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.
We investigated the effects of a cysteine residue on tyrosine nitration in several model peptides treated with myeloperoxidase (MPO), H(2)O(2), and nitrite anion (NO(2)(-)) and with horseradish peroxidase and H(2)O(2). Sequences of model peptides were acetyl-Tyr-Cys-amide (YC), acetyl-Tyr-Ala-Cys-amide (YAC), acetyl-Tyr-Ala-Ala-Cys-amide (YAAC), and acetyl-Tyr-Ala-Ala-Ala-Ala-Cys-amide (YAAAAC). Results indicate that nitration and oxidation products of tyrosyl residue in YC and other model peptides were barely detectable. A major product detected was the corresponding disulfide (e.g. YCysCysY). Spin trapping experiments with 5,5'-dimethyl-1-pyrroline N-oxide (DMPO) revealed thiyl adduct (e.g. DMPO-SCys-Tyr) formation from peptides (e.g. YC) treated with MPO/H(2)O(2) and MPO/H(2)O(2)/NO(2)(-). The steady-state concentrations of DMPO-thiyl adducts decreased with increasing chain length of model peptides. Blocking the sulfydryl group in YC with methylmethanethiosulfonate (that formed YCSSCH(3)) totally inhibited thiyl radical formation as did substitution of Tyr with Phe (i.e. FC) in the presence of MPO/H(2)O(2)/NO(2)(-). However, increased tyrosine nitration, tyrosine dimerization, and tyrosyl radical formation were detected in the MPO/H(2)O(2)/NO(2)(-)/YCSSCH(3) system. Increased formation of S-nitrosated YC (YCysNO) was detected in the MPO/H(2)O(2)/(*)NO system. We conclude that a rapid intramolecular electron transfer reaction between the tyrosyl radical and the Cys residue impedes tyrosine nitration and induces corresponding thiyl radical and nitrosocysteine product. Implications of this novel intramolecular electron transfer mechanism in protein nitration and nitrosation are discussed.  相似文献   

2.
In a series of heme and non-heme proteins the nitration of tyrosine residues was assessed by complete pronase digestion and subsequent HPLC-based separation of 3-nitrotyrosine. Bolus addition of peroxynitrite caused comparable nitration levels in all tested proteins. Nitration mainly depended on the total amount of tyrosine residues as well as on surface exposition. In contrast, when superoxide and nitrogen monoxide (NO) were generated at equal rates to yield low steady-state concentrations of peroxynitrite, metal catalysis seemed to play a dominant role in determining the sensitivity and selectivity of peroxynitrite-mediated tyrosine nitration in proteins. Especially, the heme-thiolate containing proteins cytochrome P450(BM-3) (wild type and F87Y variant) and prostacyclin synthase were nitrated with high efficacy. Nitration by co-generated NO/O(2)(-) was inhibited in the presence of superoxide dismutase. The NO source alone only yielded background nitration levels. Upon changing the NO/O(2)(-) ratio to an excess of NO, a decrease in nitration in agreement with trapping of peroxynitrite and derived radicals by NO was observed. These results clearly identify peroxynitrite as the nitrating species even at low steady-state concentrations and demonstrate that metal catalysis plays an important role in nitration of protein-bound tyrosine.  相似文献   

