Academic literature on the topic 'Homolytic cleavage of diborons'

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

Select a source type:

Consult the lists of relevant articles, books, theses, conference reports, and other scholarly sources on the topic 'Homolytic cleavage of diborons.'

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.

Journal articles on the topic "Homolytic cleavage of diborons"

1

Verma, Piyush Kumar, Naresh Kumar Meher, and K. Geetharani. "Homolytic cleavage of diboron(4) compounds using diazabutadiene derivatives." Chemical Communications 57, no. 64 (2021): 7886–89. http://dx.doi.org/10.1039/d1cc02881b.

Full text
Abstract:
Homolytic cleavage of diboron was achieved using diazabutadiene derivatives (DABs). The cleavage is accompanied by the formation of new π-bonds and the geometry of the product is highly dependent on the substituents on the DAB units.
APA, Harvard, Vancouver, ISO, and other styles
2

Wang, Pengfei, Yuli Li, and Guangwei Wang. "Tetrahydroxydiboron-Initiated Atom-Transfer Radical Cyclization." Synthesis 53, no. 19 (April 19, 2021): 3555–63. http://dx.doi.org/10.1055/a-1485-4956.

Full text
Abstract:
AbstractIn this work, the first diboron reagent initiated atom-transfer radical cyclization was reported, in which the boryl radicals were generated by the homolytic cleavage of a B–B single bond weakened by the coordination of Lewis base. To clarify the role of carbonate and DMF in the cleavage of B–B bond, we calculated the free energy diagram of two pathways by density functional theory (DFT) investigations. The DFT calculation showed that the presence of carbonate facilitates the B–B bond cleavage to form boron radicals, which can be further stabilized by DMF. Subsequent atom-transfer cyclization initiated by stabilized dihydroxyboron radical is also energetically favored.
APA, Harvard, Vancouver, ISO, and other styles
3

Cao, Levy L., and Douglas W. Stephan. "Homolytic Cleavage Reactions of a Neutral Doubly Base Stabilized Diborane(4)." Organometallics 36, no. 16 (August 9, 2017): 3163–70. http://dx.doi.org/10.1021/acs.organomet.7b00522.

Full text
APA, Harvard, Vancouver, ISO, and other styles
4

Wang, Jiaoyang, Wenrui Zheng, and Yuanyuan Zheng. "Theoretical study on homolytic B–B cleavages of diboron(4) compounds." RSC Adv. 7, no. 78 (2017): 49251–72. http://dx.doi.org/10.1039/c7ra09006d.

Full text
APA, Harvard, Vancouver, ISO, and other styles
5

Tsurugi, Hayato, Kazushi Mashima, Luis C. Misal Castro, and Ibrahim Sultan. "Pyridine-Mediated B–B Bond Activation of (RO)2B–B(OR)2 for Generating Borylpyridine Anions and Pyridine-Stabilized Boryl Radicals as Useful Boryl Reagents in Organic Synthesis." Synthesis 53, no. 18 (April 20, 2021): 3211–26. http://dx.doi.org/10.1055/a-1486-8169.

Full text
Abstract:
AbstractSignificant developments have been achieved in recent years toward the utilization of (RO)2B–B(OR)2 for exploring transition-metal-free organic transformations in organic synthesis. Among the various combinations of Lewis bases with diborons developed so far, pyridine derivatives are simple, commercially available, and cheap compounds to expand the synthetic utility of diborons by generating borylpyridine anions and pyridine-stabilized boryl radicals via B–B bond cleavage. These borylpyridine species mediate a series of transformations in both a catalytic and stoichiometric manner for C–X activation (X = halogen, CO2H, NR2) and concomitant C-borylation, hydroboryl­ation, C–C bond formation, and reduction reactions.1 Introduction2 Reaction Pathway for B–B Bond Cleavage of Diborons with Electron-Deficient Pyridines3 Pyridine-Mediated B–B Bond Activation of (RO)2B–B(OR)2 for Application in Organic Synthesis3.1 Dehalogenative C-Borylation3.2 Desulfonative C-Borylation3.3 Decarboxylative C-Borylation3.4 Deaminative C-Borylation3.5 Hydroborylation3.6 C–C Bond Formation3.7 Pyridine Functionalization3.8 Deoxygenation and N-Borylation Reactions4 Conclusions
APA, Harvard, Vancouver, ISO, and other styles
6

