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 453 毫秒
1.
Nitric-oxide synthases (NOS) are catalytically self-sufficient flavo-heme enzymes that generate NO from arginine (Arg) and display a novel utilization of their tetrahydrobiopterin (H(4)B) cofactor. During Arg hydroxylation, H(4)B acts as a one-electron donor and is then presumed to redox cycle (i.e. be reduced back to H(4)B) within NOS before further catalysis can proceed. Whereas H(4)B radical formation is well characterized, the subsequent presumed radical reduction has not been demonstrated, and its potential mechanisms are unknown. We investigated radical reduction during a single turnover Arg hydroxylation reaction catalyzed by neuronal NOS to document the process, determine its kinetics, and test for involvement of the NOS flavoprotein domain. We utilized a freeze-quench instrument, the biopterin analog 5-methyl-H(4)B, and a method that could separately quantify the flavin and pterin radicals that formed in NOS during the reaction. Our results establish that the NOS flavoprotein domain catalyzes reduction of the biopterin radical following Arg hydroxylation. The reduction is calmodulin-dependent and occurs at a rate that is similar to heme reduction and fast enough to explain H(4)B redox cycling in NOS. These results, in light of existing NOS crystal structures, suggest a "through-heme" mechanism may operate for H(4)B radical reduction in NOS.  相似文献   

2.
1-Methyl-4-phenyl-2,3-dihydropyridinium (MPDP+), a metabolic product of the nigrostriatal toxin 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine (MPTP), has been shown to generate superoxide radicals during its autoxidation process. The generation of superoxide radicals was detected as a 5,5-dimethyl-1-pyrroline-N-oxide (DMPO).O2- spin adduct by spin trapping in combination with EPR techniques. The rate of formation of spin adduct was dependent not only on the concentrations of MPDP+ and oxygen but also on the pH of the system. Superoxide dismutase inhibited the spin adduct formation in a dose-dependent manner. The ability of DMPO to trap superoxide radicals, generated during the autoxidation of MPDP+, and of superoxide dismutase to effectively compete with this reaction for the available O2-, has been used as a convenient competition reaction to quantitatively determine various kinetic parameters. Thus, using this technique the rate constant for scavenging of superoxide radical by superoxide dismutase was found to be 7.56 x 10(9) M-1 s-1. The maximum rate of superoxide generation at a fixed spin trap concentration using different amounts of MPDP+ was found to be 4.48 x 10(-10) M s-1. The rate constant (K1) for MPDP+ making superoxide radical was found to be 3.97 x 10(-6) s-1. The secondary order rate constant (KDMPO) for DMPO-trapping superoxide radicals was found to be 10.2 M-1 s-1. The lifetime of superoxide radical at pH 10.0 was calculated to be 1.25 s. These values are in close agreement to the published values obtained using different experimental techniques. These results indicate that superoxide radicals are produced during spontaneous oxidation of MPDP+ and that EPR spin trapping can be used to determine the rate constants and lifetime of free radicals generated in aqueous solutions. It appears likely that the nigrostriatal toxicity of MPTP/MPDP+ leading to Parkinson's disease may largely be due to the reactivity of these radicals.  相似文献   

3.
Endothelial nitric-oxide synthase (eNOS) plays important roles in vascular physiology and homeostasis. Whether eNOS catalyzes nitric oxide biosynthesis or the synthesis of reactive oxygen species such as superoxide, hydrogen peroxide, and peroxynitrite is dictated by the bioavailability of tetrahydrobiopterin (BH(4)) and L-arginine during eNOS catalysis. The effect of BH(4) and L-arginine on oxygen-induced radical intermediates has been investigated by single turnover rapid-freeze quench and EPR spectroscopy using the isolated eNOS oxygenase domain (eNOS(ox)). Three distinct radical intermediates corresponding to >50% of the heme were observed during the reaction between ferrous eNOS(ox) and oxygen. BH(4)-free eNOS(ox) produced the superoxide radical very efficiently in the absence of L-arginine. L-Arginine decreased the formation rate of superoxide by an order of magnitude but not its final level or EPR line shape. For BH(4)-containing eNOS(ox), only a stoichiometric amount of BH(4) radical was produced in the presence of L-arginine, but in its absence a new radical was obtained. This new radical could be either a peroxyl radical of BH(4) or an amino acid radical was in the vicinity of the heme. Formation of this new radical is very rapid, >150 s(-1), and it was subsequently converted to a BH(4) radical. The trapping of the superoxide radical by cytochrome c in the reaction of BH(4)(-) eNOS(ox) exhibited a limiting rate of approximately 15 s(-1), the time for the superoxide radical to leave the heme pocket and reach the protein surface; this reveals a general problem of the regular spin-trapping method in determining radical formation kinetics. Cytochrome c failed to trap the new radical species. Together with other EPR characteristics, our data strongly support the conclusion that this new radical is not a superoxide radical or a mixture of superoxide and biopterin radicals. Our study points out distinct roles of BH(4) and L-arginine in regulating eNOS radical intermediates. BH(4) prevented superoxide formation by chemical conversions of the Fe(II)O(2) intermediate, and l-arginine delayed superoxide formation by electronic interaction with the heme-bound oxygen.  相似文献   