3.
Nitration of protein tyrosine residues to 3-nitrotyrosine (NO2Tyr) serves as both a marker and mediator of pathogenic reactions of nitric oxide (*NO), with peroxynitrite (ONOO-) and leukocyte peroxidase-derived nitrogen dioxide (*NO2) being proximal mediators of nitration reactions in vivo. Cytochrome c is a respiratory and apoptotic signaling heme protein localized exofacially on the inner mitochondrial membrane. We report herein a novel function for cytochrome c as a catalyst for nitrite (NO2-) and hydrogen peroxide (H2O2)-mediated nitration reactions. Cytochrome c catalyzes both self- and adjacent-molecule (hydroxyphenylacetic acid, Mn-superoxide dismutase) nitration via heme-dependent mechanisms involving tyrosyl radical and *NO2 production, as for phagocyte peroxidases. Although low molecular weight phenolic nitration yields were similar for cytochrome c and the proteolytic fragment of cytochrome c microperoxidase-11 (MPx-11), greater extents of protein nitration occurred when MPx-11 served as catalyst. Partial proteolysis of cytochrome c increased both the peroxidase and nitrating activities of cytochrome c. Extensive tyrosine nitration of Mn-superoxide dismutase occurred when exposed to either cytochrome c or MPx-11 in the presence of H2O2 and NO2-, with no apparent decrease in catalytic activity. These results reveal a post-translational tyrosine modification mechanism that is mediated by an abundant hemoprotein present in both mitochondrial and cytosolic compartments. The data also infer that the distribution of specific proteins capable of serving as potent catalysts of nitration can lend both spatial and molecular specificity to biomolecule nitration reactions.  相似文献   

4.
Factors determining the selectivity of protein tyrosine nitration.   总被引:9,自引:0,他引:9  
Tyrosine nitration is a covalent posttranslational protein modification derived from the reaction of proteins with nitrating agents. Protein nitration appears to be a selective process since not all tyrosine residues in proteins or all proteins are nitrated in vivo. To investigate factors that may determine the biological selectivity of protein tyrosine nitration, we developed an in vitro model consisting of three proteins with similar size but different three-dimensional structure and tyrosine content. Exposure of ribonuclease A to putative in vivo nitrating agents revealed preferential nitration of tyrosine residue Y(115). Tyrosine residue Y(23) and to a lesser extent residue Y(20) were preferentially nitrated in lysozyme, whereas tyrosine Y(102) was the only residue modified by nitration in phospholipase A(2). Tyrosine Y(115) was the residue modified by nitration after exposure of ribonuclease A to different nitrating agents: chemically synthesized peroxynitrite, nitric oxide, and superoxide generated by SIN-1 or myeloperoxidase (MPO)/H(2)O(2) plus nitrite (NO(-2)) in the presence of bicarbonate/CO(2). The nature of the nitrating agent determined in part the protein that would be predominantly modified by nitration in a mixture of all three proteins. Ribonuclease A was preferentially nitrated upon exposure to MPO/H(2)O(2)/NO(-2), whereas phospholipase A(2) was the primary target for nitration upon exposure to peroxynitrite. The data also suggest that the exposure of the aromatic ring to the surface of the protein, the location of the tyrosine on a loop structure, and its association with a neighboring negative charge are some of the factors determining the selectivity of tyrosine nitration in proteins.  相似文献   

5.
Radiation chemical experiments demonstrate that the reaction of tyrosyl radical (TyrO(.)) with (.)NO(2) yields 45 +/- 3% 3-nitrotyrosine and that a major product of the reaction of TyrO(.) with (.)NO is 3,3'-dityrosine. Radiolysis was used to generate (.)NO and O-(2) in the presence of tyrosine and bicarbonate at pH 7.5 +/- 0.1. The nitration yield was found to be dose rate-dependent, and the yield per radical produced by pulse radiolysis was identical to that obtained with authentic peroxynitrite. The proposed mechanism that accounts for the data is as follows: (i) In the presence of CO(2) the reaction of (.)NO with O-(2) yields 33% (.)NO(2) and CO-(3), where the latter reacts rapidly with tyrosine to form TyrO(.); (ii) The formation of 3-nitrotyrosine takes place via the reaction of (.)NO(2) with TyrO(.), which is the main process at high dose rates; and (iii) Under continuous generation of (.)NO and O-(2), the formation of 3-nitrotyrosine is strongly suppressed because of efficient scavenging of (.)NO(2) by tyrosine. The proposed model shows that the highest nitration yield is obtained for similar fluxes of (.)NO and O-(2) and is completely inhibited upon excess production of O-(2) because of efficient scavenging of TyrO(.) by O-(2). The biological implications of these findings are discussed.  相似文献   

