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.
Previous studies have shown that the carcinogen N-hydroxy-2-acetylaminofluorene is converted by one-electron oxidants to a free nitroxide radical which dismutates to N-acetoxy-2-acetylaminofluorene and 2-nitrosofluorene. The present study shows that the same oxidation can be achieved with horseradish peroxidase and H2O2. The free radical intermediate was detected by its ESR signal, and the yields of N-acetoxy-2-acetylaminofluorene and of 2-nitrosofluorene were determined under a number of conditions. Addition of tRNA to the reaction mixture containing N-acetoxy-N-2-acetyl[2′-3H]aminofluorene yielded tRNA-bound radioactivity; addition of guanosine yielded a reaction product which appears to be N-guanosin-8-yl)-2-acetylaminofluorene. The latter compound has previously been identified as a reaction product of N-acetoxy-2-acetylaminofluorene and guanosine. Preliminary attempts to demonstrate the formation of a nitroxide free radical or its dismutation products with rat liver mixed function oxidase systems were not successful.  相似文献   

2.
C.L. Greenstock  R.W. Miller 《BBA》1975,396(1):11-16
The rate of reaction between superoxide anion (O¯.2) and 1,2-dihydroxybenzene-3,5-disulfonic acid (tiron) was measured with pulse radiolysis-generated O¯.2. A kinetic spectrophotometric method utilizing competition betweenp-benzoquinoneand tiron for O¯.2 was employed. In this system, the known rate of reduction ofp-benzoquinonewas compared with the rate of oxidation of tiron to the semiquinone. From the concentration dependence of the rate of tiron oxidation, the absolute second order rate constant for the reaction was determined to be 5 · 108 M?·s?1. Ascorbat reduced O¯.2 to hydrogen peroxide with a rate constant of 108 M?1 · s?1 as determined by the same method. The tiron semiquinone may be used as an indicator free radical for the formation of superoxide anion in biological systems because of the rapid rate of oxidation of the catechol by O¯.2 compared to the rate of O¯.2 formation in most enzymatic systems.Tiron oxidation was used to follow the formation of superoxide anion in swollen chloroplasts. The chloroplasts photochemically reduced molecular oxygen which was further reduced to hydrogen peroxide by tiron. Tiron oxidation specifically required O¯.2 since O2 was consumed in the reaction and tiron did not reduce the P700 cation radical or other components of Photosystem I under anaerobic conditions.  相似文献   

3.
The interaction of chlorophylls a and b with electrochemically prepared superoxide anion was studied in aprotic solvent. It was found that O2?·causes the deprotonation at carbon C-10 of ring V and production of chlorophyll enolate ions. The intermediate anions undergo rapid oxidation into corresponding chlorins. Pyrochlorophyll a, which lacks the C-10 carboxymethyl group, did not show the transformation. It is suggested that more strong free radical oxidants (e.g., HO2·, or RO2·) are capable of abstracting the hydrogen atom at C-10. The possible significance of free radical deprotonation and oxidation in chlorophyll allomerization is discussed.  相似文献   

4.
We have used the spin trap 5,5-dimethyl-1-pyrroline-1-oxide to determine if primary free radicals are involved in the hematin-cumene hydroperoxide system which has been shown to oxidize N-hydroxy-2-acetylaminofluorene into the nitroxyl free radical form of this carcinogen. We have found that the spin trap was oxidized itself rather than trapping either primary free radicals or carcinogen free radicals.  相似文献   