4.
Berka V  Wang LH  Tsai AL 《Biochemistry》2008,47(1):405-420
Fully coupled nitric oxide synthase (NOS) catalyzes formation of nitric oxide (NO), l-citrulline, NADP+, and water from l-arginine, NADPH, and oxygen. Uncoupled or partially coupled NOS catalyzes the synthesis of reactive oxygen species such as superoxide, hydrogen peroxide, and peroxynitrite, depending on the availability of cofactor tetrahydrobiopterin (BH4) and l-arginine during catalysis. We identified three distinct oxygen-induced radical intermediates in the ferrous endothelial NOS oxygenase domain (eNOSox) with or without BH4 and/or l-arginine [Berka, V., Wu, G., Yeh, H. C., Palmer, G., and Tsai, A.-L. (2004) J. Biol. Chem. 279, 32243-32251]. The effects of BH4 and l-arginine on the oxygen-induced radical intermediates in the isolated neuronal NOS oxygenase domain (nNOSox) have been similarly investigated by single-turnover stopped-flow and rapid-freeze quench EPR kinetic measurements in the presence or absence of dithiothreitol (DTT). Like for eNOSox, we found different radical intermediates in the reaction of ferrous nNOSox with oxygen. (1) nNOSox (without BH4 or l-Arg) produces superoxide in the presence or absence of DTT. (2) nNOSox (with BH4 and l-Arg) yields a typical BH4 radical in a manner independent of DTT. (3) nNOSox (with BH4 and without l-Arg) yields a new radical. Without DTT, EPR showed a mixture of superoxide and biopterin radicals. With DTT, a new approximately 75 G wide radical EPR was observed, different from the radical formed by eNOSox. (4) The presence of only l-arginine in nNOSox (without BH4 but with l-Arg) caused conversion of approximately 70% of superoxide radical to a novel radical, explaining how l-arginine decreases the level of superoxide production in nNOSox (without BH4 but with l-Arg). The regulatory role of l-arginine in nNOS is thus very different from that in eNOS where substrate was only to decrease the rate of formation of superoxide but not the total amount of radical. The role of DTT is also different. DTT prevents oxidation of BH4 in both isoforms, but in nNOS, DTT also inhibits oxidation of two key cysteines in nNOSox to prevent the loss of substrate binding. This new role of thiol found only for nNOS may be significant in neurodegenerative diseases.  相似文献   

5.
A purely chemical system for NAD(P)H oxidation to biologically active NAD(P)+ has been developed and characterized. Suitable amounts of EDTA, manganous ions and mercaptoethanol, combined at physiological pH, induce nucleotide oxidation through a chain length also involving molecular oxygen, which eventually undergoes quantitative reduction to hydrogen peroxide. Mn2+ is specifically required for activity, while both EDTA and mercaptoethanol can be replaced by analogs. Optimal molar ratios of chelator/metal ion (2:1) yield an active coordination compound which catalyzes thiol autoxidation to thiyl radical. The latter is further oxidized to disulfide by molecular oxygen whose one-electron reduction generates superoxide radical. Superoxide dismutase (SOD) inhibits both thiol oxidation and oxygen consumption as well as oxidation of NAD(P)H if present in the mixture. A tentative scheme for the chain length occurring in the system is proposed according to stoichiometry of reactions involved. Two steps appear of special importance in nucleotide oxidation: (a) the supposed transient formation of NAD(P). from the reaction between NAD(P)H and thiyl radicals; (b) the oxidation of the reduced complex by superoxide to keep thiol oxidation cycling.  相似文献   

