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 93 毫秒
1.
We report here on the role of copper (II) salts on the acceleration of peroxynitrite (ONOO-) decomposition and ONOO- reaction with the anaesthetic agent propofol (2,6-diisopropylphenol) in alkaline medium. We observed a strong acceleration of the ONOO- decomposition in alkaline medium in the presence of copper (I and II) salts. After 18 h of ONOO- reaction with propofol, we observed nitrosated, nitrated, and oxidized (quinone and diphenylquinone) derivatives of propofol, but in the presence of Cu(II) (20% molar vs ONOO-), the yields of quinone and nitrosopropofol strongly increased. We also observed that the temperature and the atmosphere influenced the effects of Cu(II) on ONOO- reactions with propofol: low temperatures promoted nitrosation and high temperatures promoted oxidation; O2 atmosphere increased the general reactivity and the yield of nitrated and oxidized products. We highlighted the influence of Cu(II) salts on the radical character of the reaction by direct EPR technique. The exact mechanism of the Cu(II) catalysis remains unexplained, but we suggest the formation of a copper complex with propofol or, more probably, the oxidation of ONOO- into ONOO. by copper ions promoting the formation of quinone and nitrosopropofol according to a previously reported mechanism [M. Cudic, C. Ducrocq, Transformations of 2,6-diisopropylphenol by NO-derived nitrogen oxides, particularly peroxynitrite, Nitric Oxide 4 (2000) 147-156].  相似文献   

2.
M Cudic  C Ducrocq 《Nitric oxide》2000,4(2):147-156
To investigate the protective effect of the anesthetic 2, 6-diisopropylphenol, or propofol, in oxidative processes in which (*)NO and peroxynitrite are involved, direct interactions were explored. The reactions of the highly lipophilic propofol with (*)NO in methanolic or aqueous buffered solutions under air were shown to produce the same compounds as those detected with peroxynitrite, but with very low yields and slow rates. In aqueous neutral medium, peroxynitrite (ONOO(-), ONOOCO(-)(2), ONOOH) was able to nitrate and oxidize propofol: In addition to oxidation products, quinone and quinone dimer, the formation of the 4-nitropropofol derivative was detected, increasing with peroxynitrite or CO(2) concentrations. Nitration reached 20% after the addition of 25 mM bicarbonate to an equimolecular mixture of peroxynitrite and propofol in methanol/phosphate-buffered solution (1/4,v/v) at pH 7.4. However, peroxynitrite either in methanol or in alkaline-buffered mixture (optimum pH 10-12) resulted in the rapid and almost complete transformation of propofol to an intermediate compound 1, which further decomposed to 4-nitrosopropofol. The transient compound 1 was obtained from either peroxynitrite or (*)NO in the presence of oxygen. From mass spectrometry determination of compound 1 we propose the involvement of the nitrosodioxyl radical ONOO(*), forming an adduct with the propofoxyl radical, to yield 4-nitrosodioxypropofol and finally 4-nitrosopropofol.  相似文献   

3.
We have reported previously that the apparent rate of peroxynitrite (ONOO(-) ) decay, as followed from its absorbance at 302 nm, decreases in the presence of hydrogen peroxide, mannitol and ethanol (Alvarez et al., 1995, Chem. Res. Toxicol. 8:859-864; Alvarez et al., 1998, Free Radic. Biol. Med. 24:1331-1337). Recently, two papers confirmed the observation and proposed that this slowing effect was due to the formation of absorbing peroxynitrate (O(2) NOO(-) ) as intermediate (Goldstein and Czapski, 1998, J. Am. Chem. Soc. 120:3458-3463; Hodges and Ingold, 1999, J. Am. Chem. Soc. 121:10695-10701). Peroxynitrate would be formed from the reaction of peroxynitrite-derived nitrogen dioxide with superoxide. Superoxide, in turn, would arise from the one-electron oxidation of hydrogen peroxide, or from the reaction of reductive radicals derived from mannitol and ethanol with dioxygen. In agreement with this concept, we show herein that under the conditions of our previous work, the slowing effect is prevented by superoxide dismutase and, in the case of mannitol and ethanol, by reducing the dioxygen concentration of the reaction solutions. Thus, superoxide formation is necessary for the decrease in the rate of absorbance decay. In addition, by simulations using known rate constants and absorption coefficients, we show that the slowing effect can be quantitatively accounted for by the formation of peroxynitrate.  相似文献   

