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.
《Free radical research》2013,47(9):1150-1156
Abstract

Oxidation of tyrosine moieties by radicals involved in lipid peroxidation is of current interest; while a rate constant has been reported for reaction of lipid peroxyl radicals with a tyrosine model, little is known about the reaction between tyrosine and alkoxyl radicals (also intermediates in the lipid peroxidation chain reaction). In this study, the reaction between a model alkoxyl radical, the tert-butoxyl radical and tyrosine was followed using steady-state and pulse radiolysis. Acetone, a product of the β-fragmentation of the tert-butoxyl radical, was measured; the yield was reduced by the presence of tyrosine in a concentration- and pH-dependent manner. From these data, a rate constant for the reaction between tert-butoxyl and tyrosine was estimated as 6?±?1 × 107 M?1 s?1 at pH 10. Tyrosine phenoxyl radicals were also monitored directly by kinetic spectrophotometry following generation of tert-butoxyl radicals by pulse radiolysis of solutions containing tyrosine. From the yield of tyrosyl radicals (measured before they decayed) as a function of tyrosine concentration, a rate constant for the reaction between tert-butoxyl and tyrosine was estimated as 7?±?3 × 107 M?1 s?1 at pH 10 (the reaction was not observable at pH 7). We conclude that reaction involves oxidation of tyrosine phenolate rather than undissociated phenol; since the pKa of phenolic hydroxyl dissociation in tyrosine is ~ 10.3, this infers a much lower rate constant, about 3 × 105 M?1 s?1, for the reaction between this alkoxyl radical and tyrosine at pH 7.4.  相似文献   

2.
The repair of tryptophan and tyrosine radicals in proteins by urate was studied by pulse radiolysis. In chymotrypsin, urate repairs tryptophan radicals efficiently with a rate constant of 2.7 × 10(8)M(-1)s(-1), ca. 14 times higher than the rate constant derived for N-acetyltryptophan amide, 1.9 × 10(7)M(-1)s(-1). In contrast, no repair of tryptophan radicals was observed in pepsin, which indicates a rate constant smaller than 6 × 10(7)M(-1)s(-1). Urate repairs tyrosine radicals in pepsin with a rate constant of 3 × 10(8)M(-1)s(-1)-ca. 12 times smaller than the rate constant reported for free tyrosine-but not in chymotrypsin, which implies an upper limit of 1 × 10(6)M(-1)s(-1) for the corresponding rate constant. Intra- and intermolecular electron transfer from tyrosine residues to tryptophan radicals is observed in both proteins, however, to different extents and with different rate constants. Urate inhibits electron transfer in chymotrypsin but not in pepsin. Our results suggest that urate repairs the first step on the long path to protein modification and prevents damage in vivo. It may prove to be a very important repair agent in tissue compartments where its concentration is higher than that of ascorbate. The product of such repair, the urate radical, can be reduced by ascorbate. Loss of ascorbate is then expected to be the net result, whereas urate is conserved.  相似文献   

3.
3-Nitrotyrosine has been reported as an important biomarker of oxidative stress that may play a role in a variety of diseases. In this work, transient UV-visible absorption spectra and kinetics observed during the reaction of the hydrated electron, e(aq)(-), with 3-nitrotyrosine and derivatives thereof were investigated. The absorption spectra show characteristics of aromatic nitro anion radicals. The absorptivity of radical anion product at 300 nm is estimated to be (1.0 ± 0.2) × 10(4) M(-1) cm(-1) at pH 7.3. The rate constants determined for the reaction of e(aq)(-) with 3-nitrotyrosine, N-acetyl-3-nitrotyrosine ethyl ester and glycylnitrotyrosylglycine at neutral pH (3.0 ± 0.3) × 10(10) M(-1) s(-1), (2.9 ± 0.2) × 10(10) M(-1) s(-1) and (1.9 ± 0.2) × 10(10) M(-1) s(-1), respectively, approach the diffusion-control limit and are almost two orders of magnitude higher than those for the reactions with tyrosine and tyrosine-containing peptides. The magnitude of the rate constants supports reaction of e(aq)(-) at the nitro group, and the product absorbance at 300 nm is consistent with formation of the nitro anion radical. The pH dependence of the second-order rate constant for e(aq)(-) decay (720 nm) in the presence of 3-nitrotyrosine shows a decrease with increasing pH, consistent with unfavorable electrostatic interactions. The pH dependence of the second-order rate constant for formation of radical anion (300 nm) product suggests that deprotonation of the amino group slows the rate, which indicates that deamination to form the 1-carboxy-2-(4-hydroxy-3-nitrophenyl)ethyl radical occurs. We conclude that the presence of the nitro group activates tyrosine and derivatives toward reaction with e(aq)(-) and can affect the redox chemistry of biomolecules exposed to oxidative stress.  相似文献   

