首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
Thermal or photolytic reactions of bioactive S‐nitrosothiols and related thiols in the presence of radical generators in deaerated DMSO or aqueous solutions under argon or saturated with nitric oxide (NO) produced nitroxides and an oxyaminyl radical, which were well characterized by EPR spectra. Nitroxides containing a thiyl substituent were obtained. Possible mechanisms are proposed. Bioactive S‐nitrosothiols such as S‐nitrosoglutathione, S‐nitroso‐N‐acetylpenicillamine and related thiols such as glutathione and N‐acetylpenicillamine were used for the investigation. Radical generators utilized as transient radical sources were 2,2‐azobisisobutyronitrile, 2,2‐azobis(2‐methylpropionamidine) dihydrochloride, tert‐butyl peroxide and benzoyl peroxide. Copyright © 2002 John Wiley & Sons, Ltd.  相似文献   

2.
Free radicals generated from UV irradiation of simple aliphatic amides in anaerobic and nitric oxide (NO)‐saturated liquid mixtures or solutions gave EPR spectra of nitroxides. The application of isotopic effects to EPR spectra and the generation of radicals by transient radical attack on substrate molecules or by photolysing amine or acetoin were used to help identify photochemically produced radicals from the amides. The aliphatic amides used were formamide, acetamide and their N‐methyl‐ or deuterium‐substituted derivatives. Transient radicals used to attack the amides via hydrogen‐atom abstraction were generated from the initiator AIBN or AAPH. The observation of various nitroxides indicates the reactivity of NO for trapping acyl, carbamoyl and other carbon‐centered radicals. Possibly mechanistic pathways diagnosed with this trap are proposed. Copyright © 2003 John Wiley & Sons, Ltd.  相似文献   

3.
The styrene polymerization initiated by benzoyl peroxide (BPO) in the presence of N‐tert‐butyl‐α‐isopropylnitrone as nitroxide precursor is well‐controlled provided that a prereaction between the nitrone and BPO is carried out in suitable conditions prior to polymerization at a higher temperature. Electron spin resonance (ESR) spectroscopy was implemented to probe the nitroxides formed during both steps, that is, the prereaction and polymerization, and to get crucial information regarding the structure of the nitroxides responsible for the polymerization control. ESR studies combined with first principles calculations have evidenced that nitroxides observed during the prereaction in the presence of styrene and during the polymerization steps consist of a mixture of two macronitroxides. One is formed by the addition of a growing polystyrene chain to the nitrone as would be expected. However, the second one results from the addition of a polystyrene chain to tert‐butyl nitroso that is in situ formed presumably by decomposition of the first macronitroxide type. © 2013 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem, 2013  相似文献   

4.
N‐Nitroso compounds containing benzene, fluorene or fluorenone rings were synthesized. Photolysis of these compounds with 312‐nm UV light provided the NO . species, the presence of which was corroborated by use of an EPR method and of 2‐phenyl‐4,4,5,5‐tetramethylimidazolin‐1‐oxyl 3‐oxide (PTIO) as a trapping agent. During irradiation of N‐methyl‐N‐nitroso‐9‐fluorenone carboxamide ( 14 c ) in the absence of PTIO, it underwent decomposition followed by recombination to give the heterocyclic nitric oxide radical 15 . Incorporation of intercalating moieties endowed the N‐nitroso compounds with DNA‐cleaving ability through single‐strand scission upon UV irradiation in a phosphate buffer (pH 5.0–8.0) under aerobic conditions.  相似文献   

5.
The mechanism of “self-spin-probing” reactions between carbanion of nitroalkanes (1) and ω-hydroperfluorodiacyl peroxides (2) initiated by electron transfer has been verified by EPR detection of new bis(ω-hydroperfluoro-dialkyl) nitroxides (11) and ω-hydroperfluoroalkyl nitroalkyl nitroxides (12) — spin adducts of radical intermediates H(CF2CF2)n · (n=1,2,3) and R1R2C·NO2 to the nitroso trap H(CFCF)nNO (9) generated in the reactions.  相似文献   