4.
Peroxynitrite (ONOO(((-)))/ONOOH) is expected in vivo to react predominantly with CO(2), thereby yielding NO(2)(.) and CO(3) radicals. We studied the inhibitory effects of ascorbate on both NADH and dihydrorhodamine 123 (DHR) oxidation by peroxynitrite generated in situ from 3-morpholinosydnonimine N-ethylcarbamide (SIN-1). SIN-1 (150 micrometer)-mediated oxidation of NADH (200 micrometer) was half-maximally inhibited by low ascorbate concentrations (61-75 micrometer), both in the absence and presence of CO(2). Control experiments performed with thiols indicated both the very high antioxidative efficiency of ascorbate and that in the presence of CO(2) in situ-generated peroxynitrite exclusively oxidized NADH via the CO(3) radical. This fact is attributed to the formation of peroxynitrate (O(2)NOO(-)/O(2)NOOH) from reaction of NO(2)(.) with O(2), which is formed from reaction of CO(3) with NADH. SIN-1 (25 micrometer)-derived oxidation of DHR was half-maximally inhibited by surprisingly low ascorbate concentrations (6-7 micrometer), irrespective of the presence of CO(2). Control experiments performed with authentic peroxynitrite revealed that ascorbate was in regard to both thiols and selenocompounds much more effective to protect DHR. The present results demonstrate that ascorbate is highly effective to counteract the oxidizing properties of peroxynitrite in the absence and presence of CO(2) by both terminating CO(3)/HO( small middle dot) reactions and by its repair function. Ascorbate is therefore expected to act intracellulary as a major peroxynitrite antagonist. In addition, a novel, ascorbate-independent protection pathway exists: scavenging of NO(2)(.) by O(2) to yield O(2)NOO(-), which further decomposes into NO(2)(-) and O(2).  相似文献   

5.
An electron-rich iron(III) porphyrin complex (meso-tetramesitylporphinato)iron(III) chloride [Fe(TMP)Cl], was found to catalyze the epoxidation of olefins by aqueous 30% H2O2 when the reaction was carried out in the presence of 5-chloro-1-methylimidazole (5-Cl-1-MeIm) in aprotic solvent. Epoxides were the predominant products with trace amounts of allylic oxidation products, indicating that Fenton-type oxidation reactions were not involved in the olefin epoxidation reactions. cis-Stilbene was stereospecifically oxidized to cis-stilbene oxide without giving isomerized trans-stilbene oxide product, demonstrating that neither hydroperoxy radical (HOO·) nor oxoiron(IV) porphyrin [(TMP)FeIV=O] was responsible for the olefin epoxidations. We also found that the reactivities of other iron(III) porphyrin complexes such as (meso-tetrakis(2,6-dichlorophenyl)porphinato)iron(III) chloride [Fe(TDCPP)Cl], (meso-tetrakis(2,6-difluorophenyl)porphinato)iron(III) chloride [Fe(TDFPP)Cl], and (meso-tetrakis(pentafluorophenyl)porphinato)iron(III) chloride [Fe(TPFPP)Cl] were significantly affected by the presence of the imidazole in the epoxidation of olefins by H2O2. These iron porphyrin complexes did not yield cyclohexene oxide in the epoxidation of cyclohexene by H2O2 in the absence of 5-Cl-1-MeIm in aprotic solvent; however, addition of 5-Cl-1-MeIm to the reaction solutions gave high yields of cyclohexene oxide with the formation of trace amounts of allylic oxidation products. We proposed, on the basis of the results of mechanistic studies, that the role of the imidazole is to decelerate the O–O bond cleavage of an iron(III) hydroperoxide porphyrin (or H2O2–iron(III) porphyrin adduct) and that the intermediate transfers its oxygen to olefins prior to the O–O bond cleavage.  相似文献   