4.
Nitrogen dioxide is a product of peroxynitrite homolysis and peroxidase-catalyzed oxidation of nitrite. It is of great importance in protein tyrosine nitration because most nitration pathways end with the addition of *NO2 to a one-electron-oxidized tyrosine. The rate constant of this radical addition reaction is high with free tyrosine-derived radicals. However, little is known of tyrosine radicals in proteins. In this paper, we have used *NO2 generated by gamma radiolysis to study the nitration of the R2 subunit of ribonucleotide reductase, which contains a long-lived tyrosyl radical on Tyr122. Most of the nitration occurred on Tyr122, but nonradical tyrosines were also modified. In addition, peptidic bonds close to nitrated Tyr122 could be broken. Nitration at Tyr122 was not observed with a radical-free metR2 protein. The estimated rate constant of the Tyr122 radical reaction with *NO2 was of 3 x 10(4) M(-1) s(-1), thus several orders of magnitude lower than that of a radical on free tyrosine. Nitration rate of other tyrosine residues in R2 was even lower, with an estimated value of 900 M(-1) s(-1). This study shows that protein environment can significantly reduce the reactivity of a tyrosyl radical. In ribonucleotide reductase, the catalytically active radical residue is very efficiently protected against nitrogen oxide attack and subsequent nitration.  相似文献   

5.
We have demonstrated that hypochlorite (HOCI/OCl-) and hypobromite (HOBr/OBr-) can react with tert-butyl hydroperoxide with close rate constants (k(HOCl) = 10,8 M(-1) x s(1); k(HOBr) = 8,9 M(-1) x (s(-1)). By means of the spin trap 4-pyridyl-1-oxide-N-tert-butyl nitron we have found that both reactions proceed through decomposition of tert-butyl hydroperoxide and generation of tert-butyl peroxyl (OOC(CH3)3) and tert-butoxyl (OC(CH3)3) radicals, the ratio of their the concentrations being dependent on the concentration of tert-butyl hydroperoxide. Thus, hypobromite, similar to hypochlorite, is a precursor of free radicals produced in the reaction with organic hydroperoxides. This reaction can be of great importance in the intensification of free radical processes, namely, in lipid peroxidation at the stage of chain branching.  相似文献   

6.
The reactivity of ebselen, 2-phenyl-1,2-benzisoselenazol-3(2H)one, and structurally related analogues was studied by pulse radiolysis. The rate constant for the reaction of ebselen with trichloromethylperoxyl radicals was determined to be 2.9 X 10(8) M-1 s-1, while its sulfur analogue, 2-phenyl-1,2-benzisothiazol-3(2H)one, was oxidized at much lower rates, k less than or equal to 10(7) M-1 s-1. Among several derivatives studied, the only other compound that exhibited a high rate constant was 2-(methylseleno)-benzoic acid-N-phenylamide. Oxidation of ebselen by other halogenated peroxyl radicals was also carried out and revealed a direct relationship between rate constant and the degree of halogenation of the oxidant. The transient radicals generated during oxidation of ebselen and the analogues were characterized by optical absorption and conductivity measurements and were attributed to one-electron-oxidized radical cations. The oxidation potentials were determined by cyclic voltammetry. Comparative evaluation of the in vitro behavior during microsomal lipid peroxidation revealed ebselen to be the most potent antioxidant of the compounds investigated, 2-(Methylseleno)-benzoic acid-N-phenylamide, despite its high rate constant for oxidation by halogenated peroxyl radicals, was found to be a poor antioxidant. The rate constant of oxidation of ebselen by trichloromethylperoxyl radicals is comparable to that of alpha-tocopherol under similar conditions, underscoring the potential pharmacological interest of ebselen as an antioxidant.  相似文献   

