of Cu, no specific process has been proposed that can explain the perfect circularly concentric banding and geometrical shapes in botryoidal malachite. These patterns of concentric equidistant laminations are comparable to those arising from chemically oscillating experiments using the classical reactants of the Belousov-Zhabotinsky (B-Z) reaction. Through optical microscopy and micro-Raman imaging, this
APA, Harvard, Vancouver, ISO, and other styles
40

Jursík, František, Jana Ondráčková та Bohumil Hájek. "Stereochemical changes accompanying acid decarboxylation of the Λ-[Co((S)-valinato)2CO3]- isomers". Collection of Czechoslovak Chemical Communications 50, № 7 (1985): 1582–87. http://dx.doi.org/10.1135/cccc19851582.

Full text
Abstract:
Λ-trans(O)-[Co((S)-Val)2CO3]- (Val = valine) reacts with 1M-HNO3) (3 min, 28 °C) to form 90% trans(O) and 10% cis(N)-[Co((S)-Val)2(H2O)2]+. Addition of solid KHCO3 to the reaction mixture containing diaquo species gives 96% (95% Λ) trans(O)- and 4% (56% Λ) C1-cis(N)-[Co((S)-Val)2CO3]- isomers indicating decarboxylation of trans(O) isomer proceeds without change in configuration. Acid decarboxylation of Λ-C1-cis(N)-[Co((S)-Val)2CO3]- leads to the mixture consisting from 65% of cis(N)- and 35% of trans(O)-[Co((S)-Val)2(H2O)2]+. When this reaction mixture is treated with solid KHCO, 61% (92% Λ) o
APA, Harvard, Vancouver, ISO, and other styles
41

Mertens, M. A. Stephanie, Daniel F. Sauer, Ulrich Markel, Johannes Schiffels, Jun Okuda, and Ulrich Schwaneberg. "Chemoenzymatic cascade for stilbene production from cinnamic acid catalyzed by ferulic acid decarboxylase and an artificial metathease." Catalysis Science & Technology 9, no. 20 (2019): 5572–76. http://dx.doi.org/10.1039/c9cy01412h.

Full text
APA, Harvard, Vancouver, ISO, and other styles
42

Yuan, Jin-Wei, Shuai-Nan Liu, and Wen-Peng Mai. "Copper-catalysed difluoroalkylation of aromatic aldehydes via a decarboxylation/aldol reaction." Organic & Biomolecular Chemistry 15, no. 36 (2017): 7654–59. http://dx.doi.org/10.1039/c7ob01739a.

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

Gunaganti, Naresh, Anupreet Kharbanda, Naga Rajiv Lakkaniga, et al. "Catalyst free, C-3 functionalization of imidazo[1,2-a]pyridines to rapidly access new chemical space for drug discovery efforts." Chemical Communications 54, no. 92 (2018): 12954–57. http://dx.doi.org/10.1039/c8cc07063f.

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

ZAHEER, KHAN, та AZIZ KHAN A. "Kinetics and Mechanism of Decarboxylation of α-Amino Acids by Ninhydrin". Journal of Indian Chemical Society Vol. 67, Dec 1990 (1990): 963–65. https://doi.org/10.5281/zenodo.6257310.

Full text
Abstract:
Department of Chemistry, Aligarh Muslim University, Aligarh-202 002 <em>Manuscript received 17 January 1990, revised 29 June 1990, accepted 21 November 1990</em> The kinetic study of decarboxylation of \(\)\(\propto\)-amino acids has been carried out at various concentrations of ninhydrin at different temperatures (50-80&deg;) and hydrogen ion concentration from 1.0 x 10<sup>-3</sup>&nbsp;to 1.0 x 10<sup>-6</sup> mol dm<sup>-3</sup>&nbsp;. The reaction follows an irreversible pseudo first order reaction in presence of excess ninhydrin. The rate of the reaction increases with increase in pH and
APA, Harvard, Vancouver, ISO, and other styles
45

Van Arnum, Susan D., Nancy Stepsus, and Barry K. Carpenter. "An unexpected oxidative decarboxylation reaction of frenolicin-B." Tetrahedron Letters 38, no. 3 (1997): 305–8. http://dx.doi.org/10.1016/s0040-4039(96)02339-8.

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

Schijndel, Jack van, Dennis Molendijk, Luiz Alberto Canalle, Erik Theodorus Rump, and Jan Meuldijk. "Temperature Dependent Green Synthesis of 3-Carboxycoumarins and 3,4-unsubstituted Coumarins." Current Organic Synthesis 16, no. 1 (2019): 130–35. http://dx.doi.org/10.2174/1570179415666180924124134.