6.
We have reported previously that the apparent rate of peroxynitrite (ONOO-) decay, as followed from its absorbance at 302 nm, decreases in the presence of hydrogen peroxide, mannitol and ethanol (Alvarez et al., 1995, Chem. Res. Toxicol. 8:859-864; Alvarez et al., 1998, Free Radic. Biol. Med. 24:1331–1337). Recently, two papers confirmed the observation and proposed that this slowing effect was due to the formation of absorbing peroxynitrate (O2NOO-) as intermediate (Goldstein and Czapski, 1998, J. Am. Chem. Soc. 120:3458–3463; Hodges and Ingold, 1999, J. Am. Chem. Soc. 121:10695–10701). Peroxynitrate would be formed from the reaction of peroxynitrite-derived nitrogen dioxide with superoxide. Superoxide, in turn, would arise from the one-electron oxidation of hydrogen peroxide, or from the reaction of reductive radicals derived from mannitol and ethanol with dioxygen. In agreement with this concept, we show herein that under the conditions of our previous work, the slowing effect is prevented by superoxide dismutase and, in the case of mannitol and ethanol, by reducing the dioxygen concentration of the reaction solutions. Thus, superoxide formation is necessary for the decrease in the rate of absorbance decay. In addition, by simulations using known rate constants and absorption coefficients, we show that the slowing effect can be quantitatively accounted for by the formation of peroxynitrate.  相似文献   

7.
Thiols represent preferential targets of peroxynitrite in biological systems. In this work, we investigated the mechanisms and kinetics of the reaction of peroxynitrite with the dithiol dihydrolipoic acid (DHLA) and its oxidized form, lipoic acid (LA). Peroxynitrite reacted with DHLA being oxidation yields higher at alkaline pH. The stoichiometry for the reaction was two thiols oxidized per peroxynitrite. LA formation accounted for approximately 50% DHLA consumption at pH 7.4, probably reflecting secondary reactions between LA and peroxynitrite. Indeed, peroxynitrous acid reacted with LA with an apparent second-order rate constant (k(2app)) of 1400 M(-1) s(-1) at pH 7.4 and 37 degrees C. Nitrite and LA-thiosufinate were formed as reaction products. Surprisingly, the k(2app) for peroxynitrite-dependent DHLA oxidation was only 250 M(-1) s(-1) per thiol, at pH 7.4 and 37 degrees C. Testing various low-molecular-weight thiols, we found that an increase in the thiol pK (pK(SH)) value correlated with a decrease of k(2app) for the reaction with peroxynitrite at pH 7.4. The pK(SH) for DHLA is 10.7, in agreement with its modest reactivity with peroxynitrite.  相似文献   

8.
Peroxynitrite, a strong oxidant formed intravascularly in vivo, can diffuse onto erythrocytes and be largely consumed via a fast reaction (2 x 10(4) m(-1) s(-1)) with oxyhemoglobin. The reaction mechanism of peroxynitrite with oxyhemoglobin that results in the formation of methemoglobin remains to be elucidated. In this work, we studied the reaction under biologically relevant conditions using millimolar oxyhemoglobin concentrations and a stoichiometric excess of oxyhemoglobin over peroxynitrite. The results support a reaction mechanism that involves the net one-electron oxidation of the ferrous heme, isomerization of peroxynitrite to nitrate, and production of superoxide radical and hydrogen peroxide. Homolytic cleavage of peroxynitrite within the heme iron allows the formation of ferrylhemoglobin in approximately 10% yields, which can decay to methemoglobin at the expense of reducing equivalents of the globin moiety. Indeed, spin-trapping studies using 2-methyl-2-nitroso propane and 5,5 dimethyl-1-pyrroline-N-oxide (DMPO) demonstrated the formation of tyrosyl- and cysteinyl-derived radicals. DMPO also inhibited covalently linked dimerization products and led to the formation of DMPO-hemoglobin adducts. Hemoglobin nitration was not observed unless an excess of peroxynitrite over oxyhemoglobin was used, in agreement with a marginal formation of nitrogen dioxide. The results obtained support a role of oxyhemoglobin as a relevant intravascular sink of peroxynitrite.  相似文献   