7.
The rate constant for the reduction of the tyrosyl radical with selenocysteine has been measured to investigate whether selenocysteine is capable of repair of protein radicals. Tyrosyl radicals, both free in solution and in insulin, were generated by means of pulse radiolysis and laser flash photolysis in aqueous solution. The rate constant for the reaction of free N-acetyl-tyrosyl-amine radicals with selenocysteine is (8 +/- 2) x 10 (8) M (-1) s (-1), and that for tyrosyl radicals in insulin is (1.6 +/- 0.4) x 10 (8) M (-1) s (-1). The rate constant for the reaction of selenoglutathione with the N-acetyl-tyrosyl-amine radical is (5 +/- 2) x 10 (8) M (-1) s (-1). In contrast, cysteine and glutathione react more slowly than their selenium analogues with the tyrosyl radical: the reactions of N-acetyl-tyrosyl-amine radicals with cysteine and glutathione are 3 and 5 orders of magnitude slower, respectively, than those with selenocysteine and selenoglutathione, while those of tyrosyl radicals in insulin are 3 and 2 orders of magnitude slower, respectively.  相似文献   

8.
Hydrogen sulfide (H(2)S) is an endogenously generated gas that can also be administered exogenously. It modulates physiological functions and has reported cytoprotective effects. To evaluate a possible antioxidant role, we investigated the reactivity of hydrogen sulfide with several one- and two-electron oxidants. The rate constant of the direct reaction with peroxynitrite was (4.8±1.4)×10(3)M(-1) s(-1) (pH 7.4, 37°C). At low hydrogen sulfide concentrations, oxidation by peroxynitrite led to oxygen consumption, consistent with a one-electron oxidation that initiated a radical chain reaction. Accordingly, pulse radiolysis studies indicated that hydrogen sulfide reacted with nitrogen dioxide at (3.0±0.3)×10(6)M(-1) s(-1) at pH 6 and (1.2±0.1)×10(7)M(-1) s(-1) at pH 7.5 (25°C). The reactions of hydrogen sulfide with hydrogen peroxide, hypochlorite, and taurine chloramine had rate constants of 0.73±0.03, (8±3)×10(7), and 303±27M(-1) s(-1), respectively (pH 7.4, 37°C). The reactivity of hydrogen sulfide was compared to that of low-molecular-weight thiols such as cysteine and glutathione. Considering the low tissue concentrations of endogenous hydrogen sulfide, direct reactions with oxidants probably cannot completely account for its protective effects.  相似文献   

9.
The effects of t-butylhydroperoxide (tBHP), its alkoxyl radical (tBuO.) and its peroxyl radical (tBuOO.) in model systems and on red blood cells were studied. Glyceraldehyde-3-phosphate dehydrogenase was strongly inhibited by tBHP via a direct reaction of the hydroperoxide with an essential sulfhydryl group in the enzyme molecule. Several other enzymes were unaffected by tBHP. Alcohol dehydrogenase was strongly inhibited by tBuO. but was much less sensitive to tBuOO.. Lysozyme, lactate dehydrogenase and trypsin, on the other hand, were very sensitive to the peroxyl and not, or much less, to the alkoxyl radical, whereas acetylcholinesterase was very sensitive to both radicals. tBuOO. caused covalent binding of tryptophan, tyrosine, histidine and methionine to serum albumin. The corresponding alkoxyl radical was ineffective in this respect. Conversely, tBuO. caused peroxidation of linolenic acid, whereas tBuOO. did not. Incubation of human erythrocytes with tBHP caused lipid peroxidation and K+ leakage. Both effects were caused by tBHP-derived radicals generated in a reaction of the hydroperoxide with hemoglobin. With radical scavengers it was possible to dissociate tBHP-induced lipid peroxidation and K+ leakage, demonstrating that these two processes are not causally related. Experimental results indicate that tBuO. causes lipid peroxidation, whereas tBuOO. is responsible for K+ leakage.  相似文献   