6.
We have studied the reaction of reduced nitric-oxide synthase (NOS) with molecular oxygen at -30 degrees C. In the first reaction cycle (from L-Arg to hydroxy-L-Arg), an oxygen adduct complex formed rapidly. Experiments in the absence of the reductase domain demonstrated that this complex was then further reduced by one electron stemming from the cofactor tetrahydrobiopterin (BH4). Spectral evidence suggested an iron(IV) porphyrin pi-cation radical as an intermediate. The nature of the oxidized BH4 was identified by EPR as a BH3* radical. Within the second cycle (from hydroxy-L-Arg to citrulline and NO), an iron(III)-NO complex could be identified clearly by its spectral characteristics. The strict requirement of BH4 for its formation suggests that BH4 plays a redox role, although transient, also in the second reaction cycle.  相似文献   

7.
1-Methyl-4-phenyl-2,3-dihydropyridinium perchlorate (MPDP+), an intermediate in the metabolism of the neurotoxin 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine, was found to generate superoxide radicals during its autoxidation process. The generation of superoxide radicals was detected by their ability to reduce ferricytochrome c. Superoxide dismutase inhibited this reduction in a dose-dependent manner. The rate of reduction of ferricytochrome c was dependent not only on the concentration of MPDP+ but also on the pH of the system. Thus, the rate of autoxidation of MPDP+ and the sensitivity of this autoxidation to superoxide dismutase-inhibitable ferricytochrome c reduction were both augmented, as the pH was raised from 7.0 to 10.5. The rate constant (Kc) for the reaction of superoxide radical with ferricytochrome c to form ferricytochrome c was found to be 3.48 x 10(5) M-1 s-1. The rate constant (KMPDP+) for the reaction of MPDP+ with ferricytochrome3+ c was found to be only 4.86 M-1 s-1. These results, in conjunction with complexities in the kinetics, lead to the proposal that autoxidation of MPDP+ proceeds by at least two distinct pathways, one of which involves the production of superoxide radicals and hence is inhibitable by superoxide dismutase. It is possible that the free radicals so generated could induce oxidative injury which may be central to the MPTP/MPDP(+)-induced neuropathy.  相似文献   

8.
A mechanism for the production of hydroxyl radical (*OH) during the oxidation of hydroquinones by laccase, the ligninolytic enzyme most widely distributed among white-rot fungi, has been demonstrated. Production of Fenton reagent (H2O2 and ferrous ion), leading to *OH formation, was found in reaction mixtures containing Pleurotus eryngii laccase, lignin-derived hydroquinones, and chelated ferric ion. The semiquinones produced by laccase reduced both ferric to ferrous ion and oxygen to superoxide anion radical (O2*-). Dismutation of the latter provided the H2O2 for *OH generation. Although O2*- could also contribute to ferric ion reduction, semiquinone radicals were the main agents accomplishing the reaction. Due to the low extent of semiquinone autoxidation, H2O2 was the limiting reagent in Fenton reaction. The addition of aryl alcohol oxidase and 4-methoxybenzyl alcohol (the natural H2O2-producing system of P. eryngii) to the laccase reaction greatly increased *OH generation, demonstrating the synergistic action of both enzymes in the process.  相似文献   

9.
Leukotriene B4, C4, D4 and E4 inactivation by hydroxyl radicals   总被引:1,自引:0,他引:1  
Leukotriene B4 chemotactic activity and leukotriene C4, D4 and E4 slow reacting substance activity were rapidly decreased by hydroxyl radicals generated by two different iron-supplemented acetaldehyde-xanthine oxidase systems. At low Fe2+, leukotriene inactivation was inhibited by catalase, superoxide dismutase, mannitol and ethanol, suggesting involvement of hydroxyl radicals generated by the iron-catalyzed interaction of superoxide and H2O2 (Haber-Weiss reaction). Leukotriene inactivation increased at high Fe2+ concentrations, but was no longer inhibitable by superoxide dismutase, suggesting that inactivation resulted from a direct interaction between H2O2 and Fe2+ to form hydroxyl radicals (Fenton reaction). The inactivation of leukotrienes by hydroxyl radicals suggests that oxygen metabolites generated by phagocytes may play a role in modulating leukotriene activity.  相似文献   

