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 15 毫秒
1.
The aim of this study was to investigate the oxidation of two common fluorescent probes, dichlorodihydrofluorescein (DCFH2) and dihydrorhodamine (DHR), and their oxidized forms, dichlorofluorescein and rhodamine, by the radical products of peroxynitrite chemistry, *OH, NO2*, and CO3*-. At pH 8.0-8.2, rate constants for the interaction of carbonate radical with probes were estimated to be 2.6 x 10(8) x M(-1) s(-1) for DCFH2 and 6.7 x 10(8) M(-1) s(-1) for DHR. Nitrogen dioxide interacted more slowly than carbonate radical with these probes: the rate constant for the interaction between NO2* and DCFH2 was estimated as 1.3 x 10(7) M(-1) s(-1). Oxidation of DHR by nitrogen dioxide led to the production of rhodamine, but the kinetics of these reactions were complex. Hydroxyl radical interacted with both probes with rate constants close to the diffusion-controlled limit. We also found that oxidized forms of these fluorescent probes reacted rapidly with carbonate, nitrogen dioxide, and hydroxyl radicals. These data suggest that probe oxidation may often be in competition with reaction of the radicals with cellular antioxidants.  相似文献   

2.
Antioxidant properties of melatonin: a pulse radiolysis study   总被引:5,自引:0,他引:5  
Various one-electron oxidants such as OH*, tert-BuO*, CCl3OO*, Br2*- and N3*, generated pulse radiolytically in aqueous solutions at pH 7, were scavenged by melatonin to form two main absorption bands with lambda(max) = 335 nm and 500 nm. The assignment of the spectra and determination of extinction coefficients of the transients have been reported. Rate constants for the formation of these species ranged from 0.6-12.5x10(9) dm3 mol(-1) s(-1). These transients decayed by second order, as observed in the case of Br2*- and N3* radical reactions. Both the NO2* and NO* radicals react with the substrate with k = 0.37x10(7) and 3x10(7) dm3 mol(-1) s(-1), respectively. At pH approximately 2.5, the protonated form of the transient is formed due to the reaction of Br2*- radical with melatonin, pKa ( MelH* <=> Mel* + H+) = 4.7+/-0.1. Reduction potential of the couple (Mel*/MelH), determined both by cyclic voltammetric and pulse-radiolytic techniques, gave a value E(1)7 = 0.95+/-0.02 V vs. NHE. Repair of guanosine radical and regeneration of melatonin radicals by ascorbate and urate ions at pH 7 have been reported. Reactions of the reducing radicals e(aq)- and H* atoms with melatonin have been shown to occur at near diffusion rates.  相似文献   

3.
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   

4.
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   

5.
Arrhenius parameters for the reactions of oxidizing hydroxyl radicals and reducing hydrated electrons with cisplatin, transplatin and carboplatin in aqueous solution have been determined using pulsed electron radiolysis and absorption spectroscopy techniques. Under physiological pH and chloride concentration conditions, hydroxyl radical reaction rate constants of (9.99 +/- 0.20) x 10(9), (8.38 +/- 0.55) x 10(9), and (6.03 +/- 0.08) x 10(9) M(-1) s(-1) at 24.0, 20.7 and 24.0 degrees C, respectively, with corresponding activation energies of 12.79 +/- 0.57, 13.88 +/- 1.14, and 14.35 +/- 0.56 kJ mol(-1) for these three reactions, were determined. These oxidations of cisplatin and transplatin to form a Pt(III) transient are significantly faster than reported previously at room temperature. The lower rate constant for carboplatin is consistent with hydroxyl radical reaction partitioning between reaction at the platinum center and the cyclobutanedicarboxylate ligand. The equivalent reductive hydrated electron reaction rate constants measured were (1.99 +/- 0.04) x 10(10) (24.0 degrees C), (1.77 +/- 0.08) x 10(10) (22.0 degrees C), and (8.92 +/- 0.06) x 10(9) M(-1) s(-1) (24.0 degrees C), with corresponding activation energies of 15.75 +/- 1.00, 19.74 +/- 1.82, and 19.99 +/- 0.34 kJ mol(-1). Again, the values determined for cisplatin and transplatin are faster than reported; however, all three values are consistent with direct reduction of the platinum center to form a Pt(I) moiety.  相似文献   