6.
A series of substituted 4‐methylphenols 1 and 2 was oxidized with PbO2 in the presence of nitroso compounds 3 – 10 . The formation of adducts of benzyl radicals with the nitroso spin traps in the reaction mixture was established, suggesting the abstraction of an H‐atom from the methyl substituent of 1 or 2 . In the consecutive steps, the adducts underwent a further rearrangement to the corresponding nitrones. When the starting phenol contained bulky tBu groups in ortho‐position (see 2,6‐di(tert‐butyl)‐4‐methylphenol ( 1a )), the stable 2,6‐di(tert‐butyl)‐4R‐phenoxy radicals (R=? CH?N+(O?)? X) were detected as the final radical products. The indirect evidence of nitrones in the reaction mixture was performed in one case by the reaction with a RO radicals.  相似文献   

7.
A kinetic analysis has been carried out for a cascade of intramolecular oxidation reactions of free radicals generated in the redox reactions of substituted cyclohexyl endoperoxides (15 compounds) with the Fe2+ ion. Each radical conversion reaction has been characterized by its enthalpy, activation energy, and rate constant. Kinetic characteristics have been calculated by the intersecting parabolas method. Depending on their structure, cyclohexyl endoperoxides generate one to three radicals. There is a linear empirical correlation between the number of radicals generated by a peroxide and its molar antimalarial activity (IC 50/M, where M is the molar mass of the peroxide). The peroxides that generate no more than one radical show no antimalarial activity.  相似文献   

8.
A new series of 2‐(p‐tolyloxy)‐3‐(5‐(pyridin‐4‐yl)‐1,3,4‐oxadiazol‐2‐yl)quinoline were synthesized from oxidative cyclization of N′‐((2‐(p‐tolyloxy)quinoline‐3‐yl)methylene)isonicotinohydrazide in DMSO/I2 at reflux condition for 3–4 h. The structures of the new compounds were confirmed by elemental analyses as well as IR, 1H‐NMR, and mass spectral data. All the synthesized compounds were screened for their antibacterial activities against various bacterial strains. Several of these compounds showed potential antibacterial activity. J. Heterocyclic Chem., (2011).  相似文献   

9.
During the decay of (15N)peroxynitrite (O?15NOO ? ) in the presence of N‐acetyl‐L ‐tyrosine (Tyrac) in neutral solution and at 268 K, the 15N‐NMR signals of 15NO and 15NO show emission (E) and enhanced absorption (A) as it has already been observed by Butler and co‐workers in the presence of L ‐tyrosine (Tyr). The effects are built up in radical pairs [CO , 15NO ]S formed by O? O bond scission of the (15N)peroxynitrite? CO2 adduct (O?15NO? OCO ). In the absence of Tyrac and Tyr, the peroxynitrite decay rate is enhanced, and 15N‐CIDNP does not occur. This is explained by a chain reaction during the peroxynitrite decay involving N2O3 and radicals NO . and NO . The interpretation is supported by 15N‐CIDNP observed with (15N)peroxynitrite generated in situ during reaction of H2O2 with N‐acetyl‐N‐(15N)nitroso‐dl ‐tryptophan ((15N)NANT) at 298 K and pH 7.5. In the presence of Na15NO2 at pH 7.5 and in acidic solution, 15N‐CIDNP appears in the nitration products of Tyrac, 1‐(15N)nitro‐N‐acetyl‐L ‐tyrosine (1‐15NO2‐Tyrac) and 3‐(15N)nitro‐N‐acetyl‐L ‐tyrosine (3‐15NO2‐Tyrac). The effects are built up in radical pairs [Tyrac . , 15NO ]F formed by encounters of independently generated radicals Tyrac . and 15NO . Quantitative 15N‐CIDNP studies show that nitrogen dioxide dependent reactions are the main if not the only pathways for yielding both nitrate and nitrated products.  相似文献   

