Your browser doesn't support javascript.
loading
Mostrar: 20 | 50 | 100
Resultados 1 - 20 de 20
Filtrar
Más filtros










Base de datos
Intervalo de año de publicación
1.
Angew Chem Int Ed Engl ; 57(29): 9146-9149, 2018 Jul 16.
Artículo en Inglés | MEDLINE | ID: mdl-29578273

RESUMEN

In 1923, Wieland and Wingler reported that in the molecular hydrogen producing reaction of hydrogen peroxide with formaldehyde in basic solution, free hydrogen atoms (H. ) are not involved. They postulated that bis(hydroxymethyl)peroxide, HOCH2 OOCH2 OH, is the intermediate, which decomposes to yield H2 and formate, proposing a mechanism that would nowadays be considered as a "concerted process". Since then, several other (conflicting) "mechanisms" have been suggested. Our NMR and Raman spectroscopic and kinetic studies, particularly the determination of the deuterium kinetic isotope effect (DKIE), now confirm that in this base-dependent reaction, both H atoms of H2 derive from the CH2 hydrogen atoms of formaldehyde, and not from the OH groups of HOCH2 OOCH2 OH or from water. Quantum-chemical CBS-QB3 and W1BD computations show that H2 release proceeds through a concerted process, which is strongly accelerated by double deprotonation of HOCH2 OOCH2 OH, thereby ruling out a free radical pathway.

2.
J Am Chem Soc ; 138(16): 5290-8, 2016 04 27.
Artículo en Inglés | MEDLINE | ID: mdl-27023326

RESUMEN

Persistent dialkylnitroxides (e.g., 2,2,6,6-tetramethylpiperidin-1-oxyl, TEMPO) play a central role in the activity of hindered amine light stabilizers (HALS)-additives that inhibit the (photo)oxidative degradation of consumer and industrial products. The accepted mechanism of HALS comprises a catalytic cycle involving the rapid combination of a nitroxide with an alkyl radical to yield an alkoxyamine that subsequently reacts with a peroxyl radical to eventually re-form the nitroxide. Herein, we offer evidence in favor of an alternative reaction mechanism involving the acid-catalyzed reaction of a nitroxide with a peroxyl radical to yield an oxoammonium ion followed by electron transfer from an alkyl radical to the oxoammonium ion to re-form the nitroxide. In preliminary work, we showed that TEMPO reacts with peroxyl radicals at diffusion-controlled rates in the presence of acids. Now, we show that TEMPO can be regenerated from its oxoammonium ion by reaction with alkyl radicals. We have determined that this reaction, which has been proposed to be a key step in TEMPO-catalyzed synthetic transformations, occurs with k ∼ 1-3 × 10(10) M(-1) s(-1), thereby enabling it to compete with O2 for alkyl radicals. The addition of weak acids facilitates this reaction, whereas the addition of strong acids slows it by enabling back electron transfer. The chemistry is shown to occur in hydrocarbon autoxidations at elevated temperatures without added acid due to the in situ formation of carboxylic acids, accounting for the long-known catalytic radical-trapping antioxidant activity of TEMPO that prompted the development of HALS.

3.
J Am Chem Soc ; 137(1): 94-7, 2015 Jan 14.
Artículo en Inglés | MEDLINE | ID: mdl-25533605

RESUMEN

Hydrogen atom transfer is central to many important radical chain sequences. We report here a method for determination of both the primary and secondary isotope effects for symmetrical substrates by the use of NMR. Intramolecular competition reactions were carried out on substrates having an increasing number of deuterium atoms at symmetry-related sites. Products that arise from peroxyl radical abstraction at each position of the various substrates reflect the competition rates for H(D) abstraction. The primary KIE for autoxidation of tetralin was determined to be 15.9 ± 1.4, a value that exceeds the maximum predicted by differences in H(D) zero-point energies (∼7) and strongly suggests that H atom abstraction by the peroxyl radical occurs with substantial quantum mechanical tunneling.