6.
Shi Y  Lin W  Fan B  Jia Z  Yao S  Kang J  Wang W  Zheng R 《Biochimica et biophysica acta》1999,1472(1-2):115-127
DNA damaged by oxygen radicals has been implicated as a causative event in a number of degenerative diseases, including cancer and aging. So it is very significant to look for ways in which either oxygen radicals are scavenged prior to DNA damage or damaged DNA is repaired to supplement the cells' inadequate repair capacity. The repair activities and reaction mechanism of phenylpropanoid glycosides (PPGs) and their derivatives, isolated from Chinese folk medicinal herbs, towards both dGMP-OH* adducts and dAMP-OH* adducts were studied with the pulse radiolytic technique. On pulse irradiation of nitrous oxide saturated 2 mM dGMP or dAMP aqueous solution containing one of the PPGs or their derivatives, the transient absorption spectra of the hydroxyl adduct of dGMP or dAMP decayed with the formation of that of phenoxyl radicals of PPGs or their derivatives within several decades of microseconds after electron pulse irradiation. The result indicated that dGMP or dAMP hydroxyl adducts can be repaired by PPGs or their derivatives. The rate constants of the repair reactions were deduced to be 0.641-1.28 x 10(9) M(-1) s(-1) for dGMP-OH* and 0.2-0.491 x 10(9) M(-1) s(-1) for dAMP-OH*, which positively correlated to the number of phenolic hydroxyl groups in the glycoside structure. A deeper understanding of this new repair mechanism may help researchers to design strategies to prevent and/or intervene more effectively in free radical related diseases.  相似文献   

7.
Sunlight mediated hydroxyl radical production from aqueous ferric perchlorate at low pH has been investigated using deoxyribose-thiobarbituric acid assay. The rate of production of hydroxyl radical was found to be dependent on the time of irradiation. Hydroxyl radical scavengers can compete with deoxyribose for hydroxyl radicals produced in the system leading to a decreased yield of thiobarbituric acid chromogen. The second-order rate constants of the added scavengers can be determined using a simple competition kinetic method. The rate constants for the reaction of hydroxyl radical with a number of purine and pyrimidine derivatives were determined using this method. The rate constants obtained (1-7 x 10(9) dm(3) mol(-1) s(-1)) were found to be in good agreement with those reported using pulse radiolysis technique. The rate constants of dimethyluracil, xanthosine, amino and methyl substituted pyrimidines, cytidine monophosphate and uridine monophosphate were also determined by this method. It is proposed that sunlight mediated production of hydroxyl radical coupled with deoxyribose-thiobarbituric acid assay is a simple and efficient method for the determination of rate constants for the reaction of hydroxyl radical with a wide range of biomolecules.  相似文献   

8.
The rate constants of the reactions of alcohol dehydrogenase and glyceraldehyde-3-phosphate dehydrogenase with hydroxyl radicals were determined using the method of steady-state competitive reactions. Ethanol was used as a scavenger of hydroxyl radicals. The rate constants of the reactions of hydroxyl radicals with alcohol dehydrogenase and glyceraldehyde-3-phosphate dehydrogenase were found to be 2.8 x 10(12) dm(3) mol(-1) s(-1), and 1.6 x 10(12) dm(3) mol(-1) s(-1), respectively.  相似文献   

9.
Shi Y  Kang J  Lin W  Fan P  Jia Z  Yao S  Wang W  Zheng R 《Biochimica et biophysica acta》1999,1472(1-2):279-289
The repair effects on deoxynucleotide radical cations of phenylpropanoid glycosides (PPGs) and their analogs, isolated from a Chinese folk medicinal herb, were studied using the pulse radiolysis technique. The radical cations of deoxynucleotides were formed by the reaction of SO4*- with deoxynucleotides. On pulse irradiation of a nitrogen saturated deoxynucleotide aqueous solution containing 20 mM K2S2O8, 200 mM t-BuOH and one of the PPGs or their analogs, the transient absorption spectra of the radical cations of nucleotide decayed with the formation of those of the radical cation of PPGs or their analogs within several tens of microseconds after electron pulse irradiation. The result indicates that deoxynucleotide radical cations can be repaired by PPGs or their analogs. The rate constants of the repair reactions were determined to be 0.48-1.1 x 10(9), 0.64-1.80 x 10(9) and 2.12-4.4 x 10(9) M(-1) s(-1) for dAMP, dGMP and dCMP radical cations respectively. It is obvious that the rate constants of the repair reaction depend on the number of phenolic hydroxyl groups contained in the PPGs and their analogs. A deeper understanding of this new repair mechanism will undoubtedly help researchers design strategies to prevent and/or intervene more effective in free radical related diseases.  相似文献   