10.
Oligodeoxyribonucleotide conjugates of ethylenediamine‐N,N,N′,N′‐tetrakis(methylenephosphonic acid) (EDTP) have been used to place a CeIII/EDTP complex in close proximity to predetermined phosphodiester linkages of a complementary target oligonucleotide. In the presence of atmospheric oxygen, the CeIII is oxidized into CeIV which, in turn, efficiently cleaves the target phosphodiester linkage. No cleavage occurs at the other single‐stranded regions, which suggests that the catalytic Ce species is strictly localized next to the target phosphodiester linkage. No decrease in the reaction rate is observed upon introduction of scavengers for hydroxyl radicals (such as DMSO or MeOH) or singlet oxygen (such as NaN3) to the system; this indicates that the reaction proceeds via a hydrolytic pathway. Any significant contribution by an oxidative pathway is further ruled out by the observation that nucleosides remain intact after incubation with CeIV/EDTP complex for extended periods.  相似文献   

11.
A synthetic method for the production of polyethylene (PE) chains carrying alkoxyamine end‐group has been proposed first by successfully reacting the well‐known 2,2,6,6‐tetramethylpiperidine‐N‐oxyl (TEMPO) and N‐(2‐methyl‐2‐propyl)‐N‐(1‐diethylphosphono‐2,2‐dimethylpropyl)‐N‐oxyl (commonly called SG1) stable radicals with dipolyethylenylmagnesium compounds to give PE‐TEMPO and PE‐SG1. Since the homolytic cleavage of these two macroalkoxyamines for the production of block copolymers using controlled radical polymerization would require temperatures higher than 160 °C, two original new nitroxides (4‐[(2,2‐dimethyl‐4‐(Ntert‐Butyl‐N‐(1‐diethoxyphosphoryl‐2,2‐dimethylpropyl)aminoxy)‐4‐n‐butoxycarbonyl)butanoyloxyl]‐2,2,6,6‐tetramethylpiperidinyl‐1‐oxy, DD1) and 4‐[(2,2‐dimethyl‐4‐(Ntert‐Butyl‐N‐(1‐diethoxyphosphoryl‐2,2‐dimethylpropyl)aminoxy)‐4‐phenyl) butanoyloxyl]‐2,2,6,6‐tetramethylpiperidinyl‐1‐oxy, DD2) containing a TEMPO moiety and incorporating an SG1‐based alkoxyamine (cleavage temperature: 60 °C) were then synthesized. NMR analyses showed that the resulting PE‐DD1 and PE‐DD2 were obtained using this functionalization strategy though with low to moderate yields (from 17% to 40%). PE‐DD2 (40% functionalization) was used under controlled radical polymerization conditions of n‐butyl acrylate. SEC analyses together with 1H NMR analysis showed that a poly(ethylene‐bn‐butyl acrylate) block copolymer was produced and that the polymerization proceeded under control. © 2007 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 45: 2705–2718, 2007  相似文献   

12.
The photophysical and photochemical properties of (OC‐6‐33)‐(2,2′‐bipyridine‐κN1,κN1′)tricarbonyl(9,10‐dihydro‐9,10‐dioxoanthracene‐2‐carboxylato‐κO)rhenium (fac‐[ReI(aq‐2‐CO2)(2,2′‐bipy)(CO)3]) were investigated and compared to those of the free ligand 9,10‐dihydro‐9,10‐dioxoanthracene‐2‐carboxylate (=anthraquinone‐2‐carboxylate) and other carboxylato complexes containing the (2,2′‐bipyridine)tricarbonylrhenium ([Re(2,2′‐bipy)(CO)3]) moiety. Flash and steady‐state irradiations of the anthraquinone‐derived ligand (λexc 337 or 351 nm) and of its complex reveal that the photophysics of the latter is dominated by processes initiated in the Re‐to‐(2,2′‐bipyridine) charge‐transfer excited state and 2,2′‐bipyridine‐ and (anthraquinone‐2‐carboxylato)‐centered intraligand excited states. In the reductive quenching by N,N‐diethylethanamine (TEA) or 2,2′,2″‐nitrilotris[ethanol] TEOA, the reactive states are the 2,2′‐bipyridine‐centered and/or the charge‐transfer excited states. The species with a reduced anthraquinone moiety is formed by the following intramolecular electron transfer, after the redox quenching of the excited state: [ReI(aq−2−CO2)(2,2′‐bipy.)(CO)3]⇌[ReI(aq−2−CO2.)(2,2′‐bipy)(CO)3] The photophysics, particularly the absence of a ReI‐to‐anthraquinone charge‐transfer excited state photochemistry, is discussed in terms of the electrochemical and photochemical results.  相似文献   