Weng, Zhiqiang, and Lai Yoong Goh. "Homolytic Cleavage and Aggregation Processes in Cyclopentadienylchromium Chemistry." Accounts of Chemical Research 37, no. 3 (March 2004): 187–99. http://dx.doi.org/10.1021/ar030003r.

Full text
APA, Harvard, Vancouver, ISO, and other styles
7

Ding, Lanlan, Wenrui Zheng, and Yingxing Wang. "Homolytic C–O Cleavage in Phosphates and Sulfonates." Journal of Physical Chemistry A 119, no. 14 (March 27, 2015): 3488–99. http://dx.doi.org/10.1021/acs.jpca.5b00569.

Full text
APA, Harvard, Vancouver, ISO, and other styles
8

Kampmeier, Jack A. "Regioselectivity in the Homolytic Cleavage of S-Adenosylmethionine." Biochemistry 49, no. 51 (December 28, 2010): 10770–72. http://dx.doi.org/10.1021/bi101509u.

Full text
APA, Harvard, Vancouver, ISO, and other styles
9

Nome, Faruk, Marcos Caroli Rezende, Claudia Maria Sabóia, and Arlindo Clemente Da Silva. "Kinetics of the thermolysis of para-substituted benzylcobalamins and derivatives." Canadian Journal of Chemistry 65, no. 9 (September 1, 1987): 2095–99. http://dx.doi.org/10.1139/v87-347.

Full text
Abstract:
The thermolysis of five para-substituted benzylcobalamins was studied at different temperatures. In the presence of KCN the observed rates increase with the cyanide concentration until a constant value is attained at high [CN−]. In both series the homolytic cleavage of the Co—C bond is slightly dependent on the para-substituent of the benzyl moiety, with ρ values of −0.1 and −0.2 when 5,6-dimethylbenzimidazole (Bzm) and cyanide are the axial ligands respectively. The Co—C bond is weakened by electron-donating axial ligands and the homolytic cleavage rates increase in the order H2O < Bzm < CN.
APA, Harvard, Vancouver, ISO, and other styles
10

Bajaj, Ashima, Rishu Khurana, and Md Ehesan Ali. "Quantum interference and spin filtering effects in photo-responsive single molecule devices." Journal of Materials Chemistry C 9, no. 34 (2021): 11242–51. http://dx.doi.org/10.1039/d1tc02200h.

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

Dissertations / Theses on the topic "Homolytic cleavage of diborons"

1

Verma, Piyush Kumar. "Cobalt-nhc Complexes and Diazabutadienes in Activation of Mono/Diboron Compounds and Their Application in C-b Coupling Reactions." Thesis, 2021. https://etd.iisc.ac.in/handle/2005/5241.