Asunto(s)
Deuterio/química , Hidrocarburos/química , Hidrógeno/química , Cinética , Espectroscopía de Resonancia Magnética , Estructura Molecular , Oxidación-Reducción , Peróxidos/química
6.
J Org Chem ; 73(5): 1830-41, 2008 Mar 07.
Artículo en Inglés | MEDLINE | ID: mdl-18260673

RESUMEN

Hydroquinones (benzene-1,4-diols) are naturally occurring chain-breaking antioxidants, whose reactions with peroxyl radicals yield 1,4-semiquinone radicals. Unlike the 1,2-semiquinone radicals derived from catechols (benzene-1,2-diols), the 1,4-semiquinone radicals do not always trap another peroxyl radical, and instead the stoichiometric factor of hydroquinones varies widely between 0 and 2 as a function of ring-substitution and reaction conditions. This variable antioxidant behavior has been attributed to the competing reaction of the 1,4-semiquinone radical with molecular oxygen. Herein we report the results of experiments and theoretical calculations focused on understanding this key reaction. Our experiments, which include detailed kinetic and mechanistic investigations by laser flash photolysis and inhibited autoxidation studies, and our theoretical calculations, which include detailed studies of the reactions of both 1,4-semiquinones and 1,2-semiquinones with O2, provide many important insights. They show that the reaction of O2 with 2,5-di-tert-butyl-1,4-semiquinone radical (used as model compound) has a rate constant of 2.4 +/- 0.9 x 10(5) M-1 s-1 in acetonitrile and as high as 2.0 +/- 0.9 x 10(6) M-1 s-1 in chlorobenzene, i.e., similar to that previously reported in water at pH approximately 7. These results, considered alongside our theoretical calculations, suggest that the reaction occurs by an unusual hydrogen atom abstraction mechanism, taking place in a two-step process consisting first of addition of O2 to the semiquinone radical and second an intramolecular H-atom transfer concerted with elimination of hydroperoxyl to yield the quinone. This reaction appears to be much more facile for 1,4-semiquinones than for their 1,2-isomers.


Asunto(s)
Oxígeno/química , Quinonas/química , Cinética , Oxidación-Reducción , Termodinámica
7.
J Org Chem ; 72(21): 8095-8, 2007 Oct 12.
Artículo en Inglés | MEDLINE | ID: mdl-17850162

RESUMEN

Spectroscopic data are consistent with computations that show that, in their most stable conformations, the peroxyl moiety is equatorial in cyclohexylperoxyl radicals and axial in oxa- and most polyoxacyclohexyl-2-peroxyl radicals.

10.
Org Lett ; 6(23): 4319-22, 2004 Nov 11.
Artículo en Inglés | MEDLINE | ID: mdl-15524473

RESUMEN

Axial and equatorial cyclohexylacyl and tetrahydropyranyl-2-acyl radicals gave distinct EPR spectra thanks to surprisingly large beta-hydrogen atom hyperfine splittings that enabled them to be characterized and monitored. DFT computations indicated that the axial species (X = CH(2)) had a higher barrier to rotation about the (O)C(alpha)-C(beta) bond. The computed difference Delta H degrees for the axial and equatorial radicals (R = H, X = CH(2)) was 0.8 kcal mol(-)(1).

11.
J Am Chem Soc ; 126(34): 10667-75, 2004 Sep 01.
Artículo en Inglés | MEDLINE | ID: mdl-15327325

RESUMEN

The O-H bond dissociation enthalpies (BDEs) of 13 oximes, RR'C=NOH, having R and/or R' = H, alkyl, and aryl are reported. Experimental anchor points used to validate the results of theoretical calculations include (1) the O-H BDEs of (t-Bu)2C=NOH, t-Bu(i-Pr)C=NOH, and t-Bu(1-Ad)C=NOH determined earlier from the heat released in the reaction of (t-Bu)2C=NO* with (PhNH)2 in benzene and EPR spectroscopy (Mahoney, L. R.; Mendenhall, G. D.; Ingold, K. U. J. Am. Chem. Soc. 1973, 95, 8610), all of which were decreased by 1.7 kcal/mol to reflect a revision to the heat of formation of (E)-azobenzene (which has significant ramifications for other BDEs) and to correct for the heat of hydrogen bonding of (t-Bu)2C=NOH (alphaH2 = 0.43 measured in this work) to benzene, and (2) the measured rates of thermal decomposition of six RR'C=NOCH2Ph at 423 or 443 K, which were used to derive O-H BDEs for the corresponding RR'C=NOH. Claims (Bordwell, F. G.; Ji, G. Z. J. Org. Chem. 1992, 57, 3019; Bordwell, F. G.; Zhang, S. J. Am. Chem. Soc. 1995, 117, 4858; and Bordwell, F. G.; Liu, W.-Z. J. Am. Chem. Soc. 1996, 118, 10819) that the O-H BDEs in mono- and diaryloximes are significantly lower than those for alkyloximes due to delocalization of the unpaired electron into the aromatic ring have always been inconsistent with the known structures of iminoxyl radicals as are the purported perpendicular structures, i.e., phi(Calpha-C=N-O*) = 90 degrees, for sterically hindered dialkyl iminoxyl radicals. The present results confirm the 1973 conclusion that simple steric effects, not electron delocalization or dramatic geometric changes, are responsible for the rather small differences in oxime O-H BDEs.

