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 925 毫秒
1.
Simultaneously produced superoxide/nitric oxide radicals (O2*-/NO*) could form peroxynitrite (OONO-) which has been found to cause atherogenic, i.e. oxidative modification of LDL. Aromatic hydroxylation and nitration of the aspirin metabolite salicylate by OONO- has been reported. Therefore we tested if salicylate may be able to protect LDL from oxidation by O2*-/NO* by scavenging the OONO reactive decomposition products. When LDL was exposed to simultaneously produced O2*-/NO* using the sydnonimine SIN-1, salicylate exerted an inhibitory effect on LDL oxidation as measured by TBARS and lipid hydroperoxide formation and alteration in electrophoretic mobility of LDL. The cytotoxic effect of SIN-1 pre-oxidised LDL to endothelial cells was also diminished when salicylate was present during SIN-1 treatment of LDL. Spectrophotometric analysis revealed that salicylate was converted to dihydroxybenzoic acid (DHBA) derivatives in the presence of SIN-1. 2,3- and 2,5-DHBA were even more effective to protect LDL from oxidation by O2*-/NO*. Because O2*-/NO* can occur in vivo, the results may indicate that salicylate could act as an efficacious inhibitor of O2*-/NO* initiated atherogenic LDL modification, thus further supporting the rationale of aspirin medication regarding cardiovascular diseases.  相似文献   

2.
Tyrosine nitration is a posttranslational modification observed in many pathologic states that can be associated with peroxynitrite (ONOO(-)) formation. However, in vitro, peroxynitrite-dependent tyrosine nitration is inhibited when its precursors, superoxide (O(2)*(-)) and nitric oxide ((*)NO), are formed at ratios (O(2)*(-)/(*)NO) different from one, severely questioning the use of 3-nitrotyrosine as a biomarker of peroxynitrite-mediated oxidations. We herein hypothesize that in biological systems the presence of superoxide dismutase (SOD) and the facile transmembrane diffusion of (*)NO preclude accumulation of O(2)*(-) and (*)NO radicals under flux ratios different from one, preventing the secondary reactions that result in the inhibition of 3-nitrotyrosine formation. Using an array of reactions and kinetic constants, computer-assisted simulations were performed in order to assess the flux of 3-nitrotyrosine formation (J(NO(2(-))Y)) during exposure to simultaneous fluxes of superoxide (J(O(2)*(-))) and nitric oxide (J((*)NO)), varying the radical flux ratios (J(O(2)*(-))/ J((*)NO)), in the presence of carbon dioxide. With a basic set of reactions, J(NO(2(-))Y) as a function of radical flux ratios rendered a bell-shape profile, in complete agreement with previous reports. However, when superoxide dismutation by SOD and (*)NO decay due to diffusion out of the compartment were incorporated in the model, a quite different profile of J(NO(2(-))Y) as a function of the radical flux ratio was obtained: despite the fact that nitration yields were much lower, the bell-shape profile was lost and the extent of tyrosine nitration was responsive to increases in either O(2)*(-) or (*)NO, in agreement with in vivo observations. Thus, the model presented herein serves to reconcile the in vitro and in vivo evidence on the role of peroxynitrite in promoting tyrosine nitration.  相似文献   

3.
After enzymic biotransformation, molsidomine (MO) acts via the metabolite 3-morpholinosydnonimine (SIN-1) through spontaneous liberation of nitric oxide (NO) and superoxide (O(2)(.-)). The aim of this study was to compare the effects of MO and its active metabolite SIN-1 on the redox status of rat erythrocytes and reticulocytes. Rat erythrocyte as well as reticulocyte-rich red blood cell (RBC) suspensions were aerobically incubated (2 h, 37 degrees C) without (control) or in the presence of different concentrations of MO or SIN-1. In rat erythrocytes, biotransformation of MO resulted in the production of NO and nitroxyl (NO(-)). Endogenous superoxide anion (O(2)(.-)) participated in peroxynitrite generation. SIN-1 simultaneously liberated NO and O(2)(.-), which formed peroxynitrite (at least in part), but the liberated NO predominantly reacted with haemoglobin, forming methaemoglobin in erythrocytes. In reticulocytes, MO and SIN-1 caused an increase in the levels of both nitrite and 3-nitrotyrosine (an indicator of peroxynitrite), whereas they decreased the level of O(2)(.-). In reticulocytes, MO was metabolized into SIN-1 which led to the generation of NO, which reacted with O(2)(.-) (endogenous or exogenous) forming reactive nitrogen species. In conclusion, there are two metabolic pathways for MO biotransformation: one causing NO and NO(-) generation predominantly in erythrocytes and the other, via SIN-1 metabolism, in reticulocytes. The main difference between the action of MO and SIN-1 was that the latter caused oxidative damage in RBCs.  相似文献   