5.
Hiroshi Seki  Masashi Imamura 《BBA》1981,635(1):81-89
The reactions of ferrocytochrome c with Br?2, (SCN)?2, N3 and OH radicals were followed by measuring the change in the optical spectra of cytochrome c on γ-irradiation as well as the rate of change of absorbance upon pulse irradiation.Ferrocytochrome c is oxidized to ferricytochrome c by Br?2, (SCN)?2 or N3 radical with an efficiency of about 100% through a second-order process in which no intermediates were observed. The rate constants in neutral solutions at I = 0.073 are 9.7 · 108 M?1 · s?1, 7.9 · 108 M?1 · s?1, 1.3 · 109 M?1 · s?1 for the oxidation by Br?2, (SCN)?2 and N3 radicals, respectively. The rate constants do not vary appreciably in alkaline solutions (pH 8.9). The ionic strength dependence was observed for the rate constants of the oxidation by Br?2 and (SCN)?2. Those rate constants estimated on the assumption that the radicals react only with the amino acid residues with the characteristic steric correction factors were less than one-tenth of the observed ones. These results suggest that the partially exposed region of the heme is the probable site of electron transfer from ferrocytochrome c to the radical.Hydroxyl radicals also oxidize ferrocytochrome c with a high rate constant (k > 1 · 1010 M?1 · s?1), but with a very small efficiency (5%).  相似文献   

6.
We previously reported that high micromolar concentrations of nitric oxide were able to oxidize mitochondrial cytochrome c at physiological pH, producing nitroxyl anion (Sharpe and Cooper, 1998 Biochem. J. 332, 9–19). However, the subsequent re-evaluation of the redox potential of the NO/NO- couple suggests that this reaction is thermodynamically unfavored. We now show that the oxidation is oxygen-concentration dependent and non stoichiometric. We conclude that the effect is due to an oxidant species produced during the aerobic decay of nitric oxide to nitrite and nitrate. The species is most probably nitrogen dioxide, NO2? a well-known biologically active oxidant. A simple kinetic model of NO autoxidation is able to explain the extent of cytochrome c oxidation assuming a rate constant of 3 × 106 M-1 s-1 for the reaction of NO2? with ferrocytochrome c. The importance of NO2? was confirmed by the addition of scavengers such as urate and ferrocyanide. These convert NO2? into products (urate radical and ferricyanide) that rapidly oxidize cytochrome c and hence greatly enhance the extent of oxidation observed. The present study does not support the previous hypothesis that NO and cytochrome c can generate appreciable amounts of nitroxyl ions (NO- or HNO) or of peroxynitrite.  相似文献   

7.
We have demonstrated that the carcinogen 2-nitrosofluorene (NOF) reacts with rat liver microsomal membranes to produce a nitroxyl free radical form of the carcinogen, designated N-?-LAF. We conclude that NOF adds to the double bond of the membrane lipids in a pseudo Diels-Alder reaction. This conclusion is based on studies involving 2,3-dimethyl-2-butene and NOF. NOF reacts with this simple unsaturated hydrocarbon to produce a stable nitroxyl free radical in a pseudo Diels-Alder reaction. NOF adds to liposomes formed from lipids extracted from the rat liver microsomes to produce a free radical identical to that produced with microsomes. NOF forms the same amount of N-?-LAF at the same rate in heated microsomes as in unheated microsomes. The observations indicate the involvement of only the lipid fraction in the reaction of NOF with membranes. The amount of N-?-LAF formed increases hyperbolically in microsomes and liposomes as a function of NOF added. The amount of N-?-LAF formed reaches a maximum, at which point the amount of free radical present is about 1% of both the amount of NOF added and the amount of phospholipid present. The half-maxima of the amount of N-?-LAF formed occurs at 50 μm NOF in liposomes but at 100 μm in microsomes. The electron spin resonance spectrum of N-?-LAF indicates that this nitroxyl free radical is in a rigidly fixed position in the membrane. N-?-LAF is reduced by NADPH. There appears to be a direct chemical reduction as well as an enzyme-mediated mechanism involving NADPH-potentiated electron flow in microsomes. The reduced compound is reoxidized by ferricyanide added to microsomal membranes.  相似文献   