9.
Peroxidases catalyze the oxidation of nitrite to nitrate in the presence of hydrogen peroxide. Two pathways may occur: one entailing the intermediate formation of NO(2) and the other implying the generation of peroxynitrite. The products of nitrite (NO(2) (-) ) oxidation by salivary peroxidase (SPO) and commercial bovine lactoperoxidase (LPO) are studied by utilizing an electrochemical assay that allows the direct, continuous monitoring of NO and/or NO(2) and by HPLC to assess nitrates at the end of the reaction. Dialyzed saliva and LPO, in the presence of H(2) O(2) , convert nitrite into nitrate and form some NO, with a molar ratio of 10(3) . In our experimental conditions, no NO(2) was detectable among the products of nitrite oxidation. SCN(-) inhibits NO formation and so does I(-) , although at higher concentrations. No effects are observed with Cl(-) or Br(-) . We conclude that SPO and LPO transform NO(2) (-) into nitrate-forming small amounts of NO in the presence of H(2) O(2) as an intermediate or a by-product, synthesized through the peroxynitrite pathway.  相似文献   

10.
Endothelial cells, macrophages, neutrophils, and neuronal cells generate superoxide (O2-) and nitric oxide (.NO) which can combine to form peroxynitrite anion (ONOO-). Peroxynitrite, known to oxidize sulfhydryls and to yield products indicative of hydroxyl radical (.OH) reaction with deoxyribose and dimethyl sulfoxide, is shown herein to induce membrane lipid peroxidation. Peroxynitrite addition to soybean phosphatidylcholine liposomes resulted in malondialdehyde and conjugated diene formation, as well as oxygen consumption. Lipid peroxidation was greater at acidic and neutral pH, with no significant lipid peroxidation occurring above pH 9.5. Addition of ferrous (Fe+2) or ferric (Fe+3) iron did not enhance lipid peroxide formation over that attributable to peroxynitrite alone. Diethylenetetraminepentacetic acid (DTPA) or iron removal from solutions by ion-exchange chromatography decreased conjugated diene formation by 25-50%. Iron did not play an essential role in initiating lipid peroxidation, since DTPA and iron depletion of reaction systems were only partially inhibitory. In contrast, desferrioxamine had an even greater concentration-dependent inhibitory effect, completely abolishing lipid peroxidation at 200 microM. The strong inhibitory effect of desferrioxamine on lipid peroxidation was due to direct reaction with peroxynitrous acid in addition to iron chelation. We conclude that the conjugate acid of peroxynitrite, peroxynitrous acid (ONOOH), and/or its decomposition products, i.e., .OH and nitrogen dioxide (.NO2), initiate lipid peroxidation without the requirement of iron. These observations demonstrate a potential mechanism contributing to O2-(-)and .NO-mediated cytotoxicity.  相似文献   

11.
Catalase-peroxidases (KatGs) are multifunctional heme peroxidases exhibiting an overwhelming catalase activity and a substantial peroxidase activity of broad specificity. Here, we show that catalase-peroxidases are also haloperoxidases capable of oxidizing chloride, bromide, and iodide in a peroxide- and enzyme-dependent manner. Recombinant KatG and the variants R119A, W122F, and W122A from the cyanobacterium Synechocystis PCC 6803 have been tested for their halogenation activity. Halogenation of monochlorodimedon (MCD), formation of triiodide and tribromide, and bromide- and chloride-mediated oxidation of glutathione have been tested. Halogenation of MCD by chloride, bromide, and iodide was shown to be catalyzed by wild-type KatG and the variant R119A. Generally, rates of halogenation increased in the order Cl(-) < Br(-) < I(-) and/or by decreasing pH. The halogenation activity of R119A was about 7-9% that of the wild-type enzyme. Upon exchange of the distal Trp122 by Phe and Ala, both the catalase and halogenation activities were lost but the overall peroxidase activity was increased. The findings suggest that the same redox intermediate is involved in H(2)O(2) and halide oxidation and that distal Trp122 is involved in both two-electron reactions. That halides compete with H(2)O(2) for the same redox intermediate is also emphasized by the fact that the polarographically measured catalase activity is influenced by halides, with bromide being more effective than chloride.  相似文献   