Full text
Abstract:
Boronic acid(esters) have been well recognized as an indispensable coupling partner in the Suzuki-Miyaura cross coupling reactions producing a vast spectrum of molecules, applicable in the diverse field ranging from medicinal to materials sciences.[1] Transition metal catalyzed synthesis of boronic esters from diborons with the assistance of bases is a well-established methodology[2]. In this thesis, the cobalt-N-Heterocyclic carbene complexes catalyzed borylation of organic compounds and interaction of diazabutadienes with diboron compounds will be discussed. (i) In the first section, Co(IMes)2Cl2 catalyzed borylation of aryl halides will be discussed. [3a] The robust protocol, operating under mild condition facilitate the borylation of a diverse range of aryl halides with great efficacy, which includes the challenging aryl chlorides. The preliminary mechanistic studies suggest that base-bis(pinacolato)diboron adduct reduces the Co(IMes)2Cl2 complex to generate Co(IMes)2Cl complex, which acts as an active catalytic species. (ii) The second section deals with catalytic synthesis of primary and secondary alkyl boronic esters using alkyl halides. [3b] The in situ generated Co-NHC complex, in assistance with base and diboron compound, produces the corresponding borylated product from alky halides. The reaction proceeds under very mild conditions and covers a wide range of alkyl halides, including chlorides having different functional groups. (iii) In the third section, development in selective hydroboration of vinyl arenes and aliphatic alkenes will be discussed. [3c] Catalyzed by Co(I)NHC complex, the alkene hydroboration by pinacol borane gives Markovnikov selective product with good selectivity, where the regio-selectivity is controlled by phenyl substituent. In absence of that, complete inversion in the selectivity has been observed. The preliminary mechanistic cycle suggests that the catalytic cycle proceeds via oxidative addition of pinacol borane to [Co] followed by alkene insertion and reduction elimination steps. (iv) The last section discusses the interaction of diazabutadiene molecules with diboron compounds. [3d] The diazabutadiene derivatives have been observed to completely cleave the B-B bond of Bis(catacolato)diboron and Bis(dithiocatacolato)diboron. The preliminary findings hint towards homolytic cleavage of the B-B bond by concerted interaction of the two nitrogen atoms of diazabutadiene with the two boron atoms of the diboron from the same face. References: [1] Boronic Acids-Preparation and Applications in Organic Synthesis, Medicine and Materials, 2nd ed.; Hall, D. G., Ed.; Wiley-VCH: Weinheim, 2011. [2] Neeve, E. C.; Geier, S. J.; Mkhalid, I. A. I.; Westcott, S. A.; Marder, T. B. Diboron(4) Compounds: From Structural Curiosity to Synthetic Workhorse. Chem. Rev. 2016, 116, 9091-9161. [3] (a) Verma, P. K.; Mandal, S.; Geetharani, K. ACS Catal. 2018, 8, 4049-4054. (b) Verma, P. K.; Prasad, K. S.; Varghese, D.; Geetharani, K. Org. Lett. 2020, 22, 4, 1431-1436. (c) Verma, P. K.; Setulekshmi, A. S.; Geetharani, K. Org. Lett. 2018, 20, 7840-7845. (d) Verma, P. K.; Meher, N. K.; Geetharani, K. Accepted for publication in Chem. Commun., Manuscript ID: CC-COM-06-2021-002881.R2.
APA, Harvard, Vancouver, ISO, and other styles
2

Wu, Yi-Wen, and 吳翌彣. "The Study of Electronic Structures of Five-coordinate Saddled Iron(III) Porphyrin Radical Cation and O-O Bond Homolytic Cleavage of (OETPP)FeIIIO(H)OtBu." Thesis, 2018. http://ndltd.ncl.edu.tw/handle/n6af3q.