4.
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   

5.
We have previously demonstrated that dopaminergic neurons in midbrain-striatum slice co-cultures are more resistant to NMDA cytotoxicity than the same neuronal population in single midbrain slice cultures. Here, we show that dopaminergic neurons in midbrain-striatum co-cultures also exhibit resistance to the cytotoxicity of nitric oxide donors, 2,2'-(hydroxynitrosohydrazono)bis-ethanamine (NOC-18) and 3-morpholinosydnonimine (SIN-1). The cytotoxicity of NMDA (30 microM) in single cultures was significantly attenuated by the nitric oxide synthase (NOS) inhibitor N(omega)-nitro-L-arginine (100 microM), whereas the toxicity in co-cultures was not. The levels of tyrosine residue nitration of tyrosine hydroxylase, a hallmark of the occurence of peroxynitrite anion in dopaminergic neurons, were lower in co-cultures than those in single cultures. Single cultures and co-cultures did not show appreciable differences in the number or distribution of NOS-containing neurons as assessed by NADPH diaphorase histochemistry. On the other hand, midbrain slices cultured with striatal slices showed higher levels of superoxide dismutase (SOD) activity as well as increased protein levels of Cu,Zn-SOD, than midbrain slices cultured alone. These results suggested that the generation of NO is involved in NMDA cytotoxicity on dopaminergic neurons, and that increased activity of SOD in co-cultures renders dopaminergic neurons resistant to NMDA cytotoxicity by preventing the formation of peroxynitrite.  相似文献   

6.
Guanidinosuccinic acid (GSA) is noted for its nitric oxide (NO) mimicking actions such as vasodilatation and activation of the N-methyl-D-aspartate (NMDA) receptor. We have reported that GSA is the product of argininosuccinate (ASA) and some reactive oxygen species, mainly the hydroxyl radical. We tested for GSA synthesis in the presence of NO donors. ASA (1 mM) was incubated with NOR-2, NOC-7 or 3-morpholinosydomine hydrochloride (SIN-1) at 37°C. GSA was determined by HPLC using a cationic resin for separation and phenanthrenequinone as an indicator. Neither NOR-2 or NOC-7 formed GSA. SIN-1, on the other hand, generates NO and the superoxide anion which, in turn, generated peroxynitrite which was then converted to the hydroxyl radical. Incubation of ASA with SIN-1 leads, via this route, to GSA. When ASA was incubated with 1 mM SIN-1, the amount of GSA produced depended on the incubation time and the concentration of ASA. Among the tested SIN-1 concentrations, from 0.5 to 5 mM, GSA synthesis was maximum at 0.5 mM and decreased with increasing concentrations of SIN-1. Carboxy-PTIO, a NO scavenger, completely inhibited GSA synthesis. SOD, a superoxide scavenger, decreased GSA synthesis by 20%, and catalase inhibited GSA synthesis only by 12%; DMSO, a hydroxyl radical scavenger completely inhibited GSA synthesis in the presence of SIN-1. These data suggest that the hydroxyl radical derived from a combination of NO and the superoxide anion generates GSA, a stable NO mimic. Meanwhile, synthesis of GSA by NO produces reactive oxygen and activates the NMDA receptor that generates NO from GSA, suggesting a positive feed back mechanism.  相似文献   