Full text
Abstract:
Aim and Objective: Because of the low abundance of 3,4-unsubstituted coumarins in plants combined with the complex purification process required, synthetic routes towards 3,4-unsubstituted coumarins are especially valuable. In the present work, we explore the possibilities of a solvent-free Green Knoevenagel condensation on various 2-hydroxybenzaldehyde derivatives and malonic acid without the use of toxic organocatalysts like pyridine and piperidine but only use ammonium bicarbonate as the catalyst. Materials and Methods: To investigate the scope of the Green Knoevenagel condensation for the
APA, Harvard, Vancouver, ISO, and other styles
47

Weliwegamage, U. S. K., S. Sotheeswaran, H. I. C. De Silva, D. G. G. P. Karunaratne, and P. H. Gamage. "Green Concept in Preparation of Biodiesel by Decarboxylation of Fatty Acids Derived from Waste Coconut and Rubber Seed Oils." International Journal of Environmental Issues 2, no. 1 (2024): 55–67. https://doi.org/10.4038/ijei.v2i1.6.

Full text
Abstract:
Green production of diesel was carried out using waste coconut oil and rubber seed oil. Experiments were carried out using preliminary apparatus constructed in house as well as using a high temperature pressure reactor. Acid values of oil samples were determined before and after the decarboxylation process. Decarboxylation of hydrolysed waste coconut oil produced the hydrocarbons, nonane (15.60% w/w), decane (46.38% w/w), undecane (23.2%w/w), dodecane (1.34%w/w) and tridecane (0.36% w/w). Decarboxylation of non-hydrolysed waste coconut oil when reacted at 350 °C using Pd/C catalyst resulted in
APA, Harvard, Vancouver, ISO, and other styles
48

BERTOLDI, Mariarita, Virginia CARBONE та Carla BORRI VOLTATTORNI. "Ornithine and glutamate decarboxylases catalyse an oxidative deamination of their α-methyl substrates". Biochemical Journal 342, № 3 (1999): 509–12. http://dx.doi.org/10.1042/bj3420509.

Full text
Abstract:
Ornithine decarboxylase (ODC) from Lactobacillus 30a catalyses the cleavage of α-methylornithine into ammonia and 2-methyl-1-pyrroline; glutamate decarboxylase (GAD) from Escherichia coli catalyses the cleavage of α-methylglutamate into ammonia and laevulinic acid. In our analyses, 2-methyl-1-pyrroline and laevulinic acid were identified by HPLC and mass spectroscopic analysis, and ammonia was identified by means of glutamate dehydrogenase. Molecular oxygen was consumed during these reactions in a 1:2 molar ratio with respect to the products. The catalytic efficiencies (kcat/Km) of the reactio
APA, Harvard, Vancouver, ISO, and other styles
49

Čolnik, Maja, Darja Pečar, Željko Knez, Andreja Goršek, and Mojca Škerget. "Kinetics Study of Hydrothermal Degradation of PET Waste into Useful Products." Processes 10, no. 1 (2021): 24. http://dx.doi.org/10.3390/pr10010024.

Full text
Abstract:
Kinetics of hydrothermal degradation of colorless polyethylene terephthalate (PET) waste was studied at two temperatures (300 °C and 350 °C) and reaction times from 1 to 240 min. PET waste was decomposed in subcritical water (SubCW) by hydrolysis to terephthalic acid (TPA) and ethylene glycol (EG) as the main products. This was followed by further degradation of TPA to benzoic acid by decarboxylation and degradation of EG to acetaldehyde by a dehydration reaction. Furthermore, by-products such as isophthalic acid (IPA) and 1,4-dioxane were also detected in the reaction mixture. Taking into acc
APA, Harvard, Vancouver, ISO, and other styles
50

Pereira, C. I., D. Matos, M. V. San Romão, and M. T. Barreto Crespo. "Dual Role for the Tyrosine Decarboxylation Pathway in Enterococcus faecium E17: Response to an Acid Challenge and Generation of a Proton Motive Force." Applied and Environmental Microbiology 75, no. 2 (2008): 345–52. http://dx.doi.org/10.1128/aem.01958-08.

Full text
Abstract:
ABSTRACT In this work we investigated the role of the tyrosine decarboxylation pathway in the response of Enterococcus faecium E17 cells to an acid challenge. It was found that 91% of the cells were able to remain viable in the presence of tyrosine when they were incubated for 3 h in a complex medium at pH 2.5. This effect was shown to be related to the tyrosine decarboxylation pathway. Therefore, the role of tyrosine decarboxylation in pH homeostasis was studied. The membrane potential and pH gradient, the parameters that compose the proton motive force (PMF), were measured at different pHs (
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!