Full text
Abstract:
碩士
國立中興大學
化學系所
106
In this study, paramagnetic NMR spectroscopy, X-ray diffraction, magnetic susceptibility and DFT calculation are employed to elucidate the strong bonding interaction between iron(III) dx2-y2 and porphyrin a2u orbitals of [Fe(OMTPP∙)Cl]SbCl6 and [Fe(OETPP∙)Cl]SbCl6. Their crystal structures clearly indicate that their saddle deformations are increased compared to those prior to oxidations. Their 1H NMR data present the S = 2 states for such one-electron oxidation states, and demonstrate that their structures in solutions remain strong saddle deformations. In light of 1H NMR spectroscopy of [Fe(TPP∙)Cl]SbCl6 resembling to the above cases, we postulate that the ring structure of [Fe(TPP∙)Cl]+ may possess strongly saddle to have great bonding interaction between iron(III) dx2-y2 and porphyrin a2u orbitals. In the DFT calculations, the degrees of phenyl ring rotation and saddled deformation will also affect their NMR spectra. In another topic, we observe O-O bond homolytic cleavage of (OETPP)FeIIIO(H)OtBu formed in the reaction of Fe(OETPP)ClO4 with TBHP and its activation parameters (ΔH≠ = 47(2) kJ mol-1, ΔS≠ = 83(9)J mol-1K-1) is measured by low-temperature UV-vis spectral data. The corresponding one-electron oxidation product is identified as [Fe(OETPP∙)OH]+, which is an isoelectronic structure as oxoiron(IV) porphyrin, by UV-vis, NMR spectroscopy and ESI-MS spectrometry. According to the experiments of NMR and ESI-MS, we also find that ·OtBu radical will convert Fe(OETPP)ClO4 to [Fe(OETPP∙)OtBu]+. Furthermore, these related iron(III) porphyrin radical cations can be carried out one more electron oxidation to isoporphyrins, an isoelectronic structure as Compound I. These iron(III) saddled isoporphyrins are shown to be reactive for highly selective chlorination of cyclohexene.
APA, Harvard, Vancouver, ISO, and other styles
3

Mala, Deep. "Chemistry of Ru(II) Complexes Bearing N-heterocyclic Carbene, Hydride, and Dihydrogen Ligands : Synthesis, Mechanistic Insights, and Applications." Thesis, 2016. http://etd.iisc.ac.in/handle/2005/4162.

Full text
Abstract:
Introduction N-heterocyclic carbenes (NHC) are strong  donors and weak acceptors. Also, NHC forms strong metal-carbon bonds in metal complexes and hence complexes bearing NHC ligands are in general, thermally stable. Chemistry of NHCs and their transition metal complexes has been explored extensively in catalysis. Metal hydride or dihydrogen complexes are reactive intermediates and also employed as catalysts in several catalytic reactions such as hydrogenation, transfer hydrogenation, and hydroformylation. The binding of H2 to a metal center and its cleavage for the oxidative addition to a metal center are key steps in these catalytic reactions. There has been a substantial development in the field of dihydrogen chemistry. In particular, elongated dihydrogen complexes are of significant interest due to their relevance in hydrogenation reactions. A large number of elongated dihydrogen complexes have been reported. However, there has been no systematic study for the elongation of the H–H bond in dihydrogen