7.
To discover the effects of nitric oxide (NO) and peroxynitrite on Uronema marinum (a ciliate responsible for systemic scuticociliatosis in cultured olive flounder Paralichthys olivaceus), the dose-dependent inhibitory effect of NO donors, S-nitroso-N-acetylpenicillamine (SNAP) and 3-morpholinosydnonimine (SIN-1) on the proliferation and survival of U. marinum was investigated. The inhibitory effects of exogenous superoxide dismutase (SOD) and catalase on the toxicity of SIN-1 were also investigated. After 24 h of incubation in the presence of 0.2 mM SNAP, the number of ciliates was not statistically different from that of the controls, whereas incubation in the presence of 0.5 mM SNAP reduced the number of parasites significantly to 59.1% of controls. Concentrations of SNAP higher than 0.5 mM resulted in greater reductions in the number of ciliates, but levels of generated NO far exceeded physiological ranges. The number of viable ciliates incubated for 24 h with 0.2 mM SIN-1 was reduced significantly to 25.0%, and all ciliates were killed by incubation in concentrations above 0.5 mM SIN-1. Although SOD decreased the toxic effect of SIN-1 on U. marinum, protection was not complete and did not improve after increasing the SOD concentration from 50 to 400 U ml(-1). Addition of catalase ranging from 500 to 10000 U ml(-1) completely protected U. marinum from SIN-1 toxicity. Ciliates exposed to catalase alone or catalase plus SIN-1 showed significantly higher and dose-dependent proliferation rates compared to controls. Addition of haemoglobin, ranging from 0.5 to 2.0 mg ml(-1), also protected U. marinum from SIN-1 toxicity, and increased the proliferation rate dose-dependently. In conclusion, resistance of U. marinum to oxidative and nitrative stress may allow this pathogen to withstand the NO- and oxygen-radical-dependent killing mechanisms of phagocytic cells.  相似文献   

8.
Independently, superoxide (O2-) and nitric oxide (NO) are biologically important signaling molecules. When co-generated, these radicals react rapidly to form powerful oxidizing and nitrating intermediates. Although this reaction was once thought to be solely cytotoxic, herein we demonstrate using MCF7, macrophage, and endothelial cells that when nanomolar levels of NO and O2- were produced concomitantly, the effective NO concentration was established by the relative fluxes of these two radicals. Differential regulation of sGC, pERK, HIF-1alpha, and p53 were used as biological dosimeters for NO concentration. Introduction of intracellular- or extracellular-generated O2- during NO generation resulted in a concomitant increase in oxidative intermediates with a decrease in steady-state NO concentrations and a proportional reduction in the levels of sGC, ERK, HIF-1alpha, and p53 regulation. NO responses were restored by addition of SOD. The intermediates formed from the reactions of NO with O2- were non-toxic, did not form 3-nitrotyrosine, nor did they elicit any signal transduction responses. H2O2 in bolus or generated from the dismutation of O2- by SOD, was cytotoxic at high concentrations and activated p53 independent of NO. This effect was completely inhibited by catalase, suppressed by NO, and exacerbated by intracellular catalase inhibition. We conclude that the reaction of O2- with NO is an important regulatory mechanism, which modulates signaling pathways by limiting steady-state levels of NO and preventing H2O2 formation from O2-.  相似文献   

9.
Virulent strains of Photobacterium damselae subsp. piscicida (Pdp) were grown in media with or without glucose supplementation (to enhance polysaccharide capsule formation) and the bactericidal action of nitric oxide (NO) and peroxynitrites was evaluated in a cell-free assay. Treatment with the NO-donor S-nitroso-acetyl-penicillamine (SNAP) induced a dose-and time-dependent decrease in Pdp survival. This effect was greater for strains grown without glucose supplementation (C forms) than for their counterparts grown with glucose supplementation (C(+) forms). Addition of superoxide anion (O2(-)) generating systems (Xanthine/Xanthine oxidase, glucose/glucose oxidase) to the culture media further enhanced the bactericidal effect of NO. A similar bactericidal effect, with the same pattern of sensitivity, was observed when C+ and C forms of the bacteria were treated with 3-morpholino-sydonimide hydrochloride (SIN-1), a compound which simultaneously generates NO and O2(-). Addition of superoxide dismutase (SOD) or SOD plus catalase (CAT) did not fully reverse the toxic action of SIN-1 and the bactericidal effect was similar for both C and C(+) forms suggesting that while NO alone is sufficient to cause damage in all strains of the pathogen tested, growth in glucose supplemented medium enhanced protection to reactive oxygen intermediates rather than NO.  相似文献   

10.
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   

11.
As a generator of peroxynitrite (ONOO), 3-morpholinosydnonimine (SIN-1) is widely used in the study of oxidative/nitrosative stress in cultured cells, although controversy exists regarding active species responsible for cytotoxicity. In this study, we report that unstable thiol-sensitive substances, generated from the reaction of SIN-1 with components in culture medium, play a crucial role in SIN-1 cytotoxicity in PC12 cells. Exposure of cells to culture medium obtained after almost complete SIN-1 decomposition at 37 °C for 2 h demonstrated almost the same degree of cytotoxicity as did fresh SIN-1. The cytotoxicity of SIN-1-decomposed medium largely depended on serum, decayed with time, and could be completely abolished by the addition of thiols. Degradation of synthetic ONOO in the culture medium did not reproduce the unstable cytotoxicity. The presence of superoxide dismutase (SOD) during SIN-1 decomposition prevented the formation of the cytotoxic substances, whereas SOD had no protection against the cytotoxicity itself, suggesting a crucial role of simultaneously generated superoxide and nitric oxide in the formation of the toxicants, but not in their cytotoxic action. The cytotoxicity of fresh SIN-1 is dramatically suppressed in a basal medium (Hanks balanced salt), suggesting that the cytotoxicity of fresh SIN-1 also requires components of culture medium. These results suggest that SIN-1 cytotoxicity in PC12 cells is mediated via the generation of cytotoxic substances in the medium during its decomposition.  相似文献   