6.
Kinetics of the reaction of peroxynitrite with ferric cytochrome c in the absence and presence of bicarbonate was studied. It was found that the heme iron in ferric cytochrome c does not react directly with peroxynitrite. The rates of the absorbance changes in the Soret region of cytochrome c spectrum caused by peroxynitrite or peroxynitrite/bicarbonate were the same as the rate of spontaneous isomerization of peroxynitrite or as the rate of the reaction of peroxynitrite with bicarbonate, respectively. This means that intermediate products of peroxynitrite decomposition, (.)OH/(.)NO(2) or, in the presence of bicarbonate, CO(3)(-)(.)/(.)NO(2), are the species responsible for the absorbance changes in the Soret band of cytochrome c. Modifications of the heme center of cytochrome c by radiolytically produced radicals, (.)OH, (.)NO(2) or CO(3)(-)(.), were also studied. The absorbance changes in the Soret band caused by radiolytically produced (.)OH or CO(3)(-)(.) were much more significant that those observed after peroxynitrite treatment, compared under similar concentrations of radicals. (.)NO(2) produced radiolytically did not interact with the heme center of cytochrome c. Cytochrome c exhibited an increased peroxidase-like activity after reaction with peroxynitrite as well as with radiolytically produced (.)OH, (.)NO(2) or CO(3)(-)(.) radicals. This means that modification of protein structure: oxidation of amino acids and/or tyrosine nitration, facilitates reaction of H(2)O(2) with the heme iron of cytochrome c, followed by reaction with the second substrate.  相似文献   

7.
Recent reports suggest that intramolecular electron transfer reactions can profoundly affect the site and specificity of tyrosyl nitration and oxidation in peptides and proteins. Here we investigated the effects of methionine on tyrosyl nitration and oxidation induced by myeloperoxidase (MPO), H2O2 and NO2 and peroxynitrite (ONOO) or ONOO and bicarbonate (HCO3) in model peptides, tyrosylmethionine (YM), tyrosylphenylalanine (YF) and tyrosine. Nitration and oxidation products of these peptides were analyzed by HPLC with UV/Vis and fluorescence detection, and mass spectrometry; radical intermediates were identified by electron paramagnetic resonance (EPR)-spin-trapping. We have previously shown (Zhang et al., J. Biol. Chem. 280 (2005) 40684-40698) that oxidation and nitration of tyrosyl residue was inhibited in tyrosylcysteine(YC)-type peptides as compared to free tyrosine. Here we show that methionine, another sulfur-containing amino acid, does not inhibit nitration and oxidation of a neighboring tyrosine residue in the presence of ONOO (or ONOOCO2) or MPO/H2O2/NO2 system. Nitration of tyrosyl residue in YM was actually stimulated under the conditions of in situ generation of ONOO (formed by reaction of superoxide with nitric oxide during SIN-1 decomposition), as compared to YF, YC and tyrosine. The dramatic variations in tyrosyl nitration profiles caused by methionine and cysteine residues have been attributed to differences in the direction of intramolecular electron transfer in these peptides. Further support for the interpretation was obtained by steady-state radiolysis and photolysis experiments. Potential implications of the intramolecular electron transfer mechanism in mediating selective nitration of protein tyrosyl groups are discussed.  相似文献   

8.
The biological relevance of tyrosine nitration is a subject of much interest, because extensive evidence supports formation of 3-nitrotyrosine in vivo under a variety of different pathological conditions. Several reagents are likely to be responsible for nitration in vivo, among others peroxynitrite and nitrite in the presence of H(2)O(2)/peroxidases. In this work we show that also metmyoglobin and methemoglobin can nitrate free tyrosine in the presence of nitrite and H(2)O(2). The results of these studies are simulated rather well by using a scheme that comprehends all the possible reactions that can take place in the system. Thus, a good understanding of the factors that determine the yields is achieved. Finally, we demonstrate that the system metMb/H(2)O(2)/NO(2)(-) can also lead to the nitration of tryptophan and produces, in particular, 6-, 4-, and 5-nitrotryptophan.  相似文献   