complexes employing NHC’s as a co-ligand. Objectives The objectives of this work are as follows: i. Synthesis and characterization of ruthenium dihydrogen complexes bearing an NHC ligand. ii. Systematic variation in the ligand environment around a metal center for the elongation of the H‒H bond in the process of oxidative addition of H2 to a metal center. iii. To synthesize new ruthenium hydride and dihydride complexes and explore their activities in catalysis such as hydrogenation, hydrodehalogenation, and CO2 activation. Significant results A series of dihydrogen complexes bearing NHC ligands of the type [RuCl(η2–H2)(CO)(IMes)(PPh3)(L)](OTf) [L = pyridine (Py), 4-methylpyridine (4MePy), acetonitrile (MeCN), pivalonitrile (Me3CCN)] have been prepared by the protonation of hydride complexes [RuHCl(CO)(IMes)(PPh3)(L)] (L = Py, 4MePy, MeCN, Me3CCN) with HOTf. The ligands (Py, 4MePy, MeCN, Me3CCN) trans to the hydride were found to be labile in all the hydride complexes. The H–H bond distances in the η2–H2 ligands of dihydrogen complexes were found to be temperature dependent. The H–H bond distances got slightly elongated (0.98-0.93 Å) with an increase in the temperature (183-233 K). Phosphine analogues, [RuCl(η2–H2)(CO)(PPh3)2(L)](OTf) [L = Py, 4MePy) of NHC dihydrogen complexes were also synthesized. The H‒H bond distances were temperature invariant (0.93 Å, 223-263 K) in these complexes. A series of homobimetallic ruthenium hydride [{RuHCl(CO)(IMes)(PPh3)}2(NN)] (NN= 4,4′-bpy, 4,4′-dpyen, 4,4′-dpyan) [4,4′-bipyridine (4,4′-bpy); 1,2-bis(4-pyridyl)ethylene (4,4′-dpyen); 1,2-bis(4-pyridyl)ethane) (4,4′-dpyan)] complexes bearing an NHC ligand and their corresponding dihydrogen complexes of the type [{RuCl(η2–H2)(IMes)(PPh3)(CO)}2(NN)][OTf]2 (NN= 4,4′-bpy, 4,4′-dpyen, 4,4′-dpyan) have been synthesized and characterized. They are the first examples of homobimetallic dihydrogen complexes bearing NHC ligands. In addition, hydrogenation of internal and terminal alkenes was carried out using these homobimetallic hydride complexes. NHC/PMe3 based cationic complexes of the type [RuH(CO)(IMes)(PMe3)3](X) (X = Cl, BPh4) have been synthesized and characterized using NMR spectroscopy. Partial hydrodechlorination of CHCl3 to CH2Cl2 was observed in the presence of both the complexes. In addition, these complexes were also found to be efficient for the activation of the C‒Cl bond in CH2Cl2 to form [CH2Cl.PMe3]Cl salt in the presence of excess PMe3. Notably, HD exchange in H2 molecules via CDCl3 was observed in the presence of both the complexes. A new and efficient route has been developed for the synthesis of the ruthenium dihydride complexes [Ru(H)2(CO)(L)(PPh3)2] (L = IMes, IPr, PPh3). The reaction pathway for the synthesis of these derivatives was established by isolation of few intermediates. Insertion of CO2 into one of the Ru‒H bond of [Ru(H)2(CO)(L)(PPh3)2] (L = IMes, IPr) complexes took place to afford bidentate formate complexes [RuH(η2-O2CH)(CO)(L)(PPh3)2] (L = IMes, IPr). Structure formulation of these complexes was done using NMR spectroscopy.
APA, Harvard, Vancouver, ISO, and other styles