8.
《Luminescence》2003,18(1):49-57
The chemiluminescence reaction of lucigenin (Luc2+?2NO3?, N,N′‐dimethyl‐9,9′‐biacridinium dinitrate) at gold electrodes in dioxygen‐saturated alkaline aqueous solutions (pH 10) was investigated in detail by the use of electrochemical emission spectroscopy. We noted that both O2 and Luc2+ are reduced on a gold electrode in aqueous solution of pH 10 in almost the same potential region. From this fact, we expected chemiluminescence based on a radical–radical coupling reaction of superoxide ion (O2·?) and one‐electron reduced form of Luc2+ (Luc·+, a radical cation). Chemiluminescence was actually observed in the potential range where O2 and Luc2+ were simultaneously reduced at the electrodes. The effects were examined upon addition of enzymes, i.e. superoxide dismutase (SOD) and catalase, into the solution and the substitution of heavy water (D2O) for light water (H2O) as a solvent on the chemiluminescence. In the presence of native and active SOD, chemiluminescence was completely absent. On the other hand, chemiluminescence was observed, unchanged in the presence of either denatured and inert SOD or catalase. In addition, the amount of chemiluminescence in D2O solution was about three times greater than that in H2O solution. These results, together with cyclic voltammetric results, suggest that O2·? participates directly in the chemiluminescence but H2O2 does not, and the chemiluminescence results from the coupling reaction between O2·? and Luc·+ under the present experimental conditions. These chemically unstable species, O2·? and Luc·+, are produced during the simultaneous electroreduction of O2 and Luc2+. The coupling reaction between those radical species would lead to the formation of a dioxetane‐type intermediate and, finally, to chemiluminescence. The chemiluminescence reaction mechanism is discussed. Copyright © 2002 John Wiley & Sons, Ltd.  相似文献   

9.
The aqueous decomposition of the iridium coordinated nitrosothiols (RSNOs) trans-K[IrCl4(CH3CN)NOSPh] (1), and K2[IrCl5(NOECyS)] (2, ECyS = cysteine ethyl ester), was studied by MS analysis of the gaseous products, ESI-MS, NMR, and UV-Vis spectroscopy. Bent NO (NO, nitroxyl anion), sulfenic acids and nitrite were observed as coordinated products in solution, while nitrous oxide (N2O) and nitrogen were detected in the gas phase. The formation of coordinated NO and N2O, a nitroxyl dimerization product, allows us to propose the formation of free nitroxyl (HNO) as an intermediate. Complex 1 decomposes 300 times slower than free PhSNO does. In both cases (1 and 2) kinetic results show a first order decomposition behavior and a very negative ΔS, which strongly indicates an associative rate-determining step. A proposed decomposition mechanism, supported by the experimental data and DFT calculations, involves, as the first step, nucleophilic attack of H2O on to the sulfur atom of the coordinated RSNO, producing an NO complex and free sulfenic acid, followed by two competing reactions: a ligand exchange reaction of this NO with the sulfenic acid or, to a minor extent, coordination of N2O to produce an NO/N2O complex which finally renders free N2 and coordinated NO2. Some of the produced NO is likely to be released from the metal center producing nitroxyl by protonation and finally N2O by dimerization and loss of H2O. In conclusion, the decomposition of these coordinated RSNOs occurs through a different mechanism than for the decomposition of free RSNOs. It involves the formation of sulfenic acids and coordinated NO, which is released from the complexes and protonated at the reaction pH producing nitroxyl (HNO), and ultimately N2O.  相似文献   

10.
The objective of this study was to determine the effect of (bi)sulfite (hydrated sulfur dioxide) on human neutrophils and the ability of these immune cells to produce reactive free radicals due to (bi)sulfite oxidation. Myeloperoxidase (MPO) is an abundant heme protein in neutrophils that catalyzes the formation of cytotoxic oxidants implicated in asthma and inflammatory disorders. In this study sulfite (?SO3?) and sulfate (SO4??) anion radicals are characterized with the ESR spin-trapping technique using 5,5-dimethyl-1-pyrroline N-oxide (DMPO) in the reaction of (bi)sulfite oxidation by human MPO and human neutrophils via sulfite radical chain reaction chemistry. After treatment with (bi)sulfite, phorbol 12-myristate 13-acetate-stimulated neutrophils produced DMPO–sulfite anion radical, –superoxide, and –hydroxyl radical adducts. The last adduct probably resulted, in part, from the conversion of DMPO–sulfate to DMPO–hydroxyl radical adduct via a nucleophilic substitution reaction of the radical adduct. This anion radical (SO4??) is highly reactive and, presumably, can oxidize target proteins to protein radicals, thereby initiating protein oxidation. Therefore, we propose that the potential toxicity of (bi)sulfite during pulmonary inflammation or lung-associated diseases such as asthma may be related to free radical formation.  相似文献   