10.
Generation of superoxide radicals (0.01-0.1 microm s(-1)) by radiolysis of aqueous solutions containing S-nitrosoglutathione (45-160 microm, pH 3.8-7.3) resulted in loss of this solute at rates varying with solute concentration, radical generation rate, and pH. The results were quantitatively consistent with the loss being attributed to competition between reaction of superoxide with S-nitrosoglutathione (rate constant 300 +/- 100 m(-1) s(-1)) and the pH-dependent disproportionation of superoxide/hydroperoxyl. This rate constant is much lower than previous estimates and seven orders of magnitude lower than the rate constants between superoxide and superoxide dismutase or superoxide and nitric oxide. This indicates that interaction between superoxide and S-nitrosoglutathione is unlikely to be biologically important, contrary to previous suggestions that reaction could serve to prevent the rapid reaction between superoxide and nitric oxide. Reductive homolysis of S-nitrosoglutathione by the carbon dioxide radical anion, a model for biological reductants such as disulfide radical anions, occurred with a rate constant of 7.4 x 10(8) m(-1) s(-1) and produced nitric oxide stoichiometrically. Thiyl radicals were not produced, indicating the alternative homolysis route to generate nitroxyl did not occur.  相似文献   

11.
The reaction of the trioxidocarbonate(*1-) radical (CO (3) (*-) , "carbonate radical anion") with cytochrome c was studied by pulse radiolysis at alkaline pH and room temperature. With iron(III) cytochrome c, CO (3) (*-) reacts with the protein moiety with rate constants of (5.1 +/- 0.6) x 10(7) M(-1) s(-1) (pH 8.4, I approximately 0.27 M) and (1.0 +/- 0.2) x 10(8) M(-1) s(-1) (pH 10, I = 0.5 M). The absorption spectrum of the haem moiety was not changed, thus, amino acid radicals produced on the protein do not reduce the haem. The pH-dependent difference in rate constants may be attributed to differences in ionization states of amino acids and to the change in the conformation of the protein. With iron(II) cytochrome c, CO (3) (*-) oxidizes the haem quantitatively, presumably via electrostatic guidance of the radical to the solvent-accessible haem edge, with a different pH dependence: at pH 8.4, the rate constant is (1.1 +/- 0.1) x 10(9) M(-1) s(-1) and, at pH 10, (7.6 +/- 0.6) x 10(8) M(-1) s(-1). We propose that CO (3) (*-) oxidizes the iron center directly, and that the lower rate observed at pH 10 is due to the different charge distribution of iron(II) cytochrome c.  相似文献   

12.
Charge transfer from tyrosine to tryptophan radicals in bovine milk casein, as observed using pulse radiolysis technique, is reported. The reactions of casein with hydroxyl, azide, Br(2)(*-) and CCl(3)O(2)(*) radicals have also been studied. Radical transformation was found to take place at a rate of 1.5 x10(4) s(-1). The effect of pH, oxidising radical and the proximity of tyrosine and tryptophan on this radical transformation, as well as repair of the casein radical by ascorbate, have also been studied.  相似文献   