Book chapters on the topic "Homolytic cleavage of diborons"

1

James, B. R., and M. T. Ashby. "Homolytic Cleavage to Give Metal-Hydrides." In Inorganic Reactions and Methods, 66–71. Hoboken, NJ, USA: John Wiley & Sons, Inc., 2007. http://dx.doi.org/10.1002/9780470145319.ch29.

Full text
APA, Harvard, Vancouver, ISO, and other styles
2

Kühn, F. E., C. C. Romäo, and W. A. Herrmann. "Homolytic Cleavage." In Compounds of Groups 7-3 (Mn..., Cr..., V..., Ti..., Sc..., La..., Ac...), 1. Georg Thieme Verlag KG, 2003. http://dx.doi.org/10.1055/sos-sd-002-00209.

Full text
APA, Harvard, Vancouver, ISO, and other styles
3

"Oxidative Degradation." In Organic Chemistry of Drug Degradation, 48–109. The Royal Society of Chemistry, 2012. http://dx.doi.org/10.1039/bk9781849734219-00048.

Full text
Abstract:
This chapter contains three functional parts: an introduction (3.1), a description of several major types of autooxidative mechanisms (3.2–3.4), and a discussion of specific oxidation pathways of drugs with various functional groups and structures in relation to each type of the major autooxidative mechanisms (3.5). In Sections 3.2–3.4, the ubiquitously known Fenton reaction and the little known, but more relevant Udenfriend reaction, are discussed in terms of their roles in free radical-mediated autooxidation by activating molecular oxygen into several reactive oxygen species (ROS), that is, O2−˙/HO2˙, H2O2, and HO˙. The radical ROS then triggers radical chain reactions, in which process organic peroxyl radicals and hydroperoxides are the predominant intermediates. The latter can undergo homolytic cleavage, owing to their relatively low O–O bond dissociation energies, as well as metal ion-catalyzed heterolytic cleavage. The homolytic cleavage generates alkoxyl and hydroxyl radicals, while the heterolytic cleavage reproduces peroxyl radical. Non-radical reactions of peroxides were then discussed, in particular those responsible for the formation of N-oxide, S-oxide, and epoxide degradants. The general mechanism for a less known autooxidative degradation pathway, carbanion/enolate-mediated autooxidation (base-catalyzed autooxidation) is also discussed. This mechanism can be significant for those drug molecules containing somewhat “acidic” carbonated CHn moieties, particularly when the drugs are formulated in liquid form. In Section 3.5, more than 60 examples of drug autooxidation in real life scenarios, that is, oxidation occurring under ambient or various stability conditions, are presented and their underlying degradation mechanisms are discussed in details. These examples cover the functional groups, moieties, and structures that are commonly seen in drug molecules.
APA, Harvard, Vancouver, ISO, and other styles
4

"S-Adenosyl Methionine: One Electron and Two Electron Reaction Manifolds in Biosyntheses." In Natural Product Biosynthesis: Chemical Logic and Enzymatic Machinery, 524–68. The Royal Society of Chemistry, 2017. http://dx.doi.org/10.1039/bk9781788010764-00524.

Full text
Abstract:
S-Adenosylmethionine (SAM), likely an ancient molecule in biological systems, is known for delivery of a [CH3+] equivalent to a host of cellular metabolites containing nucleophilic N, O, S, and C centers via heterolytic cleavage of the CH3–S bond in SAM. SAM can also undergo catalyzed hemolytic cleavage of that CH3–S bond by iron/sulfur-containing enzymes that use the resultant 5′-deoxyadenosyl radical for carbon-based radical chemistry on specific substrates. Bioinformatic analysis indicates &gt;50 000 such enzymes in microbial data bases. All these iron/sulfur cluster enzymes are predicted to be sensitive to autoxidation so, in contrast/complementarity to O2-based homolytic chemistry which is by definition aerobic, the SAM radical enzymes function anaerobically.
APA, Harvard, Vancouver, ISO, and other styles
5

"Energy of a Homolytic Cleavage of Communication OH in Replaced 2,6-di-tert.Butylphenols." In Chemistry and Physics of Complex Materials, 225–34. Apple Academic Press, 2013. http://dx.doi.org/10.1201/b16302-13.

Full text
APA, Harvard, Vancouver, ISO, and other styles
6

"Carbon-based Radicals in C–C Bond Formations in Natural Products." In Natural Product Biosynthesis: Chemical Logic and Enzymatic Machinery, 456–522. The Royal Society of Chemistry, 2017. http://dx.doi.org/10.1039/bk9781788010764-00456.

Full text
Abstract:
Molecular oxygen, O2, has a limited role in primary metabolism, albeit a key one as the terminal electron acceptor in mitochondrial respiratory chains. By contrast, oxygenases are interspersed everywhere in the biosynthetic pathways to all the major classes of secondary metabolites. Because O2 is a ground state triplet molecule it is kinetically stable in the presence of organic metabolites, including cellular metabolites. Reductive activation occurs by one-electron paths, mediated either by flavin-dependent enzymes or iron-based enzymes. The iron-based oxygenases carry out homolytic cleavage of substrate C–H bonds, generate carbon-centered radicals, and can lead not only to oxygenation products but to a variety of radical-based rearrangements in product scaffolds, as in morphine, staurosporine, rebeccamycin, penicillin, and cephalosporin biosynthesis.
APA, Harvard, Vancouver, ISO, and other styles
7

Klärner, F. G., M. K. Diedrich, G. Dierkes, and J. S. Gehrke. "Organic Reactions at High Pressure: the Effect of Pressure on cyclizations and Homolytic Bond Cleavage." In High Pressure Food Science, Bioscience and Chemistry, 3–11. Elsevier, 1998. http://dx.doi.org/10.1533/9781845698379.1.3.