9.
Peroxynitrite, formed in a rapid reaction of nitric oxide (NO) and superoxide anion radical (O(2)), is thought to mediate protein tyrosine nitration in various inflammatory and infectious diseases. However, a recent in vitro study indicated that peroxynitrite exhibits poor nitrating efficiency at biologically relevant steady-state concentrations (Pfeiffer, S., Schmidt, K., and Mayer, B. (2000) J. Biol. Chem. 275, 6346-6352). To investigate the molecular mechanism of protein tyrosine nitration in intact cells, murine RAW 264.7 macrophages were activated with immunological stimuli, causing inducible NO synthase expression (interferon-gamma in combination with either lipopolysaccharide or zymosan A), followed by the determination of protein-bound 3-nitrotyrosine levels and release of potential triggers of nitration (NO, O(2)*, H(2)O(2), peroxynitrite, and nitrite). Levels of 3-nitrotyrosine started to increase at 16-18 h and exhibited a maximum at 20-24 h post-stimulation. Formation of O(2) was maximal at 1-5 h and decreased to base line 5 h after stimulation. Release of NO peaked at approximately 6 and approximately 9 h after stimulation with interferon-gamma/lipopolysaccharide and interferon-gamma/zymosan A, respectively, followed by a rapid decline to base line within the next 4 h. NO formation resulted in accumulation of nitrite, which leveled off at about 50 microm 15 h post-stimulation. Significant release of peroxynitrite was detectable only upon treatment of cytokine-activated cells with phorbol 12-myristate-13-acetate, which led to a 2.2-fold increase in dihydrorhodamine oxidation without significantly increasing the levels of 3-nitrotyrosine. Tyrosine nitration was inhibited by azide and catalase and mimicked by incubation of unstimulated cells with nitrite. Together with the striking discrepancy in the time course of NO/O(2) release versus 3-nitrotyrosine formation, these results suggest that protein tyrosine nitration in activated macrophages is caused by a nitrite-dependent peroxidase reaction rather than peroxynitrite.  相似文献   

10.
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   

11.
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   

12.
Eosinophil peroxidase (EPO) has been implicated in promoting oxidative tissue injury in conditions ranging from asthma and other allergic inflammatory disorders to cancer and parasitic/helminthic infections. Studies thus far on this unique peroxidase have primarily focused on its unusual substrate preference for bromide (Br(-)) and the pseudohalide thiocyanate (SCN(-)) forming potent hypohalous acids as cytotoxic oxidants. However, the ability of EPO to generate reactive nitrogen species has not yet been reported. We now demonstrate that EPO readily uses nitrite (NO(2)(-)), a major end-product of nitric oxide ((.)NO) metabolism, as substrate to generate a reactive intermediate that nitrates protein tyrosyl residues in high yield. EPO-catalyzed nitration of tyrosine occurred more readily than bromination at neutral pH, plasma levels of halides, and pathophysiologically relevant concentrations of NO(2)(-). Furthermore, EPO was significantly more effective than MPO at promoting tyrosine nitration in the presence of plasma levels of halides. Whereas recent studies suggest that MPO can also promote protein nitration through indirect oxidation of NO(2)(-) with HOCl, we found no evidence that EPO can indirectly mediate protein nitration by a similar reaction between HOBr and NO(2)(-). EPO-dependent nitration of tyrosine was modulated over a physiologically relevant range of SCN(-) concentrations and was accompanied by formation of tyrosyl radical addition products (e.g. o,o'-dityrosine, pulcherosine, trityrosine). The potential role of specific antioxidants and nucleophilic scavengers on yields of tyrosine nitration and bromination by EPO are examined. Thus, EPO may contribute to nitrotyrosine formation in inflammatory conditions characterized by recruitment and activation of eosinophils.  相似文献   