12.
Org Biomol Chem ; 2(14): 2083-6, 2004 Jul 21.
Artículo en Inglés | MEDLINE | ID: mdl-15254636

RESUMEN

A combination of laser flash photolysis and competitive kinetic methods has been used to measure the absolute bimolecular rate constants for hydrogen atom abstraction in water from a series of fluorinated alkoxides and aldehyde hydrates by the perfluoroalkyl radical, *CF2CF2OCF2CF2SO3- Na+. The bimolecular rate constants observed for the beta-fluorinated alkoxides were in the 10(5) M(-1) s(-1) range, such rates representing enhancements (relative to the respective alcohols) of between 100 and almost 1000-fold, depending on the reactivity of the alkoxide. Likewise, the monobasic sodium salts of chloral and fluoral hydrate exhibit similar rate enhancements, relative to their respective hydrates.

13.
Org Lett ; 6(10): 1641-3, 2004 May 13.
Artículo en Inglés | MEDLINE | ID: mdl-15128256

RESUMEN

An unambiguous method for distinguishing between abstraction-addition and addition-abstraction mechanisms (and mixtures thereof) in the reaction of 4-hydroxy-2,2,6,6-tetramethylpiperidine-N-oxyl with a specifically deuterated cyclohexene, 1,2-dideuteriocyclohexene, is demonstrated.

14.
Org Biomol Chem ; 2(5): 689-94, 2004 03 07.
Artículo en Inglés | MEDLINE | ID: mdl-14985808

RESUMEN

A combination of laser flash photolysis and competitive kinetic methods have been used to measure the absolute bimolecular rate constants for hydrogen atom abstraction in water from a variety of organic substrates including alcohols, ethers, and carboxylic acids by the perfluoroalkyl radical, *CF(2)CF(2)OCF(2)CF(2)SO(3)(-) Na(+). Comparison, where possible, of these rate constants with those previously measured for analogous reactions in the non-polar organic solvent, 1,3-bis(trifluoromethyl)benzene (J. Am. Chem. Soc, 1999, 121, 7335) show that the alcohols react 2-5 times more rapidly in the water solvent and that the ethers react at the same rate in both solvents. A transition state for hydrogen abstraction that is more reminiscent of an "intimate ion pair" than a "solvent separated ion pair" is invoked to explain these modest solvent effects.

15.
Org Biomol Chem ; 2(3): 415-20, 2004 Feb 07.
Artículo en Inglés | MEDLINE | ID: mdl-14747871

RESUMEN

Thermolyses of seven dialkyl, two alkyl-aryl and two diaryl O-benzyl ketoxime ethers, R(1)R(2)C[double bond, length as m-dash]NOCH(2)Ph, have been examined in three hydrogen donor solvents: tetralin, 9,10-dihydrophenanthrene, and 9,10-dihydroanthracene. All the oxime ethers gave the products expected from homolytic scission of both the O-C bond (viz., R(1)R(2)C[double bond, length as m-dash]NOH and PhCH(3)) and N-O bond (viz., R(1)R(2)C[double bond, length as m-dash]NH and PhCH(2)OH). The yields of these products depended on which solvent was used and the rates of decomposition of the O-benzyl oxime ethers were greater in 9,10-dihydrophenanthrene and 9,10-dihydroanthracene than in tetralin. These results indicated that a reverse radical disproportionation reaction in which a hydrogen atom was transferred from the solvent to the oxime ether, followed by [small beta]-scission of the resultant aminoalkyl radical, must be important in the latter two solvents. Benzaldehyde was found to be an additional product from thermolyses conducted in tetralin. This, and other evidence, indicated that another induced decomposition mode involving abstraction of a benzylic hydrogen atom, followed by [small beta]-scission of the resulting benzyl radical, became important for some substrates. Participation by minor amounts of enamine tautomers of the oxime ethers was shown to be negligible by comparison of thermolysis data for the O-benzyloxime of bicyclo[3.3.1]nonan-9-one, which cannot give an enamine tautomer, with that of the O-benzyloxime of cyclohexanone.