Full text
APA, Harvard, Vancouver, ISO, and other styles
8

"Carbon Radicals." In The Chemical Biology of Carbon, 402–37. The Royal Society of Chemistry, 2023. http://dx.doi.org/10.1039/bk9781839169502-00402.

Full text
Abstract:
Chapter 12 turns from heterolytic C–C bond formations in vivo to homolytic pathways at C–H and C–C bonds that involve carbon-centered radical species rather than carbanions and carbocations. One extreme is reaction of alkane and alkene carbons with high valent oxo-iron species in the active sites of oxygenases. O2 is an obligate one electron acceptor in chemical biology. In oxygenase active sites high valent iron species cleave C–H bonds of bound substrates by hydrogen atom transfer to yield carbon radicals that are capturable by an [OH˙] equivalent in a radical rebound step. At the other end of the oxygen spectrum, substrate radicals occur under anaerobic conditions. More than 700 000 predicted protein open reading frames are proposed to cleave bound S-adenosylmethionine homolytically. The resultant 5′-deoxyadenosyl radical initiates C–H bond cleavage in a nearby bound cosubstrate to set off substrate radical chemistries.
APA, Harvard, Vancouver, ISO, and other styles
9

Martinho Simões, José A., and Manuel Minas da Piedade. "Electrochemical Measurements." In Molecular Energetics. Oxford University Press, 2008. http://dx.doi.org/10.1093/oso/9780195133196.003.0020.

Full text
Abstract:
Electrochemical measurements have been playing an increasingly important role in the thermodynamic study of reactions in solution, not only because they provide data that are difficult (or even impossible) to obtain by other methods but also because these data can often be compared with the values determined for the analogous gas-phase reactions, thus yielding information on solvation energetics. Figure 16.1 was adapted from a scheme proposed by Griller et al. It summarizes the thermochemical information on the R–X bond that can be probed by electrochemical methods. The vertical arrows represent homolytic cleavages, and the horizontal arrows depict reduction or oxidation potentials. The authors have appropriately called the scheme in figure 16.1 a “mnemonic,” rather than a “thermochemical cycle,” because not all arrow combinations define thermochemical cycles. This can be made more clear by inspecting figure 16.2, where true thermochemical cycles are defined. For example, the enthalpy of reaction 7 is not the sum of the enthalpies of reactions 1 and 4 (as might be suggested by figure 16.1) but their sum minus the enthalpy of reaction 12. In fact, true thermochemical cycles in figure 16.1 can only be defined by considering parallelograms confined either to the upper or the lower part of the mnemonic. For instance, the enthalpy of reaction 7 is given by the enthalpy of reaction 4 plus the enthalpy of reaction 9 minus the enthalpy of reaction 3, but it is not equal to the enthalpy of reaction 6 minus the enthalpy of reaction 11 plus the enthalpy of reaction 10. Also, the enthalpy of reaction 1 (the homolytic dissociation of the R–X bond in the neutral molecule RX) can be given by the sum of the enthalpies or reaction 5 and 11 minus the enthalpy of reaction 3 or, for example, by the sum of the enthalpies of reactions 7 and 12 minus the enthalpy of reaction 4. The attractive feature of the mnemonic in figure 16.1 (or the thermochemical cycles in figure 16.2) is that it depicts the seven possible R–X cleavage reactions of RX, RX−, and RX+, as well as their relationships.
APA, Harvard, Vancouver, ISO, and other styles
10

Frey, Perry A., and Adrian D. Hegeman. "Decarboxylation and Carboxylation." In Enzymatic Reaction Mechanisms. Oxford University Press, 2007. http://dx.doi.org/10.1093/oso/9780195122589.003.0012.