12.
NAD(P)H acts as a two-electron reductant in physiological, enzyme-controlled processes. Under nonenzymatic conditions, a couple of one-electron oxidants easily oxidize NADH to the NAD(.) radical. This radical reduces molecular oxygen to the superoxide radical (O-(2)) at a near to the diffusion-controlled rate, thereby subsequently forming hydrogen peroxide (H(2)O(2)). Because peroxynitrite can act as a one-electron oxidant, the reaction of NAD(P)H with both authentic peroxynitrite and the nitric oxide ((. )NO) and O-(2) releasing compound 3-morpholinosydnonimine N-ethylcarbamide (SIN-1) was studied. Authentic peroxynitrite oxidized NADH with an efficiency of approximately 25 and 8% in the absence and presence of bicarbonate/carbon dioxide (HCO(3)(-)/CO(2)), respectively. NADH reacted 5-100 times faster with peroxynitrite than do the known peroxynitrite scavengers glutathione, cysteine, and tryptophan. Furthermore, NADH was found to be highly effective in suppressing peroxynitrite-mediated nitration reactions even in the presence of HCO(3)(-)/CO(2). Reaction of NADH with authentic peroxynitrite resulted in the formation of NAD(+) and O-(2) and, thus, of H(2)O(2) with yields of about 3 and 10% relative to the added amounts of peroxynitrite and NADH, respectively. Peroxynitrite generated in situ from SIN-1 gave virtually the same results; however, two remarkable exceptions were recognized. First, the efficiency of NADH oxidation increased to 60-90% regardless of the presence of HCO(3)(-)/CO(2), along with an increase of H(2)O(2) formation to about 23 and 35% relative to the amounts of added SIN-1 and NADH. Second, and more interesting, the peroxynitrite scavenger glutathione (GSH) was needed in a 75-fold surplus to inhibit the SIN-1-dependent oxidation of NADH half-maximal in the presence of HCO(3)(-)/CO(2). Similar results were obtained with NADPH. Hence, peroxynitrite or radicals derived from it (such as, e.g. the bicarbonate radical or nitrogen dioxide) indeed oxidize NADH, leading to the formation of NAD(+) and, via O-(2), of H(2)O(2). When peroxynitrite is generated in situ in the presence of HCO(3)(-)/CO(2), i.e. under conditions mimicking the in vivo situation, NAD(P)H effectively competes with other known scavengers of peroxynitrite.  相似文献   

13.
A study was performed on the oxidizing degradation of 2-chloroaniline (used as a model pollutant in water) by photolysis (lambda = 254 nm). The change of spectrum and substrate concentration of treated solutions was measured spectrophotometrically as well as by HPLC. The yields of the degradation products (chloride ions, ammonium ions, formaldehyde, etc.) were studied as a function of UV-dose. Their initial quantum yields (Qi) were determined by specific analysis. It was shown that the substrate photolysis in the presence of N2O is most efficient, followed by degradation in media saturated with pure oxygen and air. A probable reaction mechanism for the photo-induced degradation of 2-ClA is presented.  相似文献   

14.
An electron-rich iron(III) porphyrin complex (meso-tetramesitylporphinato)iron(III) chloride [Fe(TMP)Cl], was found to catalyze the epoxidation of olefins by aqueous 30% H2O2 when the reaction was carried out in the presence of 5-chloro-1-methylimidazole (5-Cl-1-MeIm) in aprotic solvent. Epoxides were the predominant products with trace amounts of allylic oxidation products, indicating that Fenton-type oxidation reactions were not involved in the olefin epoxidation reactions. cis-Stilbene was stereospecifically oxidized to cis-stilbene oxide without giving isomerized trans-stilbene oxide product, demonstrating that neither hydroperoxy radical (HOO·) nor oxoiron(IV) porphyrin [(TMP)FeIV=O] was responsible for the olefin epoxidations. We also found that the reactivities of other iron(III) porphyrin complexes such as (meso-tetrakis(2,6-dichlorophenyl)porphinato)iron(III) chloride [Fe(TDCPP)Cl], (meso-tetrakis(2,6-difluorophenyl)porphinato)iron(III) chloride [Fe(TDFPP)Cl], and (meso-tetrakis(pentafluorophenyl)porphinato)iron(III) chloride [Fe(TPFPP)Cl] were significantly affected by the presence of the imidazole in the epoxidation of olefins by H2O2. These iron porphyrin complexes did not yield cyclohexene oxide in the epoxidation of cyclohexene by H2O2 in the absence of 5-Cl-1-MeIm in aprotic solvent; however, addition of 5-Cl-1-MeIm to the reaction solutions gave high yields of cyclohexene oxide with the formation of trace amounts of allylic oxidation products. We proposed, on the basis of the results of mechanistic studies, that the role of the imidazole is to decelerate the O–O bond cleavage of an iron(III) hydroperoxide porphyrin (or H2O2–iron(III) porphyrin adduct) and that the intermediate transfers its oxygen to olefins prior to the O–O bond cleavage.  相似文献   