10.
During autoxidation of 1,4-hydroquinone (H2Q, less than 1 mM) at pH 7.4 and 37 degrees C, stoichiometric amounts of 1,4-benzoquinone (Q) and hydrogen peroxide were formed during the initial reaction. The reaction kinetics showed a significant induction period which was abolished by minute amounts of Q. Hydrogen peroxide and catalase were without effect on the autoxidation process. Transition metals apparently were not involved, since chelators like EDTA, DETAPAC, and desferrioxamine or FeSO4 had no influence on the autoxidation kinetics. Superoxide dismutase (SOD) did not abolish the induction period but dramatically enhanced the autoxidation rate by more than two orders of magnitude. The stimulatory effect was first-order in SOD concentration but showed saturation kinetics. The dependence of Q and hydrogen peroxide formation rates on H2Q concentration shows a biphasic behaviour: dependence on the square at low H2Q, but on the square root at high H2Q concentration. As revealed by calculatory simulations the results can be adequately described by the known reaction rate constants. The reaction starts with the comproportionation of H2Q and Q to yield two semiquinone molecules which autoxidize to give two superoxide radicals and two molecules of Q which enter into a new cycle of comproportionation. Because of unfavourable equilibria the autocatalytic reaction soon comes to steady state, and the further reaction is governed by the rate of superoxide removal. At excess SOD, the comproportionation reaction is rate-limiting, thus explaining the saturation effects of SOD. The experiments do not allow a decision between the two functions of SOD; the conventional action as a superoxide:superoxide oxidoreductase or as a semiquinone:superoxide oxidoreductase. In the latter reaction SOD is thought to be reduced by semiquinone with Q formation. In the second step the reduced enzyme would be re-oxidized by a superoxide radical which is formed during autoxidation of the second semiquinone molecule generated in the comproportionation reaction. From thermodynamic considerations, the latter function of SOD appears to be plausible.  相似文献   

11.
A Rhodococcus sp. BPG-8 produces 1,2,4-benzenetriol during the transformation of resorcinol by phloroglucinol induced cell-free extract. The oxidation of 1,2,4-benzenetriol to 2-hydroxy-1,4-benzoquinone produces superoxide radicals that may have potential deleterious effects on cellular integrity. It has been shown that both superoxide dismutase (SOD) and catalase retard the autoxidation of 1,2,4-benzenetriol to 2-hydroxy-1,4-benzoquinone. Termination of the free radical chain reaction between superoxide radical and 1,2,4-benzenetriol seems to prevent this autoxidation. A NAD(P)H-dependent reductase appears to convert the 2-hydroxy-1,4-benzoquinone back to 1,2,4-benzenetriol. Both of these mechanisms appear to stabilize 1,2,4-benzenetriol so that it may be cleaved by meta cleavage enzymes. The enzymes responsible for the stabilization of 1,2,4-benzenetriol appear not to be inducible.  相似文献   

12.
The ESR spectra of the free radicals formed by the autoxidation of serotonin, 5-hydroxyindole, and 5-hydroxytryptophan in 1 N NaOH are presented. The analysis of the hyperfine splitting constants in H2O and D2O characterize these free radicals as semiquinone-imines, the one-electron oxidation product of the corresponding indole. At alkaline pH, autoxidation of these compounds ultimately leads to solid precipitate and unresolved ESR spectra characteristic of polymeric material. The reduction of cytochrome c at pH 7.4 by a wide variety of indoles correlates with the amplitude of the ESR signal in 1 N NaOH, as do other processes thought to be related to 5-hydroxyindole free radical formation. Relative to the rate of cytochrome c reduction, neither serotonin nor the serotonin free radical appears to react with oxygen to form superoxide. In the presence of NAD(P)H, the serotonin radical most probably oxidizes NAD(P)H to form the NAD(P). radical. The NAD(P). radical then reacts with oxygen to form superoxide, which ultimately reduces cytochrome c.  相似文献   