12.
Using oxidant-induced hyperpermeability of monolayers of intestinal (Caco-2) cells as a model for the pathophysiology of inflammatory bowel disease (IBD), we previously showed that oxidative injury to the F-actin cytoskeleton is necessary for the disruption of monolayer barrier integrity. We hypothesized that this cytoskeletal damage is caused by upregulation of an inducible nitric oxide (NO) synthase (iNOS)-driven pathway that overproduces reactive nitrogen metabolites (RNMs) such as NO and peroxynitrite (OONO(-)), which cause actin nitration and disassembly. Monolayers were exposed to H(2)O(2) or to RNMs with and without pretreatment with antioxidants or iNOS inhibitors. H(2)O(2) concentrations that disassembled and/or disrupted the F-actin cytoskeleton and barrier integrity also caused rapid iNOS activation, NO overproduction, and actin nitration. Added OONO(-) mimicked H(2)O(2); iNOS inhibitors and RNM scavengers were protective. Our results show that oxidant-induced F-actin and intestinal barrier disruption are caused by rapid iNOS upregulation that further increases oxidant levels; a similar positive feedback mechanism may underlie the episodic recurrence of the acute IBD attack. Confirming these mechanisms in vivo would provide a rationale for developing novel anti-RNM therapies for IBD.  相似文献   

13.
Peroxynitrite, formed by nitric oxide (NO) and superoxide, can alter protein function by nitrating amino acids such as tyrosine, cysteine, trytophan, or methionine. Inducible nitric oxide synthase (Type 2 NOS or iNOS) converts arginine to citrulline, releasing NO. We hypothesized that peroxynitrite could function as a negative feedback modulator of NO production by nitration of iNOS. Confluent cultures of the murine lung epithelial cell line, LA-4 were stimulated with cytokines to express iNOS, peroxynitrite was added, and the flasks sealed. After 3 h, NO in the headspace above the culture was sampled. Peroxynitrite caused a concentration-dependent decrease in NO. Similar results were obtained when 3-morpholinosydnonimine (SIN-1), a peroxynitrite generator, was added to the flasks. PAPA-NONOate, the NO generator, did not affect the headspace NO. Nitration of the iNOS was confirmed by detection of 3-nitrotyrosine by Western blotting. These data suggest a mechanism for inhibition of NO synthesis at inflammatory sites where iNOS, NO, and superoxide would be expected.  相似文献   

14.
We studied how the nitric oxide (NO*) donor 3-morpholinosydnonimine (SIN-1) alters the response to beta-adrenergic stimulation in cardiac rat myocytes. We found that SIN-1 decreases the positive inotropic effect of isoproterenol (Iso) and decreases the extent of both cell shortening and Ca2+ transient. These effects of SIN-1 were associated with an increased intracellular concentration of cGMP, a decreased intracellular concentration of cAMP, and a reduction in the levels of phosphorylation of phospholamban (PLB) and troponin I (TnI). The guanylyl cyclase inhibitor 1H-8-bromo-1,2,4-oxadiazolo (3,4-d)benz(b)(1,4)oxazin-1-one (ODQ) was not able to prevent the SIN-1-induced reduction of phosphorylation levels of PLB and TnI. However, the effects of SIN-1 were abolished in the presence of superoxide dismutase (SOD) or SOD and catalase. These data suggest that, in the presence of Iso, NO-related congeners, rather than NO*, are responsible for SIN-1 effects. Our results provide new insights into the mechanism by which SIN-1 alters the positive inotropic effects of beta-adrenergic stimulation.  相似文献   