13.
H Zhang  J Joseph  J Feix  N Hogg  B Kalyanaraman 《Biochemistry》2001,40(25):7675-7686
It has been reported that peroxynitrite will initiate both oxidation and nitration of tyrosine, forming dityrosine and nitrotyrosine, respectively. We compared peroxynitrite-dependent oxidation and nitration of a hydrophobic tyrosine analogue in membranes and tyrosine in aqueous solution. Reactions were carried out in the presence of either bolus addition or slow infusion of peroxynitrite, and also using the simultaneous generation of superoxide and nitric oxide. Results indicate that the level of nitration of the hydrophobic tyrosyl probe located in a lipid bilayer was significantly greater than its level of oxidation to the corresponding dimer. During slow infusion of peroxynitrite, the level of nitration of the membrane-incorporated tyrosyl probe was greater than that of tyrosine in aqueous solution. Evidence for hydroxyl radical formation from decomposition of peroxynitrite in a dimethylformamide/water mixture was obtained by electron spin resonance spin trapping. Mechanisms for nitration of the tyrosyl probe in the membrane are discussed. We conclude that nitration but not oxidation of a tyrosyl probe by peroxynitrite is a predominant reaction in the membrane. Thus, the local environment of target tyrosine residues is an important factor governing its propensity to undergo nitration in the presence of peroxynitrite. This work provides a new perspective on selective nitration of membrane-incorporated tyrosine analogues.  相似文献   

14.
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   

15.
Nitration of unsaturated fatty acids is a (patho)physiologically important pathway of lipid modification induced by nitric oxide-derived species. We report herein on the unexpected chain rearrangement undergone by (13S,9Z,11E)-13-hydroxyoctadeca-9,11-dienoic acid (1), a linoleic acid metabolite, when exposed to nitrating agents of biological relevance. At pH 7.4 and at room temperature, reaction of 1 with peroxynitrite (ONOO-) as well as Fe2+-EDTA/H2O2/NO2- and horseradish peroxidase/H2O2/NO2- led to the formation of two nitration products, which could be isolated as the methyl esters and were identified as diastereoisomeric methyl (12S)-10,11-epoxy-12-hydroxy-9-nitromethylheptadecanoates by extensive 1H, 13C, 15N NMR and MS analysis.  相似文献   

16.
Nitrotyrosine (NO(2)Tyr) formation is a hallmark of acute and chronic inflammation and has been detected in a wide variety of human pathologies. However, the mechanisms responsible for this posttranslational protein modification remain elusive. While NO(2)Tyr has been considered a marker of peroxynitrite (ONOO(-)) formation previously, there is growing evidence that heme-protein peroxidase activity, in particular neutrophil-derived myeloperoxidase (MPO), significantly contributes to NO(2)Tyr formation in vivo via the oxidation of nitrite (NO(2)(-)) to nitrogen dioxide (.NO(2)). Coronary arteries from a patient with coronary artery disease, liver and lung tissues from a sickle cell disease patient, and an open lung biopsy from a lung transplant patient undergoing rejection were analyzed immunohistochemically to map relative tissue distributions of MPO and NO(2)Tyr. MPO immunodistribution was concentrated along the subendothelium in coronary tissue and hepatic veins as well as in the alveolar epithelial compartment of lung tissue from patients with sickle cell disease or acute rejection. MPO immunoreactivity strongly colocalized with NO(2)Tyr formation, which was similarly distributed in the subendothelial and epithelial regions of these tissues. The extracellular matrix protein fibronectin (FN), previously identified as a primary site of MPO association in vascular inflammatory reactions, proved to be a major target protein for tyrosine nitration, with a strong colocalization of MPO, NO(2)Tyr, and tissue FN occurring. Finally, lung tissue from MPO(-/-) mice, having tissue inflammatory responses stimulated by intraperitoneal zymosan administration, revealed less subendothelial NO(2)Tyr immunoreactivity than tissue from wild-type mice, confirming the significant role that MPO plays in catalyzing tissue nitration reactions. These observations reveal that (i) sequestration of neutrophil-derived MPO in vascular endothelial and alveolar epithelial compartments is an important aspect of MPO distribution and action in vivo, (ii) MPO-catalyzed NO(2)Tyr formation occurs in diverse vascular and pulmonary inflammatory pathologies, and (iii) extracellular matrix FN is an important target of tyrosine nitration in these inflammatory processes.  相似文献   

