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.
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   

2.
L-Tyrosine and L-tyrosine residues in proteins are attacked by various reactive-nitrogen species (RNS) including peroxynitrite to form 3-nitrotyrosine (NO(2)Tyr) and protein-associated 3-nitrotyrosine (NO(2)TyrProt). Circulating NO(2)Tyr and NO(2)TyrProt have been suggested and are widely used as biomarkers of oxidative stress in humans. In this article the mass spectrometry (MS)-based analytical methods recently reported for the quantification of circulating levels of NO(2)Tyr and NO(2)TyrProt are discussed. These methodologies differ in sensitivity, selectivity, specificity and accessibility to interferences with the latter mainly arising from artifactual formation of NO(2)Tyr and NO(2)TyrProt during sample treatment such as acidification and chemical derivatization. Application of these methodologies to healthy normal humans revealed basal circulating levels for NO(2)Tyr which range between 0.7 and 64 nM, i.e. by two orders of magnitude. Application of gas chromatography-tandem mass spectrometry (GC-tandem MS) methods by two independent research groups by using two different protocols to avoid artifactual nitration of L-tyrosine revealed almost identical mean plasma levels of the order of 1.0 nM in healthy humans. The lower limits of quantitation (LOQ) of these methods were 0.125 and 0.3n M, respectively. This order of magnitude for basal NO(2)Tyr is supported by two liquid chromatography-tandem mass spectrometry (LC-tandem MS) methods with LOQ values of 4.4 and 1.4 nM. On the basis of the data provided by GC-tandem MS and LC-tandem MS the use of a range of 0.5-3 nM for NO(2)Tyr and of 0.6 pmol/mg plasma protein or a molar ratio of 3-nitrotyrosine to tyrosine in plasma proteins of the order of 1:10(6) for NO(2)TyrProt in plasma of healthy humans as reference values appear reasonably justified. Recently reported clinical studies involving 3-nitrotyrosine as a biomarker of oxidative stress are discussed in particular from the analytical point of view.  相似文献   

3.
The intermittent vascular occlusion occurring in sickle cell disease (SCD) leads to ischemia-reperfusion injury and activation of inflammatory processes including enhanced production of reactive oxygen species and increased expression of inducible nitric-oxide synthase (NOS2). Appreciating that impaired nitric oxide-dependent vascular function and the concomitant formation of oxidizing and nitrating species occur in concert with increased rates of tissue reactive oxygen species production, liver and kidney NOS2 expression, tissue 3-nitrotyrosine (NO(2)Tyr) formation and apoptosis were evaluated in human SCD tissues and a murine model of SCD. Liver and kidney NOS2 expression and NO(2)Tyr immunoreactivity were significantly increased in SCD mice and humans, but not in nondiseased tissues. TdT-mediated nick end-label (TUNEL) staining showed apoptotic cells in regions expressing elevated levels of NOS2 and NO(2)Tyr in all SCD tissues. Gas chromatography mass spectrometry analysis revealed increased plasma protein NO(2)Tyr content and increased levels of hepatic and renal protein NO(2)Tyr derivatives in SCD (21.4 +/- 2.6 and 37.5 +/- 7.8 ng/mg) versus wild type mice (8.2 +/- 2.2 and 10 +/- 1.2 ng/mg), respectively. Western blot analysis and immunoprecipitation of SCD mouse liver and kidney proteins revealed one principal NO(2)Tyr-containing protein of 42 kDa, compared with controls. Enzymatic in-gel digestion and MALDI-TOF mass spectrometry identified this nitrated protein as actin. Electrospray ionization and fragment analysis by tandem mass spectrometry revealed that 3 of 15 actin tyrosine residues are nitrated (Tyr(91), Tyr(198), and Tyr(240)) at positions that significantly modify actin assembly. Confocal microscopy of SCD human and mouse tissues revealed that nitration led to morphologically distinct disorganization of filamentous actin. In aggregate, we have observed that the hemoglobin point mutation of sickle cell disease that mediates hemoglobin polymerization defects is translated, via inflammatory oxidant reactions, into defective cytoskeletal polymerization.  相似文献   