13.
New water-soluble analogues of 1,3,7-trimethyluric acid with N-1 methyl replaced by various groups were prepared and evaluated for their ability to scavenge hydroxyl radicals as well as their protective potential against lipid peroxidation in erythrocyte membranes. The deoxyribose degradation method indicates that all the analogues tested effectively scavenge hydroxyl radicals and some of them show better activity than uric acid and methyluric acids. These effects are shown to be concentration dependent and are more potent at low concentrations (10-50 microM). Among the analogues tested, 1-butenyl-, 1-propargyl- and 1-benzyl-3,7-dimethyluric acids show high hydroxyl radical scavenging property with a reaction rate constant (Ks) of 3.2-6.7 x 10(10) M(-1) S(-1), 2.3-3.7 x 10(10) M(-1) S(-1) and 2.4-3.7 x 10(10) M(-1) S(-1), respectively. The effectiveness of these analogues as hydroxyl radical scavengers appears to be better than mannitol (Ks, 1.9-2.5 x 10(9) M(-1) S(-1)). With the exception of 1-pentyl- and 1-(2'-oxopropyl)-3,7-dimethyluric acids, all other analogues tested are effective inhibitors of tert-butylhydroperoxide-induced lipid peroxidation in human erythrocyte membranes. All the analogues tested are susceptible to peroxidation in the presence of hemoprotein and hydrogen peroxide. The present study has pointed out that it is possible to significantly enhance the antioxidant property of 1,3,7-trimethyluric acid by structural modification at N-1 position. Such compounds may be useful as antioxidants in vivo.  相似文献   

14.
By application of pulse radiolysis it was demonstrated that nitrogen dioxide (NO2.) oxidizes Gly-Tyr in aqueous solution with a strongly pH-dependent rate constant (k6 = 3.2 X 10(5) M-1 S-1 at pH 7.5 and k6 = 2.0 X 10(7) M-1 S-1 at pH 11.3), primarily generating phenoxyl radicals. The phenoxyl can react further with NO2. (k7 approximately 3 X 10(9) M-1 S-1) to form nitrotyrosine, which is the predominant final product in neutral solution and at low tyrosyl concentrations under gamma-radiolysis conditions. Tyrosine nitration is less efficient in acidic solution, due to the natural disproportionation of NO2., and in alkaline solutions and at high tyrosyl concentrations due to enhanced tyrosyl dimerization. Selective tyrosine nitration by interaction of NO2. with proteins (at pH 7 to 9) was demonstrated in the case of histone, lysozyme, ribonuclease A, and subtilisin Carlsberg. Nitrotyrosine developed slowly also under incubation of Gly-Tyr with nitrite at pH 4 to 5, where NO2. is formed by acid decomposition of HONO. It is recalled in this context that NO2.-induced oxidations, by regenerating NO2-, can propagate NO2./NO2- redox cycling under acidic conditions. Even faster than with tyrosine is the NO2.-induced oxidation of cysteine-thiolate (k9 = 2.4 X 10(8) M-1 S-1 at pH 9.2), involving the transient formation of cystinyl radical anions. The interaction of NO2. with Gly-Trp was comparably slow (k approximately 10(6) M-1 S-1), and no reaction was detectable by pulse radiolysis with Met-Gly and (Cys-Gly)2, or with DNA. Slow reactions of NO2. were observed with arachidonic acid (k approximately 10(6) M-1 S-1 at pH 9.0) and with linoleate (k approximately 2 X 10(5) M-1 S-1 at pH 9.4), indicating that NO2. is capable of initiating lipid peroxidation even in an aqueous environment. NO2.-Induced tyrosine nitration, using 50 microM Gly-Tyr at pH 8.2, was hardly inhibited, however, in the presence of 1 mM linoleate, and was not affected at all in the presence of 5 mM dimethylamine (a nitrosamine precursor). It is concluded that protein modifications, and particularly phenol and thiol oxidation, may be an important mechanism, as well as initiation of lipid peroxidation, of action of NO2. in biological systems.  相似文献   