11.
The spectrum of mutations induced by the carcinogen N-2-acetylaminofluorene (AAF) was analysed in Saccharomyces cerevisiae using a forward mutation assay, namely the inactivation of the URA3 gene. The URA3 gene, carried on a yeast/bacterial shuttle vector, was randomly modified in vitro using N-acetoxy-N-2-acetylaminofluorene (N-AcO-AAF) as a model reactive metabolite of the carcinogen AAF. The binding spectrum of AAF to the URA3 gene was determined and found to be essentially random, as all guanine residues reacted about equally well with N-AcO-AAF. Independent Ura? mutants were selected in vivo after transformation of the modified plasmid into a ura3Δ yeast strain. Plasmid survival decreased as a function of AAF modification, leading to one lethal hit (37% relative survival) for an average of ≈ 50 AAF adducts per plasmid molecule. At this level of modification the mutation frequency was equal to ≈ 70 × 10?4, i.e. ≈ 50-fold above the background mutation frequency. UV irradiation of the yeast cells did not further stimulate the mutagenic response, indicating the lack of an SOS-like mutagenic response in yeast. Sequence analysis of the URA3 mutants revealed ≈ 48% frameshifts, 44% base substitutions and ≈ 8 % complex events. While most base substitutions (74%) were found to be targeted at G residues where AAF is known to form covalent C8 adducts, frameshift mutations were observed at GC base pairs in only≈ 24% of cases. Indeed, more than 60% of frameshift events occurred at sequences such as 5′-(A/T)nG-3′ where a short (n = 2 or 3) monotonous run of As or Ts is located on the 5' side of a guanine residue. We refer to these mutations as semi-targeted events and present a potential mechanism that explains their occurrence.  相似文献   

12.
The hypervalent muscle pigment ferrylmyoglobin, formed by activation of metmyoglobin by hydrogen peroxide, was found to be reduced in a second-order reaction by N-tert-butyl-α-phenylnitrone (PBN, often used as a spin trap). In acidic aqueous solution at ambient temperature, the reduction is relatively slow (δH? = 65 ± 2 kJ · mol-1 and δS? = -54 ± 7 J · mol-1. K-1 for pH = 5.6), but phase transitions during freezing of the buffered solutions accelerates the reaction between ferrylmyoglobin and PBN. In these heterogenous systems at low temperature (but not when ice-formation was inhibited by glycerol), a PBN-derived radical intermediate was detected by ESR-spectroscopy, identified as a nitroxyl radical by a parallel nitrogen hyperfine coupling constant of 31.8 G, and from microwave power saturation behavior concluded not to be located in the heme-cleft of the protein. The acceleration of the reaction is most likely caused by a lowering of the pH during the freezing of the buffered solutions whereby ferrylmyoglobin becomes more oxidizing.  相似文献   

13.
Electron transfer between horse heart and Candida krusei cytochromes c in the free and phosvitin-bound states was examined by difference spectrum and stopped-flow methods. The difference spectra in the wavelength range of 540–560 nm demonstrated that electrons are exchangeable between the cytochromes c of the two species. The equilibrium constants of the electron transfer reaction for the free and phosvitin-bound forms, estimated from these difference spectra, were close to unity at 20°C in 20 mM Tris-HCl buffer (pH 7.4). The electron transfer rate for free cytochrome c was (2–3) · 104 M?1 · s?1 under the same conditions. The transfer rate for the bound form increased with increase in the binding ratio at ratios below half the maximum, and was almost constant at higher ratios up to the maximum. The maximum electron exchange rate was about 2 · 106 M?1 · s?1, which is 60–70 times that for the free form at a given concentration of cytochrome c. The activation energy of the reaction for the bound cytochrome c was equal to that for the free form, being about 10 kcal/mol. The dependence of the exchange rate on temperature, cytochrome c concentration and solvent viscosity suggests that enhancement of the electron transfer rate between cytochromes c on binding to phosvitin is due to increase in the collision frequency between cytochromes c concentrated on the phosvitin molecule.  相似文献   