Full text
Abstract:
Decarboxylation is an essential process in catabolic metabolism of essentially all nutrients that serve as sources of energy in biological cells and organisms. The most widely known biological process leading to decarboxylation is the metabolism of glucose, in which all of the carbon in the molecule is oxidized to carbon dioxide by way of the glycolytic pathway, the pyruvate dehydrogenase complex, and the tricarboxylic acid cycle. The decarboxylation steps take place in thiamine pyrophosphate (TPP)–dependent α-ketoacid dehydrogenase complexes and isocitrate dehydrogenase. The latter enzyme does not require a coenzyme, other than the cosubstrate NAD+. Many other decarboxylations require coenzymes such as pyridoxal-5'-phosphate (PLP) or a pyruvoyl moiety in the peptide chain. Biological carboxylation is the essential process in the fixation of carbon dioxide by plants and of bicarbonate by animals, plants, and bacteria. Carboxylation by enzymes requires the action of biotin or a divalent metal cofactor, and it requires ATP when the carboxylating agent is the bicarbonate ion. The most prevalent enzymatic carboxylation is that of ribulose bisphosphate carboxylase (rubisco), which is responsible for carbon dioxide fixation in plants. The basic chemistry of decarboxylation is illustrated by mechanisms A to D in fig. 8-1. The mechanisms all require some means of accommodation for the electrons from the cleavage of the bond linking the carboxylate group to the α-carbon. In mechanism A, an electron sink at the β-carbon provides a haven for two electrons. Acetoacetate decarboxylase functions by this mechanism (see chap. 1), as well as PLP- and TPP-dependent decarboxylases (see chap. 3). In mechanism B, a leaving group at the β-carbon departs with two electrons. Mevalonate-5-diphosphate decarboxylate functions by mechanism B and is discussed in a later section. In mechanism C, a leaving group replaces the α-carbon and departs with a pair of electrons. A biological example is formate dehydrogenase, in which the leaving group is a hydride that is transferred to NAD+. In mechanism D, a free radical center is created adjacent to the α-carbon and potentiates the homolytic scission of the bond to the carboxylate group. Mechanism D requires secondary electron transfer processes to create the radical center and quench the formyl radical.
APA, Harvard, Vancouver, ISO, and other styles

Conference papers on the topic "Homolytic cleavage of diborons"

1

Jeremić, Svetlana R., Jelena R. Đorović Jovanović, Marijana S. Stanojević Pirković, and Zoran S. Marković. "THERMODYNAMICALLY INVESTIGATIONS OF FREE RADICAL SCAVENGER POTENCY OF 1,2,4-TRIHYDROXYTHIOXANTHONE." In 1st INTERNATIONAL Conference on Chemo and BioInformatics. Institute for Information Technologies, University of Kragujevac, 2021. http://dx.doi.org/10.46793/iccbi21.414j.

Full text
Abstract:
The operative mechanism of the antioxidative action of 1,2,4-trihydroxythioxanthone (TX) is investigated in this contribution. Conclusions are made based on enthalpy values, as thermodynamical parameters. All calculations are done using the M06-2X/6-311++G(d,p) level of theory. To imitate polar and non-polar environments, calculations are done in water and benzene as the medium. It is found that, among three possible radicals that TX can generate, the most stable is the one obtained by homolytic cleavage of the O-H group in position 4. It was found that HAT (Hydrogen Atom Transfer) is the most plausible mechanism for that purpose in benzene. On the other hand, the most favorable mechanism in water is SPLET (Sequential Proton Loss Electron Transfer). Here is estimated the capacity of TX to deactivate hydroxyl (HO●), hydroperoxyl (HOO●) and methylperoxyl radical (CH3OO●). It is found that TX can deactivate all three free radicals following HAT and SPLET reaction mechanisms competitively, in the polar and non-polar environment. SET-PT (Single-Electron Transfer followed by Proton Transfer) is the inoperative mechanism for radicals scavenging, in the polar and non-polar environment.
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