10.
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   

11.
The kinetics and spectra of the reactions of clozapine with compounds I and II of myeloperoxidase were investigated using both single- and sequential-mixing stopped-flow techniques, steady-state kinetics, and spectrophotometric measurements. The results show conclusively that both compounds I and II are reduced in one-electron reactions with clozapine. At pH 7.0 the rate constant for compound I reacting with clozapine is (1.5 +/- 0.1) x 10(6) M(-1) s(-1) and for compound II (4.8 +/- 0.1) x 10(4) M(-1) s(-1). The physiological pH of 7.4 was found to be optimal for the oxidation of clozapine by compound I. The rate constant for compound I reacting with ascorbate is (1.1 +/- 0.1) x 10(6) M(-1) s(-1) and for compound II (1.1 +/- 0.2) x 10(4) M(-1) s(-1), both obtained at pH 7.0. Experiments with both clozapine and ascorbate present showed that ascorbate acts both as a competitive inhibitor and free radical scavenger.  相似文献   

12.
Herold S  Exner M  Nauser T 《Biochemistry》2001,40(11):3385-3395
The second-order rate constants for the reactions between nitrogen monoxide and oxymyoglobin or oxyhemoglobin, determined by stopped-flow spectroscopy, increase with increasing pH. At pH 7.0 the rates are (43.6 +/- 0.5) x 10(6) M(-1) x s(-1) for oxymyoglobin and (89 +/- 3) x 10(6) M(-1) x s(-1) for oxyhemoglobin (per heme), whereas at pH 9.5 they are (97 +/- 3) x 10(6) M(-1) x s(-1) and (144 +/- 3) x 10(6) M(-1) x s(-1), respectively. The rate constants for the reaction between oxyhemoglobin and NO* depend neither on the association grade of the protein (dimer/tetramer) nor on the concentration of the phosphate buffer (100-1 mM). The nitrogen monoxide-mediated oxidations of oxymyoglobin and oxyhemoglobin proceed via intermediate peroxynitrito complexes which were characterized by rapid scan UV/vis spectroscopy. The two complexes MbFe(III)OONO and HbFe(III)OONO display very similar spectra with absorption maxima around 500 and 635 nm. These species can be observed at alkaline pH but rapidly decay to the met-form of the proteins under neutral or acidic conditions. The rate of decay of MbFe(III)OONO increases with decreasing pH and is significantly larger than those of the analogous complexes of the two subunits of hemoglobin. No free peroxynitrite is formed during these reactions, and nitrate is formed quantitatively, at both pH 7.0 and 9.0. This result indicates that, as confirmed from protein analysis after reacting the proteins with NO* for 10 times, when peroxynitrite is coordinated to the heme of myoglobin or hemoglobin it rapidly isomerizes to nitrate without nitrating the globins in physiologically significant amounts.  相似文献   

13.
The reaction rate constants and transient spectra of 11 flavonoids and 4 phenolic acids reacting with e(aq)- at neutral pH were measured. Absorption bands of the transients of e(aq)- reacting with the above compounds all located at a wavelength shorter than 400 nm. The e(aq)- scavenging abilities were divided into three groups: (+)catechin ((1.2 +/-0.1) x 10(8) M(-1)s(-1)) < 4-chromanol ((4.4 +/- 0.4) x 10(8) M(-1)s(-1)) < genistein ((6.2+/-0.4) x 10(9) M (-1) s(-1) approximately genistin ((8 +/- 1) x 10(9) M(-1)s(-1)) approximately rutin ((7.6 +/- 0.4) x M(-1)s(-1) approximately caffeic acid ((8.3 +/- 0.5) x 10(9)M(-1)s(-1)) < transcinnamic acid((1.1 +/- 0.1) x 10(10) M(-1)s(-1)) approximately p-coumaric acid ((1.1 +/- 0.1) x 10(10) M(-1)s(-1) approximately 2,4,6-trihydroxylbenzoic acid((1.1 +/- 0.1) x 10(10) M(-1)s(-1)) approximately baicalein ((1.1 +/- 0.5) x 10(10) M(-1)s(-1)) approximately baicalin((1.3 + 0.1) X 10(10) M(-1)s(-1)) approximately naringenin ((1.2 +/- 0.1) x 10(10) M(-1)s(-1)) approximately naringin ((1.0 +/- 0.1) x 10(10) M(-1)s(-1)) approximately gossypin((1.2 +/- 0.1) x 10(10) M(-1)s(-1)) approximately quercetin((1.3 +/- 0.5) x 10(10) M(-1)s(-1)). These results suggested that C4 keto group is the active site for e(aq)- to attack on flavonoids and phenolic acids, whereas the o-dihydroxy structure in B ring, the C2,3 double bond, the C3-OH group, and glucosylation, which are key structures that influence the antioxidant activities of flavonoids and phenolic acids, have little effects on the e(aq)- scavenging activities.  相似文献   