14.
Fluorescence yield dependence on external magnetic field (0–600 G) was measured for chlorophyll-protein complexes enriched with Photosystem I. Maximal relative changes of fluorescence yield at room temperature (1.0–2.5%) were dependent on the chlorphyll a:P-700 ratio. Magnetic field-induced changes were observed only in the presence of dithionite. At low temperatures (down to ?160°C) the magnetic field-induced effect decreased. The effect is obviously connected with the functions of reaction centers in Photosystem I. An explanation of the effect is proposed based on the hypothesis of radical pairs recombination within the reaction center. For the radical pair (P-700 A), an intermediate acceptor, A, with a g-value approximately equal to that of P-700 is proposed.  相似文献   

15.
Transient spectra and kinetic data of Tiron (1,2-dihydroxybenzene-3,5-disulphonic acid) are reported, obtained after pulse-radiolytic oxidation by hydroxyl radicals (°OH), superoxide anions (O2?) or a combination of both oxygen radicals. The rate constant with °OH radicals was determined at 1.0·109 M?1·s?1. Contrary to a previous report (Greenstock, C.L. and Miller, R.W. (1975) Biochim. Biophys. Acta 396, 11–16), the rate constant with O2? of 1.0·107 M?1·s?1 is lower by one order of magnitude; also the semiquinone absorbs at 300 nm rather than at 400 nm. The ratio of the rate constants with °OH and O2? of 100 again demonstrates that any oxidation reaction by the latter radical is unspecific due to the more efficient reaction of °OH radicals, leading to the same products with catechol compounds.  相似文献   

16.
Kinetic study of the reaction between tocopheroxyl (vitamin E radical) and reduced ubiquinone, n = 10) has been performed. The rates of reaction of ubiquinol with α-tocopheroxyl 1 and seven kinds of alkyl substituted tocopheroxyl radicals 2–8 in solution have been determined spectrophotometrically, using a stopped-flow technique. The result shows that the rate constants decrease as the total electron-donating capacity of the alkyl substituents on the aromatic ring of tocopheroxyls increases. For the tocopheroxyls with two alkyl substituents at ortho positions (C-5 and C-7), the second-order rate constants, k1, obtained vary i n the order of 102, and decrease predominantly, as the size of two ortho-alkyl groups (methyl, ethyl, isopropyl and tert-buty) in tocopheroxyl increases. On the other hand, the reaction between tocopheroxyl and ubiquinone-10 (oxidized ubiquinone) has not been observed. The result indicates that ubiquinol-10 regenerates tocopherol by donating a hydrogen atom of the 1-OH and/or 4-OH group to the tocopheroxyl radical. For instance, the k1 values obtained for α-tocopheroxyl are 3.74 · 105 M?1 · s?1 and 2.15 · 5 M?1 · s?1 in benzene and ethanol solution at 25°C, respectively. The above reaction rates, k1, obtained were compared with those of vitamin C with α-tocopheroxyl reported by Packer et al. (k2 = 1.55 · 106 M?1 · s?1) and Scarpa et al. (K2 = 2 · 105 105 M?1 · s?1), which is well known as a usual regeneration reaction of tocopheroxyl in biomembrane systems. The result suggests that ubiquinol-10 also regenerates the tocopheroxyl to tocopherol and prevents lipid peroxidation in various tissues and mitochondria.  相似文献   

17.
A method utilizing high pressure liquid chromatography has been developed for the separation and quantitative estimation of all the major metabolites of the carcinogen 2-acetylaminofluorene in a single chromatographic determination. The method was used to separate 7-hydroxy-2-acetylaminofluorene, 5-hydroxy-2-acetylaminofluorene, 3-hydroxy-2-acetylaminofluorene, 1-hydroxy-2-acetyl-aminofluorene, 2-aminofluorene, N-hydroxy-2-acetylaminofluorene, and 2-acetylaminofluorene when 2-acetylaminofluorene was incubated with mouse liver microsomes and NADPH.This new high pressure liquid chromatography method for separating the metabolites arising from hydroxylations of 2-acetylaminofluorene should also prove useful in the isolation and quantitative analysis of metabolites from other N-acetylarylamines.  相似文献   