15.
Nitroxyl anion (NO-), and/or its conjugate acid, HNO, may be formed in the cellular milieu by several routes under both physiological and pathophysiological conditions. Since experimental evidence suggests that certain reactive nitrogen oxide species can contribute significantly to cerebral ischemic injury, we investigated the neurotoxic potential of HNO/NO- using Angeli's salt (AS), a spontaneous HNO/NO(-)-generating compound. Exposure to AS resulted in a time- and concentration-dependent increase in neural cell death that progressed markedly following the initial exposure. Coadministration of the donor with Tempol (1 mM), a one-electron oxidant that converts NO- to NO, prevented its toxic effect, as did the concomitant addition of Fe(III)TPPS. Media containing various chelators, catalase, Cu/Zn superoxide dismutase, or carboxy-PTIO did not ameliorate AS-mediated neurotoxicity, ruling out the involvement of transition metal complexes, H2O2, O2-, and NO, respectively. A concentration-dependent increase in supernatant protein 3-nitrotyrosine immunoreactivity was observed when cultures were exposed to AS under aerobic conditions, an effect lost in the absence of oxygen. A bell-shaped curve for augmented AS-mediated nitration was observed with increasing Fe(III)TPPS concentration, which contrasted with its linear effect on abating cytotoxicity. Finally, addition of glutamate receptor antagonists, MK-801 (10 microM) and CNQX (30 microM) to the cultures abrogated toxicity when given during, but not following, AS exposure; as did pretreatment with the exocytosis inhibitor, tetanus toxin (300 ng/ml). Taken together, our data suggest that under aerobic conditions, AS toxicity is initiated via HNO/NO- but progresses via secondary excitotoxicity.  相似文献   

16.
Exposure of cerebellar granule neurones in 25 mm KCl HEPES-containing Locke's buffer (pH 7.4) to 50-100 microm SIN-1 during 2 h decreased the steady-state free cytosolic Ca2+ concentration ([Ca2+]i) from 168 +/- 33 nm to 60 +/- 10 nm, whereas exposure to > or = 0.3 mm SIN-1 produced biphasic kinetics: (i) decrease of [Ca2+]i during the first 30 min, reaching a limiting value of 75 +/- 10 nm (due to inactivation of L-type Ca2+ channels) and (ii) a delayed increase of [Ca2+]i at longer exposures, which correlated with SIN-1-induced necrotic cell death. Both effects of SIN-1 on [Ca2+]i are blocked by superoxide dismutase plus catalase and by Mn(III)tetrakis(4-benzoic acid)porphyrin chloride. Supplementation of Locke's buffer with catalase before addition of 0.5-1 mm SIN-1 had no effect on the decrease of [Ca2+]i but further delayed and attenuated the increase of [Ca2+]i observed after 60-120 min exposure to SIN-1 and also protected against SIN-1-induced necrotic cell death. alpha-Tocopherol, the potent NMDA receptor antagonist (+)-MK-801 and the N- and P-type Ca2+ channels blocker omega-conotoxin MVIIC had no effect on the alterations of [Ca2+]i upon exposure to SIN-1. However, inhibition of the plasma membrane Ca2+ ATPase can account for the increase of [Ca2+]i observed after 60-120 min exposure to 0.5-1 mm SIN-1. It is concluded that L-type Ca2+ channels are a primary target of SIN-1-induced extracellular nitrosative/oxidative stress, being inactivated by chronic exposure to fluxes of peroxynitrite of 0.5-1 microm/min, while higher concentrations of peroxynitrite and hydrogen peroxide are required for the inhibition of the plasma membrane Ca2+ ATPase and induction of necrotic cell death, respectively.  相似文献   

17.
Human lymphocytes were exposed to increasing concentrations of SIN-1, which generates superoxide and nitric oxide, and the formation of single-strand breaks (SSB) in individual cells was determined by the single-cell gel electrophoresis assay (comet assay). A dose- and time-dependent increase in SSB formation was observed rapidly after the addition of SIN-1 (0.1-15 mM). Exposure of the cells to SIN-1 (5 mM) in the presence of excess of superoxide dismutase (0.375 mM) increased the formation of SSB significantly, whereas 1000 U/ml catalase significantly decreased the quantity of SSB. The simultaneous presence of both superoxide dismutase and catalase before the addition of SIN-1 brought the level of SSB to that of the untreated cells. Moreover, pretreatment of the cells with the intracellular Ca(2+)-chelator BAPTA/AM inhibited SIN-1-induced DNA damage, indicating the involvement of intracellular Ca(2+) changes in this process. On the other hand, pretreatment of the same cells with ascorbate or dehydroascorbate did not offer any significant protection in this system. The data suggest that H2O2-induced changes in Ca(2+) homeostasis are the predominant pathway for the induction of SSB in human lymphocytes exposed to oxidants.  相似文献   