4.
3-nitrotyrosine (NO2Tyr), an L-tyrosine derivative during nitrative stress, can substitute the COOH-terminal tyrosine of alpha-tubulin, posttranslationally altering microtubular functions. Because infection of the cells by respiratory syncytial virus (RSV) may require intact microtubules, we tested the hypothesis that NO2Tyr would inhibit RSV infection and intracellular signaling via nitrotyrosination of alpha-tubulin. A human bronchial epithelial cell line (BEAS-2B) was incubated with RSV with or without NO2Tyr. The release of chemokines and viral particles and activation of interferon regulatory factor-3 (IRF-3) were measured. Incubation with NO2Tyr increased nitrotyrosinated alpha-tubulin, and NO2Tyr colocalized with microtubules. RSV-infected cells released viral particles, RANTES, and IL-8 in a time- and dose-dependent manner, and intracellular RSV proteins coprecipitated with alpha-tubulin. NO2Tyr attenuated the RSV-induced release of RANTES, IL-8, and viral particles by 50-90% and decreased alpha-tubulin-associated RSV proteins. 3-chlorotyrosine, another L-tyrosine derivative, had no effects. NO2Tyr also inhibited the RSV-induced shift of the unphosphorylated form I of IRF-3 to the phosphorylated form II. Pre-exposure of the cells to NO(2) (0.15 ppm, 4 h), which produced diffuse protein tyrosine nitration, did not affect RSV-induced release of RANTES, IL-8, or viral particles. NO2Tyr did not affect the potential of viral spreading to the neighboring cells since the RSV titers were not decreased when the uninfected cells were cocultured with the preinfected cells in NO2Tyr-containing medium. These results indicate that NO2Tyr, by replacing the COOH-terminal tyrosine of alpha-tubulin, attenuated RSV infection, and the inhibition appeared to occur at the early stages of RSV infection.  相似文献   

5.
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   

6.
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   

7.
Nitric oxide (NO) is a precursor of reactive nitrating species, peroxynitrite and nitrogen dioxide, which modify proteins to generate oxidized species such as 3-nitrotyrosine that has been used as a hallmark of peroxynitrite-mediated oxidative stress on proteins. In the last few years however, a growing body of evidence indicates that NO also regulates a myriad of physiologic responses by modifying tyrosine residues. Looking for the molecular event triggered by NO in nerve growth factor (NGF)-induced neuronal differentiation, we recently reported that in differentiating PC12 cells, the cytoskeleton becomes the main cellular fraction containing nitrotyrosinated proteins, and alpha-tubulin is the major target. In the present work, we focus on the investigation of the sites of tyrosine nitration in alpha-tubulin purified by two-dimensional gel electrophoresis following anti-alpha-tubulin immunoprecipitation of protein extract from NGF-treated PC12 cells. Using Western blotting and matrix-assisted laser desorption/ionization-time of flight analysis, we show for the first time, both in vivo and in vitro, that nitration can occur on alpha-tubulin at sites other than the C-terminus and we positively identify Tyr 161 and Tyr 357 as two specific amino acids endogenously nitrated.  相似文献   

8.
We noted differences in the antibody response to 3-nitrotyrosine (NO(2)Tyr) in fixed and non-fixed tissues, and studied therefore potential problems associated with non-fixed tissues in Western blot analyses. Three different monoclonal anti-nitrotyrosine antibodies in Western blot analysis of inflammatory stimulated rat abdominal, liver and lung tissue homogenates caused no immunoreactivity, in contrast to a polyclonal nitrotyrosine antibody applied in fixed and non-fixed tissues. Western blot studies using both mono- and polyclonal antibodies showed a temperature- and heme group-dependent reduction of NO(2)Tyr in nitrated rat and bovine serum albumin incubated with dithiothreitol. Mass spectrometric analyses of a nitrated peptide angiotensin II revealed under similar conditions a positive temperature effect between 56 and 70 degrees C on reduction of NO(2)Tyr to 3-aminotyrosine which is not detected by anti-NO(2)Tyr antibodies. Western blot analysis may therefore underestimate the level of tissue nitration, and factors causing a reduction of NO(2)Tyr during sample preparation might conceal the actual nitration of proteins.  相似文献   

9.
3-nitrotyrosine (NO2-Tyr) is thought to be a specific marker of cell injury during oxidative damage. We have evaluated the role of poly(ADP-ribose)polymerase-1 (PARP-1) in protein nitration after treatment of immortalized fibroblasts parp-1+/+ and parp-1-/- with the alkylating agent 2'-methyl-2'-nitroso-urea (MNU). Both cell lines showed increased iNOS expression following MNU treatment in parallel with a selective induction of tyrosine nitration of different proteins. PARP-1 deficient cells displayed a delayed iNOS accumulation, reduced number of nitrated proteins, and a lower global nitrotyrosine "footprint." We have identified the mitochondrial compartment as the major site of oxidative stress during DNA damage, being MnSOD one of the NO2-Tyr-modified proteins, but not in parp-1-/- cells. These results suggest that NO-derived injury can be modulated by proteins involved in the response to genotoxic damage, such as PARP-1, and may account for the limited oxidative injury in parp-1 knockout mice during carcinogenesis and inflammation.  相似文献   