13.
The title compounds, bis­(di­methyl­form­amide)‐1κO,3κO‐bis{μ‐2,2′‐[2,2′‐di­methyl­propane‐1,3‐diyl­bis­(nitrilo­methylidyne)]­diphenolato}‐1κ4N,N′,O,O′:2κ2O,O′;2κ2O,O′:3κ4N,N′,O,O′‐di‐μ‐nitrito‐1:2κ2N:O;2:3κ2O:N‐dinickel(II)­cobalt(II), [CoNi2(NO2)2(C19H22N2O2)2(C3H7NO)2], (I), ‐copper(II), [CuNi2(NO2)2(C19H22N2O2)2(C3H7NO)2], (II), and ‐manganese(II), [MnNi2(NO2)2(C19H22N2O2)2(C3H7NO)2], (III), consist of centrosymmetric linear heterotrinuclear metal complexes. The three complexes are isostructural. There are three bridges across the Ni–M atom pairs (M is Co2+, Cu2+ or Mn2+) in each complex, involving two O atoms of a μ‐N,N′‐bis­(salicyl­idene)‐2,2′di­methyl‐1,3‐propane­diaminate ligand and an N—O moiety of a μ‐nitrito group. The coordination sphere around each metal atom, whether Co2+, Cu2+, Mn2+ or Ni2+, can be described as distorted octahedral. The Ni?M distances are 2.9988 (5) Å in (I), 2.9872 (5) Å in (II) and 3.0624 (8) Å in (III).  相似文献   

14.
A mixture of hypophosphorous, phosphorous, and phosphoric acids is formed during the anaerobic oxidation of white phosphorus by peroxides [ROOН; R = Н, 3-ClC6H4CO, (СН3)3С] in water. The rate of reactions grows considerably upon adding nonpolar organic solvents. The activity series of peroxides and solvents are determined experimentally. NMR spectroscopy shows that the main product of the reaction is phosphorous acid, regardless of the nature of the peroxide and solvent. A radical mechanism of oxidation of white phosphorus by peroxides in water is proposed. It is initiated by the homolysis of peroxide with the formation of НO? radicals that are responsible for the homolytic opening of phosphoric tetrahedrons. Further oxidation and stages of the hydrolysis of intermediate phosphorus-containing compounds yield products of the reaction.  相似文献   

15.
Bleaching of a preformed solution of the blue‐green radical cation 2,2′‐azinobis (3‐ethylbenzothizoline‐6‐sulfonic acid) (ABTS) has been extensively used to evaluate the antioxidant capacity of complex mixtures and individual compounds. The reaction of the preformed radical with free‐radical scavengers can be easily monitored by following the decay of the sample absorbance at 734 nm. The ABTS radical cation can be prepared employing different oxidants. Results obtained using MnO2 as oxidant show that the presence of manganese ions increases the rate of [ABTS] autobleaching in a concentration‐dependent manner. The radicals can also be obtained by oxidizing ABTS with 2,2‐azobis(2‐amidinopropane)hydrochloride (AAPH) or peroxodisulfate (PDS). The oxidation by AAPH takes place with a large activation energy and a low reaction order in ABTS. The data support a mechanism in which the homolysis of AAPH is the rate‐limiting step, followed by the reaction of ABTS with the peroxyl radicals produced after the azocompound thermolysis. On the other hand, the low activation energy measured employing PDS, as well as the kinetic law, are compatible with the occurrence of a bimolecular reaction between the oxidant and ABTS. Regarding the use of ABTS‐based methodologies for the evaluation of free radical scavengers, radical cations obtained employing AAPH as oxidant can be used only at low temperatures, conditions where further decomposition of the remaining AAPH is minimized. The best results are obtained with ABTS derived radicals generated in the reaction of PDS with an ABTS/PDS concentration ratio equal (or higher) to two. However, even with radicals prepared by this procedure, stoichiometric coefficients considerably larger than two are obtained for the consumption of the radical cation employing tryptophane or p‐terbutylphenol as reductants. This casts doubts on the use of ABTS‐based procedures for the estimation of antioxidant capacities. © 2002 Wiley Periodicals, Inc. Int J Chem Kinet 34: 659–665, 2002  相似文献   