13.
The metabolic disorder, alkaptonuria, is distinguished by elevated serum levels of 2,5-dihydroxyphenylacetic acid (homogentisic acid), pigmentation of cartilage and connective tissue and, ultimately, the development of inflammatory arthritis. Oxygen radical generation during homogentisic acid autoxidation was characterized in vitro to assess the likelihood that oxygen radicals act as molecular agents of alkaptonuric arthritis in vivo. For homogentisic acid autoxidized at physiological pH and above, yielding superoxide (O2-)2 and hydrogen peroxide (H2O2), the homogentisic acid autoxidation rate was oxygen dependent, proportional to homogentisic acid concentration, temperature dependent and pH dependent. Formation of the oxidized product, benzoquinoneacetic acid was inhibited by the reducing agents, NADH, reduced glutathione, and ascorbic acid and accelerated by SOD and manganese-pyrophosphate. Manganese stimulated autoxidation was suppressed by diethylenetriaminepentaacetic acid (DTPA). Homogentisic acid autoxidation stimulated a rapid cooxidation of ascorbic acid at pH 7.45. Hydrogen peroxide was among the products of cooxidation. The combination of homogentisic acid and Fe3+-EDTA stimulated hydroxyl radical (OH.) formation estimated by salicylate hydroxylation. Ferric iron was required for the reaction and Fe3+-EDTA was a better catalyst than either free Fe3+ or Fe3+-DTPA. SOD accelerated OH. production by homogentisic acid as did H2O2, and catalase reversed much of the stimulation by SOD. Catalase alone, and the hydroxyl radical scavengers, thiourea and sodium formate, suppressed salicylate hydroxylation. Homogentisic acid and Fe3+-EDTA also stimulated the degradation of hyaluronic acid, the chief viscous element of synovial fluid. Hyaluronic acid depolymerization was time dependent and proportional to the homogentisic acid concentration up to 100 microM. The level of degradation observed was comparable to that obtained with ascorbic acid at equivalent concentrations. The hydroxyl radical was an active intermediate in depolymerization. Thus, catalase and the hydroxyl radical scavengers, thiourea and dimethyl sulfoxide, almost completely suppressed the depolymerization reaction. The ability of homogentisic acid to generate O2-, H2O2 and OH. through autoxidation and the degradation of hyaluronic acid by homogentisic acid-mediated by OH. production suggests that oxygen radicals play a significant role in the etiology of alkaptonuric arthritis.  相似文献   

14.
Pterin-free inducible nitric oxide synthase (iNOS) was reconstituted with tetrahydrobiopterin (H(4)B) or tetrahydrobiopterin analogues (5-methyl-H(4)B and 4-amino-H(4)B), and the ability of bound 5-methyl-H(4)B and 4-amino-H(4)B to support catalysis by either full-length iNOS (FLiNOS) or the isolated heme domain (HDiNOS) was examined. In a single turnover with HDiNOS, 5-methyl-H(4)B forms a very stable radical, 5-methyl-H(3)B(*), that accumulates in the arginine reaction to approximately 60% of the HDiNOS concentration and decays approximately 400-fold more slowly than H(3)B(*) (0.0003 vs 0.12 s(-1)). The amount of radical (5-methyl-H(3)B(*) or H(3)B(*)) observed in the NHA reaction is very small (<3% of HDiNOS). The activity of 5-methyl-H(4)B-saturated FLiNOS and HDiNOS is similar to that when H(4)B is bound: arginine is hydroxylated to NHA, and NHA is oxidized exclusively to citrulline and (*)NO. A pterin radical was not observed with 4-amino-H(4)B- or pterin-free HDiNOS with either substrate. The catalytic activity of 4-amino-H(4)B-bound FLiNOS and HDiNOS resembles that of pterin-free iNOS: the hydroxylation of arginine is very unfavorable (<2% that of H(4)B-bound iNOS), and NHA is oxidized to a mixture of amino acid products (citrulline and cyanoornithine) and NO(-) rather than (*)NO. These results demonstrate that the bound pterin cofactor undergoes a one-electron oxidation (to form a pterin radical), which is essential to its ability to support normal NOS turnover. Although binding of H(4)B also stabilizes the NOS structure and active site, the most critical role of the pterin cofactor in NOS appears to be in electron transfer.  相似文献   