15.
It has been proposed (S. Christen et al. Proc. Natl. Acad. Sci. USA 94, 3217-3222, 1997) that although alpha-tocopherol (alpha-TH) is an efficient antioxidant, the presence of gamma-tocopherol (gamma-TH) may be required to scavenge peroxynitrite-derived reactive nitrogen species. To investigate the reactions between alpha-TH, gamma-TH, and peroxynitrite, endogenous levels of both alpha-TH and gamma-TH were monitored when low-density lipoprotein was oxidized in the presence of the peroxynitrite generator 5-amino-3-(4-morpholinyl)-1, 2,3-oxadiazolium (SIN-1). SIN-1 oxidized alpha-TH while gamma-TH levels remained constant. The sparing of gamma-TH was also demonstrated when 1,2-dilauroyl-sn-glycero-3-phosphocholine liposomes containing alpha-TH and gamma-TH were incubated with either SIN-1 or peroxynitrite. Our data show that alpha-TH inhibits peroxynitrite-mediated gamma-TH nitration, i.e., 5-NO2-gamma-tocopherol formation. The rate constants for the reactions between both alpha-TH and gamma-TH with peroxynitrite suggest that the sparing of gamma-TH by alpha-TH does not occur by competitive scavenging, but may be due to the formation of a transient gamma-TH intermediate. Nitration of gamma-TH becomes significant only after alpha-TH levels have been depleted. We conclude alpha-TH alone is sufficient to remove any peroxynitrite-derived reactive nitrogen species, as the presence of alpha-TH attenuates nitration of both gamma-TH and tyrosine. The present results also indicate that a bolus addition of peroxynitrite or SIN-1 to liposomes containing gamma-TH forms 5-NO2-gamma-tocopherol in similar yields. This is in contrast to their reaction profile with tyrosine in aqueous solution. Under these conditions, SIN-1 does not form nitrotyrosine at detectable yields.  相似文献   

16.
HE (hydroethidine), a widely used fluorescent dye for detecting intracellular superoxide, undergoes specific oxidation and hydroxylation reactions. The reaction between HE and O2?- (superoxide radical) yields a diagnostic marker product, 2-hydroxyethidium. This is contrary to the popular notion that O2?- oxidizes HE to form ethidium. HE, however, undergoes a non-specific oxidation to form ethidium in the presence of other oxidants (hydroxyl radical, peroxynitrite and perferryl iron) and other dimeric products. The mitochondria-targeted HE analogue Mito-SOX? undergoes the same type of oxidative chemistry to form products similar to those formed from HE. On the basis of the oxidative chemical mechanism of HE and Mito-SOX?, we conclude that flurorescence microscopy or related techniques are not sufficient to measure the superoxide-specific hydroxylated products. HPLC methodologies are required to separate and identify these products. Peroxynitrite reacts rapidly and stoichiometrically with boronates to form specific products. Assays using fluorescent-based boronate probes will be more reliable for peroxynitrite determination than those using either dichlorodihydrofluorescein or dihydrorhodamine.  相似文献   

17.
The close association of the heme enzyme myeloperoxidase to phosphatidylserine epitopes on the surface of non-vital polymorphonuclear leukocytes (PMNs) and other apoptotic cells at inflammatory sites favours modifications of this phospholipid by myeloperoxidase products. As detected by matrix-assisted laser desorption/ionization time-of-flight (MALDI-TOF) mass spectrometry, ammonium ions inhibit in a concentration-dependent manner the hypochlorous acid-mediated formation of aldehyde and nitrile products from 1,2-dipalmitoyl-sn-glycero-3-phosphoserine (DPPS). Concomitantly, the formation of monochloramine (NH2Cl) raises with increasing NH4+ concentrations. A transchlorination from monochlorinated O-phospho-l-serine to NH4+ with the formation of NH2Cl occurs only when extraordinary high NH4+ concentrations are applied. Due to the low rate of 0.044 M− 1 s− 1 for this process, a transhalogenation reaction from transient chlorinated intermediates of the serine moiety to NH4+ can be ruled out as an important process contributing to the HOCl-mediated formation of NH2Cl. A significant formation of NH2Cl by myeloperoxidase interacting with DPPS in the presence of ammonium ions takes only place at acidic pH values around 5, a scenario that may occur in phagosomes of macrophages after the uptake of apoptotic PMNs.  相似文献   