15.
The radiation-induced reactions of a water-soluble coumarin derivative, coumarin-3-carboxyl acid (C3CA), have been investigated in aqueous solutions by pulse radiolysis with a 35 MeV electron beam, final product analysis following (60)Co γ-irradiations and deterministic model simulations. Pulse radiolysis revealed that C3CA reacted with both hydroxyl radicals ((?)OH) and hydrated electrons (e(-) (aq)) with near diffusion-controlled rate constants of 6.8 × 10(9) and 2.1 × 10(10) M(-1) s(-1), respectively. The reactivity of C3CA towards O(2)(? -) was not confirmed by pulse radiolysis. Production of the fluorescent molecule, 7-hydroxy-coumarin-3-carboxylic acid (7OH-C3CA), was confirmed by final product analysis with a fluorescence spectrometer coupled to a high performance liquid chromatography (HPLC) system. Production yields of 7OH-C3CA following (60)Co γ-irradiations depended on the irradiation conditions and ranged from 0.025 to 0.18 (100 eV) (-1). Yield varied with saturating gas, additive and C3CA concentration, implying the presence of at least two pathways capable of providing 7OH-C3CA as a stable product following the scavenging reaction of C3CA with (?)OH, including a peroxidation/elimination sequence and a disproportionation pathway. A reaction mechanism for the two pathways was proposed and incorporated into a deterministic simulation, showing that the mechanism can explain experimentally measured 7OH-C3CA yields with a constant conversion factor of 4.7% from (?)OH scavenging to 7OH-C3CA production, unless t-BuOH was added.  相似文献   

16.
The azide, dibromide and dichloride radicals oxidize one or more tryptophan side chains in hen egg-white lysozyme. The indolyl radical produced in this second-order 1-electron oxidation subsequently oxidizes a tyrosine side chain to the phenoxy radical in an intramolecular reaction with a rate constant of 130 +/- 10 s-1 at pH 7, 25 degrees C. The final indolyl and phenoxy equilibrium mixture then decays with a t1/2 approximately 2 s. The faster intramolecular reaction exhibits a pH dependence; on decreasing the pH from 9 the first-order rate constant increases to a maximum near pH 5.4 and then declines as the pH is lowered further. In contrast, the first-order rate constant for the intramolecular electron transfer between the tyrosine and tryptophan of the peptide trpH-pro-tyrOH remains unchanged between approx. pH 11 and 6.5 and then increases as the pH is lowered further. This difference in the observed pH dependence suggests that changes in structure or ionization state influence the protein electron transfer rate. We also discuss the radiation inactivation of lysozyme in light of these observations.  相似文献   

17.
Baicalein (5, 6, 7-trihydroxy-2-phenyl-4H-1-benzopyran-4-one), a naturally occurring flavone present in some of the medicinal plants is known for its potential therapeutic effects, such as cardioprotective, anticancer and anti-inflammatory properties. However, detailed role and mechanisms behind its protective properties against different generators for oxidative stress have not been examined. In the present study, we investigated the possible protective ability of baicalein against the membrane damage caused by reactive oxygen species (ROS) and reactive nitrogen species (RNS) and the mechanisms involved using pulse radiolysis technique. Baicalein offered efficient protection even at a concentration of 10 microM towards membrane damage caused by lipid peroxidation induced by the gamma-radiation, peroxyl radicals, ascorbate-Fe2+ and peroxynitrite in rat liver mitochondria and heart homogenate. To elucidate its reaction mechanisms with biologically relevant radicals, transient absorption spectroscopy employing pulse radiolysis technique was used. Baicalein showed fairly high rate constants (3.7 x 10(9), 1.3 x 10(9) and 8.0 x 10(8) dm3 mol(-1) s(-1) for hydroxyl, azidyl and alkylchloroperoxyl radicals, respectively), suggesting that baicalein can act as an effective scavenger of these radicals. In each case, the phenoxyl radical of baicalein was generated. Thus, it was evident that the phenolic moiety of baicalein was responsible for the free radical scavenging process. Baicalein also reacts with linoleic acid peroxyl radical (LOO*), indicating its ability to act as a chain breaking antioxidant. Peroxynitrite-mediated radicals were shown to be reactive towards baicalein and the bimolecular rate constants were 2.5 x 10(7) and 3 x 10(8) dm3 mol(-1) s(-1) for *NO2 and CO3*(-) radicals, respectively. In conclusion, our results revealed the potential of baicalein in protecting mitochondrial membrane against oxidative damage induced by the four different agents. We propose that the protective effect is mediated via scavenging of primary and secondary radicals generated during oxidative stress.  相似文献   