15.
Oxidative deposition of iron in ferritin or the autoxidation of iron in the absence of protein produces radicals from Good's buffers. Radical species are formed from the piperazine ring-based buffers Hepes (4-(2-hydroxyethyl)-1-piperazineethanesulfonic acid), Epps 4-(2-hydroxyethyl)-1-piperazinepropanesulfonic acid, and Pipes 1,4-piperazinediethanesulfonic acid, but not from Mes (4-morpholineethanesulfonic acid) which contains a morpholine ring. The radicals all have half-lives around 10 min and display very similar electron paramagnetic resonance spectra consisting of at least 30 lines. The Hepes radical can be formed by the addition of potassium superoxide directly to the buffer and its production during iron(II) autoxidation is inhibited by superoxide dismutase (EC 1.15.1.1). Catalase (EC 1.11.1.6) accelerates the decay of the EPR spectrum. Thus the buffer radicals are secondary radical species produced from oxygen radicals formed during the iron catalyzed Haber-Weiss process. The deoxyribose/thiobarbituric acid assay for hydroxyl radical production shows that Hepes is an effective hydroxyl radical scavenging agent. The Hepes radical can also be formed electrolytically at a potential of +0.8 V (vs standard hydrogen electrode). Oxidation of Hepes at pH 10 during the autoxidation of iron(II) or by the addition of hydrogen peroxide produces a nitroxide radical. These results indicate that piperazine ring Good buffers should be avoided in studies of redox processes in biochemistry.  相似文献   

16.
Tetrahydrobiopterin (BH4) serves as a critical co-factor for the endothelial nitric-oxide synthase (eNOS). A deficiency of BH4 results in eNOS uncoupling, which is associated with increased superoxide and decreased NO* production. BH4 has been suggested to be a target for oxidation by peroxynitrite (ONOO-), and ascorbate has been shown to preserve BH4 levels and enhance endothelial NO* production; however, the mechanisms underlying these processes remain poorly defined. To gain further insight into these interactions, the reaction of ONOO- with BH4 was studied using electron spin resonance and the spin probe 1-hydroxy-3-carboxy-2,2,5-tetramethyl-pyrrolidine. ONOO- reacted with BH4 6-10 times faster than with ascorbate or thiols. The immediate product of the reaction between ONOO- and BH4 was the trihydrobiopterin radical (BH3.), which was reduced back to BH4 by ascorbate, whereas thiols were not efficient in recycling of BH4. Uncoupling of eNOS caused by peroxynitrite was investigated in cultured bovine aortic endothelial cells (BAECs) by measuring superoxide and NO* using spin probe 1-hydroxy-3-methoxycarbonyl-2,2,5,5-tetramethyl-pyrrolidine and the NO*-spin trap iron-diethyldithiocarbamate. Bolus ONOO-, the ONOO- donor 3-morpholinosydnonimine, and an inhibitor of BH4 synthesis (2,4-diamino-6-hydroxypyrimidine) uncoupled eNOS, increasing superoxide and decreasing NO* production. Exogenous BH4 supplementation restored endothelial NO* production. Treatment of BAECs with both BH4 and ascorbate prior to ONOO- prevented uncoupling of eNOS by ONOO-. This study demonstrates that endothelial BH4 is a crucial target for oxidation by ONOO- and that the BH4 reaction rate constant exceeds those of thiols or ascorbate. We confirmed that ONOO- uncouples eNOS by oxidation of tetrahydrobiopterin and that ascorbate does not fully protect BH4 from oxidation but recycles BH3. radical back to BH4.  相似文献   