18.
By the observation of chemically induced dynamic nuclear polarization in (15)N NMR spectroscopy it has been shown that nitration of N-acetyltyrosine, even under acidic conditions, is largely a radical process. In the alkaline reaction of tyrosine with peroxynitrite the main products are nitrite and nitrate, both produced by a radical pathway, and tyrosine nitration is a minor reaction. It is suggested that tyrosine catalyzes the production of NO(*)(2) and HO(*) from peroxynitrite.  相似文献   

19.
Peroxynitrite, a reactive nitrogen species generated from nitric oxide and superoxide anion radical, is an endogenous potential risk factor for human cancer. When 2′-deoxycytidine was incubated with peroxynitrite at neutral pH and 37 °C, the reaction was greatly enhanced by the addition of ammonium bromide. Both ammonium ion and bromide ion were required to exert the enhancing effect. In addition to ammonium ion, methylamine and dimethylamine exerted the enhancing effect in the presence of bromide ion. Two major products were identified as 5-hydroxy-2′-deoxycytidine and 5-bromo-2′-deoxycytidine. Hypochlorite solution and bromine water reacted with 2′-deoxycytidine generating 5-hydroxy-2′-deoxycytidine and 5-bromo-2′-deoxycytidine in the presence of ammonium bromide with the yields similar to those of the reaction of peroxynitrite with ammonium bromide. Fenton reaction of 2′-deoxycytidine was suppressed by the addition of ammonium bromide. Nitrogen dioxide gas did not react with 2′-deoxycytidine in the presence or the absence of ammonium bromide. These results suggest that in the presence of ammonium ion or amines, bromide ion interacts with peroxynitrous acid, which is a protonated form of peroxynitrite, but not with hydroxyl radical or nitrogen dioxide generated by homolysis of peroxynitrous acid, to form hypobromous acid. In the presence of ammonium ion or amines, bromide ion may play a role in enhancing the genotoxic effects of peroxynitrite in humans.  相似文献   

20.
Formation of peroxynitrite from NO and O-(*2) is considered an important trigger for cellular tyrosine nitration under pathophysiological conditions. However, this view has been questioned by a recent report indicating that NO and O-(*2) generated simultaneously from (Z)-1-(N-[3-aminopropyl]-N-[4-(3-aminopropylammonio)butyl]-amino) diazen-1-ium-1,2-diolate] (SPER/NO) and hypoxanthine/xanthine oxidase, respectively, exhibit much lower nitrating efficiency than authentic peroxynitrite (Pfeiffer, S. and Mayer, B. (1998) J. Biol. Chem. 273, 27280-27285). The present study extends those earlier findings to several alternative NO/O-(*2)-generating systems and provides evidence that the apparent lack of tyrosine nitration by NO/O-(*2) is due to a pronounced decrease of nitration efficiency at low steady-state concentrations of authentic peroxynitrite. The decrease in the yields of 3-nitrotyrosine was accompanied by an increase in the recovery of dityrosine, showing that dimerization of tyrosine radicals outcompetes the nitration reaction at low peroxynitrite concentrations. The observed inverse dependence on peroxynitrite concentration of dityrosine formation and tyrosine nitration is predicted by a kinetic model assuming that radical formation by peroxynitrous acid homolysis results in the generation of tyrosyl radicals that either dimerize to yield dityrosine or combine with (*)NO(2) radical to form 3-nitrotyrosine. The present results demonstrate that very high fluxes (>2 microM/s) of NO/O-(*2) are required to render peroxynitrite an efficient trigger of tyrosine nitration and that dityrosine is a major product of tyrosine modification caused by low steady-state concentrations of peroxynitrite.  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