16.
Chem Res Toxicol ; 16(9): 1118-23, 2003 Sep.
Artículo en Inglés | MEDLINE | ID: mdl-12971799

RESUMEN

Earlier work (Paul, T., et al. (2000) Biochemistry 39, 4129-4135) has demonstrated that the water soluble positively charged peroxyl radical, (H(2)N)(2)(+)CC(CH(3))(2)OO(*) ((+)AOO(*)), caused direct strand scission of the Escherichia coli plasmid supercoiled DNA, pBR 322, with ca. 50% scission occurring at a (+)AOO(*)/base pair (bp) ratio of 0.2. There was no measurable direct scission with a negatively charged peroxyl ((-)BOO(*)) at (-)BOO(*)/bp = 24, nor with a neutral peroxyl (COO(*)) at COO(*)/bp = 5. Base modification (BM) of the same DNA by the same peroxyls has now been investigated using four base excision repair (BER) glycosylases. At (+)AOO(*)/bp = 0.04, there is 10% direct strand scission, and the Fpg protein recognized an additional 25% BM, while endonuclease (Endo) IV recognized an additional 20% BM and the other two BER enzymes did not give statistically significant BMs. None of the BER enzymes showed BMs in the DNA treated with -BOO(*). However, Fpg and Endo IV showed that at COO(*)/bp = 3.4 there was a BM comparable to that observed at (+)AOO(*)/bp = 0.04. Thus, COO(*) radicals are only ca. 1.2% as reactive toward the DNA's bases as (+)AOO(*). These results underline the importance of Coulombic forces in DNA reactions. It is also proposed that (+)AOO(*) has a higher intrinsic reactivity in H-atom abstractions and electron transfer processes than (-)BOO(*) or COO(*) radicals.


Asunto(s)
Daño del ADN/efectos de los fármacos , Enzimas Reparadoras del ADN/metabolismo , ADN Superhelicoidal/efectos de los fármacos , Estrés Oxidativo , Superóxidos/efectos adversos , Amidinas/efectos adversos , Reparación del ADN/efectos de los fármacos , Reparación del ADN/fisiología , Escherichia coli/genética , Escherichia coli/metabolismo , Glicosilación , Solubilidad , Superóxidos/química , Agua
17.
Biochem Biophys Res Commun ; 309(4): 823-9, 2003 Oct 03.
Artículo en Inglés | MEDLINE | ID: mdl-13679047

RESUMEN

The synapsins (I, II, and III) comprise a family of peripheral membrane proteins that are involved in both regulation of neurotransmitter release and synaptogenesis. Synapsins are concentrated at presynaptic nerve terminals and are associated with the cytoplasmic surface of synaptic vesicles. Membrane-binding of synapsins involves interaction with both protein and lipid components of synaptic vesicles. Synapsin I binds rapidly and with high affinity to liposomes containing anionic lipids. The binding of bovine synapsin I to liposomes was studied using fluoresceinphosphatidyl-ethanolamine (FPE) to measure membrane electrostatic potential. Synapsin binding to liposomes caused a rapid increase in FPE fluorescence, indicating an increase in positive charge at the membrane surface. Synapsin I binding to monolayers resulted in a substantial increase in monolayer surface pressure. At higher initial surface pressures, the synapsin-induced increase in monolayer surface pressure is dependent on the presence of anionic lipids in the monolayer. Synapsin I also induced rapid aggregation of liposomes, but did not induce leakage of entrapped carboxyfluorescein, while other aggregation-inducing agents promoted extensive leakage. These results are in agreement with the presence of amphipathic stretches of amino acids in synapsin I that exhibit both electrostatic and hydrophobic interactions with membranes, and offer a molecular explanation for the high affinity binding of synapsin I to liposomes and for stabilization of membranes by synapsin I.