17.
One-electron oxidation of (6R)-5,6,7,8-tetrahydrobiopterin (H(4)B) by the azide radical generates the radical cation (H(4)B(*)(+)) which rapidly deprotonates at physiological pH to give the neutral trihydrobiopterin radical (H(3)B(*)); pK(a) (H(4)B(*)(+) <==> H(3)B(*) + H(+)) = (5.2 +/- 0.1). In the absence of ascorbate both the H(4)B(*)(+) and H(3)B(*) radicals undergo disproportionation to form quinonoid dihydrobiopterin (qH(2)B) and the parent H(4)B with rate constants k(H(4)B(*)(+) + H(4)B(*)(+)) = 6.5 x 10(3) M(-1) s(-1) and k(H(3)B(*) + H(3)B(*)) = 9.3 x 10(4) M(-1) s(-1), respectively. The H(3)B(*) radical is scavenged by ascorbate (AscH(-)) with an estimated rate constant of k(H(3)B(*) + AscH(-)) similar 1.7 x 10(5) M(-1) s(-1). At physiological pH the pterin rapidly scavenges a range of biological oxidants often associated with cellular oxidative stress and nitric oxide synthase (NOS) dysfunction including hydroxyl ((*)OH), nitrogen dioxide (NO(2)(*)), glutathione thiyl (GS(*)), and carbonate (CO(3)(*-)) radicals. Without exception these radicals react appreciably faster with H(4)B than with AscH(-) with k(*OH + H(4)B) = 8.8 x 10(9) M(-1) s(-1), k(NO(2)(*) + H(4)B) = 9.4 x 10(8) M(-1) s(-1), k(CO(3)(*-) + H(4)B) = 4.6 x 10(9) M(-1) s(-1), and k(GS(*) + H(4)B) = 1.1 x 10(9) M(-1) s(-1), respectively. The glutathione disulfide radical anion (GSSG(*-)) rapidly reduces the pterin to the tetrahydrobiopterin radical anion (H(4)B(*-)) with a rate constant of k(GSSG(*-) + H(4)B) similar 4.5 x 10(8) M(-1) s(-1). The results are discussed in the context of the general antioxidant properties of the pterin and the redox role played by H(4)B in NOS catalysis.  相似文献   

18.
7,8-Dihydroneopterin and neopterin are secreted by human and primate macrophages after activation by interferon-gamma in a ratio of 2:1. 7,8-Dihydroneopterin is known to suppress radical-mediated processes, but it is also able in the presence of iron ions to generate superoxide radical anion and hydroxyl radicals from molecular oxygen. Effects of 7,8-dihydroneopterin were investigated on (met)myoglobin and (met)hemoglobin. Addition of 7,8-dihydroneopterin to heme proteins in air-saturated solution resulted in dose-dependent cleavage of the porphyrin moiety. The liberation of non-heme iron and carbon monoxide originating from the cleaved porphyrin was quantified. Both were generated at equimolar concentrations with a linear correlation coefficient of 0.9. Addition of ferrous iron significantly accelerated the pteridine-mediated cleaving of the porphyrin. However, the total yield of porphyrin cleaved was controlled by the pterin rather than by the ferrous ion concentration. 7,8-Dihydroneopterin is assumed to reduce the heme iron in intact protein molecules, thereby preparing the conditions for binding of oxygen and carbon monoxide as ligands. Beyond that, it is concluded that hydroxyl radicals might be generated via reduction of molecular oxygen to superoxide anion in the autoxidation process and dismutation to hydrogen peroxide and subsequent Fenton reaction.  相似文献   

19.
Flavonoids containing phenol B rings, e.g. naringenin, naringin, hesperetin and apigenin, formed prooxidant metabolites that oxidised NADH upon oxidation by peroxidase/H2O2. Extensive oxygen uptake occurred which was proportional to the NADH oxidised and was increased up to twofold by superoxide dismutase. Only catalytic amounts of flavonoids and H2O2 were required indicating a redox cycling mechanism that activates oxygen and generates H2O2. NADH also prevented the oxidative destruction of flavonoids by peroxidase/H2O2 until the NADH was depleted. These results suggest that prooxidant phenoxyl radicals formed by these flavonoids cooxidise NADH to form NAD radicals which then activated oxygen. Similar oxygen activation mechanisms by other phenoxyl radicals have been implicated in the initiation of atherosclerosis and carcinogenesis by xenobiotic phenolic metabolites. This is the first time that a group of flavonoids have been identified as prooxidants independent of transition metal catalysed autoxidation reactions.  相似文献   

20.
Using Electron Spin Resonance spectroscopy at low temperatures, we find that thiyl radicals resulting from irradiation of frozen aqueous solutions of a variety of thiols, including cysteine, glutathione, and penicillamine react with oxygen to form sulfinyl (RSO.) radicals. The identity of the cysteine sulfinyl radical has been confirmed by the use of molecular oxygen isotopically labeled with 17O. Previous workers have suggested the reaction of thiyl radicals and molecular oxygen resulted in the formation of the potentially damaging thiol peroxyl radical, RSOO.; our work shows no evidence for this species. The sulfinyl radicals are suggested to result from a direct reaction between thiyl radicals and molecular oxygen. This reaction results in the cleavage of the dioxygen bond.  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