17.
Jang B  Han S 《Biochimie》2006,88(1):53-58
Nitration of tyrosine residues is taken as evidence for intracellular formation of peroxynitrite. Cytochrome c (cyt c) can be nitrated by peroxynitrite and nitrated cyt c has been observed in cells and tissues under stress conditions. Here we studied the biochemical properties of nitrated cyt c in order to understand its potential roles in nitrative stress. Nitration of cyt c resulted in disruption of the heme-methionine bond and rapid binding to cyanide. Equilibrium unfolding by guanidine hydrochloride showed that cyt c was slightly destabilized upon nitration but the unfolding transition of nitrated cyt c was highly cooperative indicating that the overall folding was largely preserved. Nitrated cyt c could not be reduced by superoxide and did not support electron transfer between ascorbate and cyt c oxidase. Nitration of cyt c resulted in a tremendous increase in peroxidase activity so that nitrated cyt c rapidly oxidized dihydrodichlorofluorescein even in the presence of a high concentration of glutathione. Enhanced peroxidase activity of nitrated cyt c was responsible for H2O2-induced oxidation of phospholipid membranes and H2O2/NO2--mediated nitration of other proteins. These results suggest that nitration of cyt c by peroxynitrite may exacerbate oxidative damage to mitochondrial proteins and membranes.  相似文献   

18.
Hemodynamics, specifically, fluid shear stress, modulates the focal nature of atherogenesis. Superoxide anion (O2(-.)) reacts with nitric oxide (.NO) at a rapid diffusion-limited rate to form peroxynitrite (O2(-.) + .NO-->ONOO(-)). Immunohistostaining of human coronary arterial bifurcations or curvatures, where OSS develops, revealed the presence of nitrotyrosine staining, a fingerprint of peroxynitrite; whereas in straight segments, where PSS occurs, nitrotyrosine was absent. We examined vascular nitrative stress in models of oscillatory (OSS) and pulsatile shear stress (PSS). Bovine aortic endothelial cells (BAEC) were exposed to fluid shear stress that simulates arterial blood flow: (1) PSS at a mean shear stress (tau(ave)) of 23 dyn cm(-2) and a temporal gradient (partial differential(tau)/partial differential(t)) at 71 dyn cm(-2) s(-1), and (2) OSS at tau(ave) = 0.02 dyn cm(- 2) and partial differential(tau)/partial differential(t) = +/- 3.0 dyn cm(-2) s(-1) at a frequency of 1 Hz. OSS significantly up-regulated one of the NADPH oxidase subunits (NOx4) expression accompanied with an increase in O2(-.) production. In contrast, PSS up-regulated eNOS expression accompanied with .NO production (total NO(2)(-) and NO(3)(-)). To demonstrate that O2(-.) and .NO are implicated in ONOO(-) formation, we added low-density lipoprotein cholesterol (LDL) to the medium in which BAEC were exposed to the above flow conditions. The medium was analyzed for LDL apo-B-100 nitrotyrosine by liquid chromatography electrospray ionization tandem mass spectrometry (LC/ESI/MS/MS). OSS induced higher levels of 3-nitrotyrosine, dityrosine, and o-hydroxyphenylalanine compared with PSS. In the presence of ONOO(-), specific apo-B-100 tyrosine residues underwent nitration in the alpha and beta helices: alpha-1 (Tyr(144)), alpha-2 (Tyr(2524)), beta-2 (Tyr(3295)), alpha-3 (Tyr(4116)), and beta-2 (Tyr(4211)). Hence, the characteristics of shear stress in the arterial bifurcations influenced the relative production of O2(-.) and .NO with an implication for ONOO(-) formation as evidenced by LDL protein nitration.  相似文献   