18.
Radical scavenging properties of genistein   总被引:20,自引:0,他引:20  
The reactivity of genistein toward reactive radical species has been investigated by means of pulse radiolysis. The values of rate constants, respectively 2.3 x 10(10) M(-1)s(-1) and 1.3 x 10(10) M(-1)s(-1) for the reaction with hydroxyl radical at pH 8.3 and 3.0, are close to diffusion limit indicating that genistein is a potent hydroxyl radical scavenger. The reactivity of genistein towards one-electron oxidants has also been investigated. The rate constants k = 4.6 x 10(9) M(-1)s(-1) (pH 8.3) and 6.7 x 10(8) M(-1)s(-1) (pH 7.6) have been determined for the reaction of genistein with *N3 and Br2*- radicals, respectively. For both oxidants the rate constants at pH 3 does not exceed 10(8) M(-1)s(-1). The differences in reactivity of genistein towards the oxidants at different acidity of the solution have been assumed to arise from the acid-base equilibria of genistein. The dissociation constants for genistein (pKa: 7.2, 10.0, and 13.1) have been evaluated spectroscopically. The influence of acid-base equilibria on bond dissociation energy and ionization potential for genistein has also been investigated by means of DFT calculations. It has been concluded on the basis of these calculations that monoanionic form of genistein existing at physiological pH is more powerful radical scavenger than the neutral molecule.  相似文献   

19.
The excited triplet state of 1-nitronaphthalene ((3)1NN*) reacts with OH(-) with a second-order reaction rate constant of (1.66 ± 0.08)×10(7) M(-1) s(-1) (μ±σ). The reaction yields the ˙OH radical and the radical anion 1NN(-)˙. In aerated solution, the radical 1NN(-)˙ would react with O(2) to finally produce H(2)O(2) upon hydroperoxide/superoxide disproportionation. The photolysis of H(2)O(2) is another potential source of ˙OH, but such a pathway would be a minor one in circumneutral (pH 6.5) or in basic solution ([OH(-)] = 0.3-0.5 M). The oxidation of H(2)O by (3)1NN*, with rate constant 3.8 ± 0.3 M(-1) s(-1), could be the main ˙OH source at pH 6.5.  相似文献   

20.
The iron chelating agent desferrioxamine inhibits peroxynitrite-mediated oxidations and attenuates nitric oxide and oxygen radical-dependent oxidative damage both in vitro and in vivo. The mechanism of protection is independent of iron chelation and has remained elusive over the past decade. Herein, stopped-flow studies revealed that desferrioxamine does not react directly with peroxynitrite. However, addition of peroxynitrite to desferrioxamine in both the absence and the presence of physiological concentrations of CO2 and under excess nitrite led to the formation of a one-electron oxidation product, the desferrioxamine nitroxide radical, consistent with desferrioxamine reacting with the peroxynitrite-derived species carbonate (CO3*-) and nitrogen dioxide (*NO2) radicals. Desferrioxamine inhibited peroxynitrite-dependent free radical-mediated processes, including tyrosine dimerization and nitration, oxyhemoglobin oxidation in the presence of CO2, and peroxynitrite plus carbonate-dependent chemiluminescence. The direct two-electron oxidation of glutathione by peroxynitrite was unaffected by desferrioxamine. The reactions of desferrioxamine with CO3*- and *NO2 were unambiguously confirmed by pulse radiolysis studies, which yielded second-order rate constants of 1.7 x 10(9) and 7.6 x 10(6) M(-1) s(-1), respectively. Desferrioxamine also reacts with tyrosyl radicals with k = 6.3 x 10(6) M(-1) s(-1). However, radical/radical combination reactions between tyrosyl radicals or of tyrosyl radical with *NO2 outcompete the reaction with desferrioxamine and computer-assisted simulations indicate that the inhibition of tyrosine oxidation can be fully explained by scavenging of the peroxynitrite-derived radicals. The results shown herein provide an alternative mechanism to account for some of the biochemical and pharmacological actions of desferrioxamine via reactions with CO3*- and *NO2 radicals.  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