18.
Endothelium-dependent, nitric oxide (NO)-mediated vasodilation can be impaired by reactive oxygen species (ROS), and this deleterious effect of ROS on NO availability may increase with aging. Endothelial function declines rapidly after menopause, possibly because of loss of circulating estrogen and its antioxidant effects. The purpose of the current study was to determine the role of O(2)(-) and H(2)O(2) in regulating flow-induced dilation in coronary arterioles of young (6-mo) and aged (24-mo) intact, ovariectomized (OVX), or OVX + estrogen-treated (OVE) female Fischer 344 rats. Both aging and OVX reduced flow-induced NO production, whereas flow-induced H(2)O(2) production was not altered by age or estrogen status. Flow-induced vasodilation was evaluated before and after treatment with the superoxide dismutase (SOD) mimetic Tempol (100 μM) or the H(2)O(2) scavenger catalase (100 U/ml). Removal of H(2)O(2) with catalase reduced flow-induced dilation in all groups, whereas Tempol diminished vasodilation in intact and OVE, but not OVX, rats. Immunoblot analysis revealed elevated nitrotyrosine with aging and OVX. In young rats, OVX reduced SOD protein while OVE increased SOD in aged rats; catalase protein did not differ in any group. Collectively, these studies suggest that O(2)(-) and H(2)O(2) are critical components of flow-induced vasodilation in coronary arterioles from female rats; however, a chronic deficiency of O(2)(-) buffering by SOD contributes to impaired flow-induced dilation with aging and loss of estrogen. Furthermore, these data indicate that estrogen replacement restores O(2)(-) homeostasis and flow-induced dilation of coronary arterioles, even at an advanced age.  相似文献   

19.
Protein oxidation and tyrosine nitration are two major post-translational modifications of protein by reactive nitrogen oxide species, which are mainly produced by peroxynitrite and heme peroxidases (hemin)-H(2)O(2)-NO(2)(-) system. We report herein some novel phenomena between hemin-H(2)O(2)-NO(2)(-) and 3-morpholinosydnonimine hydrochloride (SIN-1)-mediated oxidation and nitration reactions of glutamate dehydrogenase (GDH). Hemin-H(2)O(2) could effectively induce GDH protein oxidation and reduce its activity. Although the addition of low concentration of nitrite promoted protein oxidation, protein oxidation was weakened with the increase of nitrite concentration, meanwhile, tyrosine nitration was increased and the enzyme activity was partially restored. However, with the increase of SIN-1 concentration, protein oxidation and tyrosine nitration were increased, enzyme activity was decreased. The presence of desferrioxamine and/or catechin inhibit tyrosine nitration both in hemin-H(2)O(2)-NO(2)(-) and in SIN-1, but they promoted protein oxidation and reduced the enzyme activity in hemin-H(2)O(2)-NO(2)(-) system, while inhibited protein oxidation and recover the enzyme activity in SIN-1 system. These results reveal both hemin-H(2)O(2)-NO(2)(-) and SIN-1 can cause inactivation of GDH through protein oxidation and tyrosine nitration, but the impact of the effect of protein oxidation (not thiol oxidation) on enzyme activity is stronger than that of protein tyrosine nitration. Moreover, mass spectrometric analysis indicated that nitrated tyrosine residues by hemin-H(2)O(2)-NO(2)(-) were Tyr262 and Tyr471 while by SIN-1 were Tyr401 and Tyr493. It meant that protein oxidation and tyrosine nitration of GDH induced by hemin-H(2)O(2)-NO(2)(-) were different from those induced by SIN-1.  相似文献   

20.
The survival of skeletal muscle myoblasts in culture after exposure either to a donor of NO, sodium nitroprusside (SNP), or ethanamine, 2,2'-(hydroxynitrosohydrazono)bis-(DETA NONOate), or to a donor of both NO and O(-)(2), 3-morpholinosydnonimine hydrochloride (SIN-1), was investigated. SIN-1 reduced clonogenic survival markedly but donors of NO alone did not. The injurious effect of SIN-1 was prevented by oxyhemoglobin or by uric acid but not by superoxide dismutase. The exposure of myoblasts to authentic peroxynitrite (ONOO(-)) or to DETA NONOate in the presence of an O(-)(2)-generating system did not reduce their survival. The results show that NO or ONOO(-) alone is not detrimental to myoblast survival and suggest that SIN-1 toxicity is, at least in part, mediated by H(2)O(2) in this myoblast culture system.  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