14.
The rate constants for reduction of methemerythrin from Phascolopsis gouldii and Themiste pyroides by hydrated electrons are 2.0 and 3.9 x 10(9) M(-1)s(-1), respectively, at pH 8.2, I = 0.03 M, and 25 degrees C. There is only a small increase in rate when the pH is lowered to 6.3 and a very small decrease when the ionic strength is raised to 0.1 M. Adding azide ion (to form the met-azide adduct) has little effect on the reactivity towards e-aq. For the monomer form, metmyohemerythrin from T. pyroides, the reaction rate constant is 4.5 x 10(9) M(-1)s(-1). Methemerythrin from T. pyroides reacts with CO2- with a rate constant 6.8 x 10(7) M(-1)s(-1). The reactivity sequence e-aq greater than CO2- greater than SO2- (from dithionite reduction) towards methemerythrin is the same as that observed with reduction of heme proteins but the rate constants are some 10 to 100 times smaller for the former. Only 10 to 20% of the e-aq or CO2- radicals generated effect reduction of the iron centers in methemerythrin.  相似文献   

15.
Quantifying oxygen radicals that arise during the redox cycling of quinone-containing anticancer agents such as diaziquone (AZQ) has been difficult, as has been their detection at low drug concentrations. This is due to the fact that EPR spin trapping, the method most often used for *OH detection, requires the use of high drug concentrations. Using a new highly sensitive technique that employs a fluorescamine-derivatized nitroxide, we show that low levels of NADPH-cytochrome P450 reductase (4.25 microg/ml) catalyze the production of hydroxyl radicals at very low, clinically relevant AZQ concentrations. Thus, at this enzyme concentration, we were able to detect a rate of 0.10 nM s(-1) hydroxyl radical production by 5 microM AZQ, a clinically relevant concentration. The Michaelis-Menten constants for AZQ-mediated hydroxyl radical production are: K(M) = 10.7 +/- 1.4 microM, and V(max) = 5.2 +/- 0.9 x 10(-8) M s(-1) (mg protein)(-1). Experiments employing catalase, superoxide dismutase, and NADPH-cytochrome P450 reductase, confirm the previously deduced conclusions from high drug concentrations, that is, that at low concentrations, AZQ acts to shuttle reducing equivalents from the enzyme to oxygen, thus generating the redox cycle. The data presented here suggest that the levels and locations of redox active metal ions may be the principal controlling factor in the pathway of AZQ activity that involves oxidative stress.  相似文献   