Asunto(s)
Sinapsinas/metabolismo , Fusión Celular , Membrana Celular/metabolismo , Colorantes Fluorescentes , Liposomas , Fosfolípidos/metabolismo , Vesículas Sinápticas/metabolismo
18.
J Biol Chem ; 278(40): 38384-94, 2003 Oct 03.
Artículo en Inglés | MEDLINE | ID: mdl-12878609

RESUMEN

The inhibition of post-Amadori advanced glycation end product (AGE) formation by three different classes of AGE inhibitors, carbonyl group traps, chelators, and radical-trapping antioxidants, challenge the current paradigms that: 1) AGE inhibitors will not increase the formation of any AGE product, 2) transition metal ions are required for oxidative formation of AGE, and 3) screening AGE inhibitors only in systems containing transition metal ions represents a valid estimate of potential in vivo mechanisms. This work also introduces a novel multifunctional AGE inhibitor, 6-dimethylaminopyridoxamine (dmaPM), designed to function as a combined carbonyl trap, metal ion chelator, and radical-trapping antioxidant. Other AGE inhibitors including pyridoxamine, aminoguanidine, o-phenylenediamine, dipyridoxylamine, and diethylenetriaminepentaacetic acid were also examined. The results during uninterrupted and interrupted ribose glycations show: 1) an unexpected increase in the yield of pentosidine in the presence of radical-trapping phenolic antioxidants such as Trolox and dmaPM, 2) significant formation of Nepsilon-carboxymethyllysine (CML) in the presence of strong chelators and phenolic antioxidants, which implies that there must be nonradical routes to CML, 3) prevention of intermolecular cross-links with radical-trapping inhibitors, and 4) that dmaPM shows excellent inhibition of AGE. Glucose glycations reveal the expected inhibition of pentosidine and CML with all compounds tested, but in a buffer free of trace metal ions the yield of CML in the presence of radical-trapping antioxidants was between the metal ion-free and metal ion-containing controls. Protein molecular weight analyses support the conclusion that Amadori decomposition pathways are constrained in the presence of metal ion chelators and radical traps.


Asunto(s)
Antioxidantes/farmacología , Arginina/análogos & derivados , Arginina/química , Productos Finales de Glicación Avanzada/antagonistas & inhibidores , Lisina/análogos & derivados , Lisina/química , Piridoxamina/química , Bioquímica/métodos , Cromatografía Líquida de Alta Presión , Reactivos de Enlaces Cruzados/farmacología , Relación Dosis-Respuesta a Droga , Cromatografía de Gases y Espectrometría de Masas , Glucosa/metabolismo , Iones , Modelos Químicos , Ribosa/química , Espectrometría de Masa por Láser de Matriz Asistida de Ionización Desorción , Factores de Tiempo
19.
Free Radic Biol Med ; 34(1): 77-83, 2003 Jan 01.
Artículo en Inglés | MEDLINE | ID: mdl-12498982

RESUMEN

We have reexamined claims that melatonin directly scavenges hydrogen peroxide and shown them to be unfounded. Relative hydrogen peroxide concentrations were determined in the absence and presence of melatonin using both an isoluminol-based chemiluminescence assay (with measurements at circa 40 s, 6 h, and 24 h after mixing) and the phenol red/horseradish peroxidase assay employed by two earlier groups of workers (with measurements at 5 s, then every minute for the first 5 min, and then every hour to 5 h). Both assay procedures were in agreement. There was no significant change in the hydrogen peroxide concentrations over 24 h, and, furthermore, the concentrations of H(2)O(2) in the presence and absence of melatonin were the same within experimental error. Our results were obtained in metal ion-free systems. It therefore appears likely that the claims for a direct melatonin/H(2)O(2) reaction were due to contamination by traces of transition metal ions.


Asunto(s)
Depuradores de Radicales Libres/farmacología , Peróxido de Hidrógeno/química , Melatonina/farmacología , Peroxidasa de Rábano Silvestre/química , Humanos , Mediciones Luminiscentes , Fenolsulfonftaleína/química
SELECCIÓN DE REFERENCIAS
DETALLE DE LA BÚSQUEDA
...