19.
Excess superoxide (O(2)(-)) and nitric oxide (NO) forms peroxynitrite (ONOO(-)) during cardiac ischemia reperfusion (IR) injury, which in turn induces protein tyrosine nitration (tyr-N). Mitochondria are both a source of and target for ONOO(-). Our aim was to identify specific mitochondrial proteins that display enhanced tyr-N after cardiac IR injury, and to explore whether inhibiting O(2)(-)/ONOO(-) during IR decreases mitochondrial protein tyr-N and consequently improves cardiac function. We show here that IR increased tyr-N of 35 and 15kDa mitochondrial proteins using Western blot analysis with 3-nitrotyrosine antibody. Immunoprecipitation (IP) followed by LC-MS/MS identified 13 protein candidates for tyr-N. IP and Western blot identified and confirmed that the 35kDa tyr-N protein is the voltage-dependent anion channel (VDAC). Tyr-N of native cardiac VDAC with IR was verified on recombinant (r) VDAC with exogenous ONOO(-). We also found that ONOO(-) directly enhanced rVDAC channel activity, and rVDAC tyr-N induced by ONOO(-) formed oligomers. Resveratrol (RES), a scavenger of O(2)(-)/ONOO(-), reduced the tyr-N levels of both native and recombinant VDAC, while L-NAME, which inhibits NO generation, only reduced tyr-N levels of native VDAC. O(2)(-) and ONOO(-) levels were reduced in perfused hearts during IR by RES and L-NAME and this was accompanied by improved cardiac function. These results identify tyr-N of VDAC and show that reducing ONOO(-) during cardiac IR injury can attenuate tyr-N of VDAC and improve cardiac function.  相似文献   

20.
In this study, we investigated the effects of various nitrogen oxide (NO(x)) species on the extent of prostaglandin H(2) synthase-1 (PGHS-1) nitration in purified protein and in vascular smooth muscle cells. We also examined PGHS-1 activity under these conditions and found the degree of nitration to correlate inversely with enzyme activity. In addition, since NO(x) species are thought to invoke damage during the pathogenesis of atherosclerosis, we examined human atheromatous tissue for PGHS-1 nitration. Both peroxynitrite and tetranitromethane induced Tyr nitration of purified PGHS-1, whereas 1-hydroxy-2-oxo-3-(N-methyl-aminopropyl)-3-methyl-1-triazene (NOC-7; a nitric oxide-releasing compound) did not. Smooth muscle cells treated with peroxynitrite showed PGHS-1 nitration. The extent of nitration by specific NO(x) species was determined by electrospray ionization mass spectrometry. Tetranitromethane was more effective than peroxynitrite, NOC-7, and nitrogen dioxide at nitrating a tyrosine-containing peptide (12%, 5%, 1%, and <1% nitration, respectively). Nitrogen dioxide and, to a lesser extent, peroxynitrite, induced dityrosine formation. Using UV/Vis spectroscopy, it was estimated that the reaction of PGHS-1 with excess peroxynitrite yielded two nitrated tyrosines/PGHS-1 subunit. Finally, atherosclerotic tissue obtained from endarterectomy patients was shown to contain nitrated PGHS-1. Thus, prolonged exposure to elevated levels of peroxynitrite may cause oxidative damage through tyrosine nitration.  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