16.
We directly measured the absolute reactivity of 17-beta-estradiol (E2) and several phenolic model compounds for E2 toward t-butoxy radical (t-BuO*) by nanosecond time-resolved optical spectroscopy. Compared to other phenols, E2 is a moderate, but not strong deactivator of oxyradicals. The absolute bimolecular rate constant for H-atom transfer from E2 to t-BuO* is 1.3 +/- 0.3 x 10(9) M(-1) x s(-1) (23 degrees C, benzene). We estimate the O-H bond strength of 17-beta-estradiol to be approximately 85 +/- 2 kcal/mol and calculate the reaction rate constant of E2 toward peroxy radical to be 10(5) M(-1) x s(-1) at 37 degrees C. The conjugate phenoxy radical of 17-beta-estradiol, E2O*, is unusually reactive toward alpha-tocopherol and ascorbate by H-atom transfer in homogeneous solution (10(8)-10(9) M(-1) x s(-1)). Our findings suggest that E2 functions in vivo as a highly localized, synergistic biological antioxidant. This may partly explain the clinical effectiveness of ovarian steroids in delaying the manifestations of Alzheimer's Disease as well as in protecting against cardiovascular pathologies. In the absence of complementary antioxidant synergists, E2O* is expected to be a pro-oxidant.  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.
Nitric oxide-derived oxidants (e.g., peroxynitrite) are believed to participate in antimicrobial activities as part of normal host defenses but also in oxidative tissue injury in inflammatory disorders. A similar role is ascribed to the heme enzyme myeloperoxidase (MPO), the most abundant protein of polymorphonuclear leukocytes, which are the terminal phagocytosing effector cells of the innate immune system. Concomitant production of peroxynitrite and release of millimolar MPO are characteristic events during phagocytosis. In order to understand the mode of interaction between MPO and peroxynitrite, we have performed a comprehensive stopped-flow investigation of the reaction between all physiological relevant redox intermediates of MPO and peroxynitrite. Both iron(III) MPO and iron(II) MPO are rapidly converted to compound II by peroxynitrite in monophasic reactions with calculated rate constants of (6.8+/-0.1) x 10(6) M(-1)s(-1) and (1.3+/-0.2) x 10(6) M(-1)s(-1), respectively (pH 7.0 and 25 degrees C). Besides these one- and two-electron reduction reactions of peroxynitrite, which produce nitrogen dioxide and nitrite, a one-electron oxidation to the oxoperoxonitrogen radical must occur in the fast monophasic transition of compound I to compound II mediated by peroxynitrite at pH 7.0 [(7.6+/-0.1) x 10(6) M(-1)s(-1)]. In addition, peroxynitrite induced a steady-state transition from compound III to compound II with a rate of (1.0+/-0.3) x 10(4) M(-1)s(-1). Thus, the interconversion among the various oxidation states of MPO that is prompted by peroxynitrite is remarkable. Reaction mechanisms are proposed and the physiological relevance is discussed.  相似文献   

19.
The oxidation of serotonin (5-hydroxytryptamine) by the myeloperoxidase intermediates compounds I and II was investigated by using transient-state spectral and kinetic measurements at 25.0 +/- 0.1 degrees C. Rapid scan spectra demonstrated that both compound I and compound II oxidize serotonin via one-electron processes. Rate constants for these reactions were determined using both sequential-mixing and single-mixing stopped-flow techniques. The second order rate constant obtained for the one-electron reduction of compound I to compound II by serotonin is (1.7 +/- 0.1) x 10(7) M(-1) x s(-1), and that for compound II reduction to native enzyme is (1.4 +/- 0.1) x 10(6) M(-1) x s(-1) at pH 7.0. The maximum pH of the compound I reaction with serotonin occurs in the pH range 7.0-7.5. At neutral pH, the rate constant for myeloperoxidase compound I reacting with serotonin is an order of magnitude larger than for its reaction with chloride, (2.2 +/- 0.2) x 10(6) M(-1) x s(-1). A direct competition of serotonin with chloride for myeloperoxidase compound I oxidation was observed. Our results suggest that serotonin may have a role to protect lipoproteins from oxidation and to prevent enzymes from inactivation caused by the potent oxidants HOCl and active oxygen species.  相似文献   

20.
The reactions of the carbonate and dichloride radical anions, CO3- and Cl2-, with the extracellular matrix glycosaminoglycan hyaluronan (HA) have been studied using the kinetic technique of pulse radiolysis and also by steady-state irradiation combined with gel permeation chromatography/multiangle laser light scattering(gpc/MALLS) to measure the rates of reaction with HA and the yield of HA chain scission, respectively. For comparison, the same measurements were made for the reactions of the free radicals *OH, Br2*-, and N3*. The carbonate and dichloride radical anions were found to react relatively quickly with HA (7.0 x 10(5) and 6.9 x 10(6) dm3 mol(-1) s(-1), respectively) although they are much less reactive than the hydroxyl radical, *OH. Significant yields (20 and 38%, respectively) of chain scission of HA by these radical anions were also determined from the gpc/MALLS experiments, providing some support for their potential participation in the depolymerization of HA in vivo. These results are compared with data obtained for the other free radicals (hydroxyl, azide radicals, and dibromide radical anions) investigated in this study in order to gain an insight into their mechanism of reaction with HA. Earlier chain scission yields of HA by hydroxyl radicals determined by the authors have also been revised using the gpc/MALLS technique employed in the current study. The yields of 52% (absence of air) and 44% (in air) are much lower than the previous values. In the current study, the effect of oxygen on the yields of HA chain breaks is discussed in terms of the reactivity of HA peroxyl radicals in the presence of superoxide radical anions. The relevance of the results of this study to mechanisms of inflammation is discussed.  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