16.
A new route for the synthesis of substituted 8‐methyl‐6‐phenyl‐5,6‐dihydro‐4H‐1,3,2‐benzodioxaphosphocine‐2‐oxide derivatives has been developed by using cinnamic acid and p‐cresol via condensation, reduction, and followed by phosphorylation steps. The title compounds were characterized by IR, 1H, 13C, 31P, and mass spectral studies and elemental analysis. The title compounds have been investigated for their antioxidant activity with respect to their IC50 values using 2,2‐diphenyl‐1‐picrylhydrazyl, NO radical scavenging activities, and reducing power assay. The results obtained from the aforementioned methods revealed that 2‐phenylamino derivatives have shown greater free radical scavenging activity when compared with those of the phenoxy derivatives and is attributed to the presence of secondary amino group, which is able to produce free radicals easily.  相似文献   

17.
We report a simple protocol for the photochemical Giese addition of C(sp3)‐centered radicals to a variety of electron‐poor olefins. The chemistry does not require external photoredox catalysts. Instead, it harnesses the excited‐state reactivity of 4‐alkyl‐1,4‐dihydropyridines (4‐alkyl‐DHPs) to generate alkyl radicals. Crucial for reactivity is the use of a catalytic amount of Ni(bpy)32+ (bpy=2,2′‐bipyridyl), which acts as an electron mediator to facilitate the redox processes involving fleeting and highly reactive intermediates.  相似文献   

18.
The synthesis and characterization of the first catalytic manganese N‐heterocyclic carbene complexes are reported: MnBr(N‐methyl‐N′‐2‐pyridylbenzimidazol‐2‐ylidine)(CO)3 and MnBr(N‐methyl‐N′‐2‐pyridylimidazol‐2‐ylidine)(CO)3. Both new species mediate the reduction of CO2 to CO following two‐electron reduction of the MnI center, as observed with preparative scale electrolysis and verified with 13CO2. The two‐electron reduction of these species occurs at a single potential, rather than in two sequential steps separated by hundreds of millivolts, as is the case for previously reported MnBr(2,2′‐bipyridine)(CO)3. Catalytic current enhancement is observed at voltages similar to MnBr(2,2′‐bipyridine)(CO)3.  相似文献   

19.
A series of novel substituted (E)‐N′‐benzylidene‐4‐methyl‐2‐(2‐propylpyridin‐4‐yl)thiazole‐5‐carbohydrazide derivatives ( 6a‐l ) have been synthesized by following the multistep synthetic route starting from prothionamide. The resulting compounds were characterized via 1H, 13C NMR, and HRMS spectral data. The synthesized carbohydrazides were evaluated for their in vitro antimicrobial and antioxidant activities. Tested molecules have displayed moderate to good growth inhibition activity. Among the screened compounds, 6b , 6e , 6j, and 6k are found to be the more promising antimicrobial agents. A 2,2‐diphenyl‐1‐picrylhydrazyl assay was used to test the antioxidant activity of the carbohydrazides. The carbohydrazide derivatives 6b and 6i have shown better free‐radical scavenging ability than the other investigated compounds.  相似文献   

20.
The title compound, tetrakis(μ‐2,3‐di­methoxy­benzoato)‐κ4O:O′;κ6O,O′:O′‐bis[(2,2′‐bi­pyridine‐N,N′)(2,3‐di­methoxy­benzoato‐O,O′)lanthanum(III)], [La2(2,3‐DMOBA)6(2,2′‐bpy)2], where 2,3‐DMOBA is 2,3‐di­methoxy­benzoate (C9H9O4) and 2,2′‐bpy is 2,2′‐bi­pyridine (C10H8N2), is a dimer with a centre of inversion between the La atoms bridged by four carboxyl­ate ligands. The central La atom is ennea‐coordinated and has a distorted monocapped square‐antiprism geometry.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号