18.
Acetylene inhibition of nitrous oxide reduction by denitrifying bacteria   总被引:1,自引:0,他引:1  
Acetylene (0.1 atm) caused complete or almost complete inhibition of reduction of N2O by whole cell suspensions of Pseudomonas perfectomarinus, P. aeruginosa and Micrococcus denitrificans. Acetylene did not inhibit reduction of NO3? or NO2? by these organisms. In the presence of acetylene there was stoichiometric conversion of NO3? or NO2? to N2O with negligible subsequent reduction of the latter. In the absence of acetylene there was no or only transient accumulation of N2O. The data are consistent with the view that N2O is an obligatory intermediate in the reduction of NO2? to N2 in all of the three organisms studied.  相似文献   

19.
Ascorbate-reduced horse heart cytochrome c reduces photo-oxidized bacterial reaction centres with a second-order rate constant of (5–8) · 108 M?1 · s?1 at an ionic strength of 50 mM. In the absence of cytochrome c, the cytochrome c1 in the ubiquinol:cytochrome c oxidoreductase is oxidized relatively slowly (k = 3.3 · 105 M?1 · s?1). Ferrocytochrome c binds specifically to ascorbate-reduced reductase, with a Kd of 0.6 μM, and only the free cytochrome c molecules are involved in the rapid reduction of photo-oxidized reaction centres. The electron transfer between ferricytochrome c and ferrocytochrome c1 of the reductase is rapid, with a second-order rate constant of 2.1 · 108 M?1 · s?1 at an ionic strength of 50 mM. The rate of electron transfer from the Rieske iron-sulphur cluster to cytochrome c1 is even more rapid. The cytochrome b of the ubiquinol:cytochrome c oxidoreductase can be reduced by electrons from the reaction centres through two pathways: one is sensitive to antimycin and the other to myxothiazol. The amount of cytochrome b reduced in the absence of antimycin is dependent on the redox potential of the system, but in no case tested did it exceed 25% of the amount of photo-oxidized reaction centres.  相似文献   

20.
Evidence presented in this report suggests that the hydroxyl radical (OH.), which is generated from liver microsomes is an initiator of NADPH-dependent lipid peroxidation. The conclusions are based on the following observations: 1) hydroxyl radical production in liver microsomes as measured by esr spin-trapping correlates with the extent of NADPH induced microsomal lipid peroxidation as measured by malondialdehyde formation; 2) peroxidative degradation of arachidonic acid in a model OH · generating system, namely, the Fenton reaction takes place readily and is inhibited by thiourea, a potent OH · scavenger, indicating that the hydroxyl radical is capable of initiating lipid peroxidation; 3) trapping of the hydroxyl radical by the spin trap, 5,5-dimethyl-1-pyrroline-1-oxide prevents lipid peroxidation in liver microsomes during NADPH oxidation, and in the model system in the presence of linolenic acid. The possibility that cytochrome P-450 reductase is involved in NADPH-dependent lipid peroxidation is discussed. The optimal pH for the production of the hydroxyl radical in liver microsomes is 7.2. The generation of the hydroxyl radical is correlated with the amount of microsomal protein, possibly NADPH cytochrome P-450 reductase. A critical concentration of EDTA (5 × 10?5m) is required for maximal production of the hydroxyl radical in microsomal lipid peroxidation during NADPH oxidation. High concentrations of Fe2+-EDTA complex equimolar in iron and chelator do not inhibit the production of the hydroxyl radical. The production of the hydroxyl radical in liver microsomes is also promoted by high salt concentrations. Evidence is also presented that OH radical production in microsomes during induced lipid peroxidation occurs primarily via the classic Fenton reaction.  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