10.
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   

11.
Myoglobin-catalyzed tyrosine nitration: no need for peroxynitrite.   总被引:1,自引:0,他引:1  
The nitration of tyrosine residues in protein to yield 3-nitrotyrosine derivatives has been suggested to represent a specific footprint for peroxynitrite formation in vivo. However, recent studies suggest that certain hemoproteins such as peroxidases catalyze the H(2)O(2)-dependent nitration of tyrosine to yield 3-nitrotyrosine in a peroxynitrite-independent reaction. Because 3-nitrotyrosine has been shown to be present in the postischemic myocardium, we wished to assess the ability of myoglobin to catalyze the nitration of tyrosine in vitro. We found that myoglobin catalyzed the oxidation of nitrite and promoted the nitration of tyrosine. Both nitrite oxidation and tyrosine nitration were H(2)O(2)-dependent and required the formation of ferryl (Fe(+4)) myoglobin. In addition, nitrite oxidation and tyrosine nitration were pH-dependent with a pH optimum of approximately 6.0. Taken together, these data suggest that the acidic pH and low oxygen tension produced during myocardial ischemia will facilitate myoglobin-catalyzed, peroxyntrite-independent formation of 3-nitrotyrosine.  相似文献   

12.
3-Morpholinosyndnomine (SIN-1) has been reported to be a peroxynitrite (OONO(-)) donor because it produces both nitric oxide (NO) and superoxide (O(2)(-).) upon decomposition in aqueous solution. However, SIN-1 can decompose to primarily NO in the presence of electron acceptors, including those found in biological tissues, making it necessary to determine the release product(s) formed in any given biological system. In a mixed cortical cell culture system, SIN-1 caused a concentration-dependent increase in cortical cell injury with a parallel increase in the release of cellular proteins containing 3-nitrotyrosine into the culture medium. The increase in 3-nitrotyrosine immunoreactivity, a footprint of OONO(-) production, was specific for SIN-1 as exposure to neurotoxic concentrations of an NO donor (Z)-1-[2-aminoethyl)-N-(2-ammonioethyl) aminodiazen-1-ium-1,2-diolate (DETA/NO), or NMDA did not result in the nitration of protein tyrosine residues. Both SIN-1-induced injury and 3-nitrotyrosine staining were prevented by the addition of either 5,10,15,20-Tetrakis (4-sulfonatophenyl) prophyrinato iron (III) [FeTPPS], an OONO(-) decomposition catalyst, or uric acid, an OONO(-) scavenger. Removal of NO alone was sufficient to inhibit the formation of OONO(-) from SIN-1 as well as its cytotoxicity. Removal of O(2)(-). and the subsequently formed H(2)O(2) by superoxide dismutase (SOD) plus catalase likewise prevented the nitration of protein-bound tyrosine but actually enhanced the cytotoxicity of SIN-1, indicating that cortical cells can cope with the oxidative but not the nitrosative stress generated. Finally, neural injury induced by SIN-1 in unadulterated cortical cells was prevented by antagonism of AMPA/kainate receptors, while blockade of the NMDA receptor was without effect. In contrast, activation of both NMDA and non-NMDA receptors contributed to the SIN-1-mediated neurotoxicity when cultures were exposed in the presence of SOD plus catalase. Thus, whether SIN-1 initiates neural cell death in an OONO(-)-dependent or -independent manner is determined by the antioxidant status of the cells. Further, the mode of excitotoxicity by which injury progresses is determined by the NO-related species generated.  相似文献   

13.
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   

14.
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   

15.
Reactive intermediates derived from nitric oxide ((*)NO) are thought to play a contributing role in disease states associated with inflammation and infection. We show here that glutathione S-transferases (GSTs), principal enzymes responsible for detoxification of endogenous and exogenous electrophiles, are susceptible to inactivation by reactive nitrogen species (RNS). Treatment of isolated GSTs or rat liver homogenates with either peroxynitrite, the myeloperoxidase/hydrogen peroxide/nitrite system, or tetranitromethane, resulted in loss of GST activity with a concomitant increase in the formation of protein-associated 3-nitrotyrosine (NO(2)Tyr). This inactivation was only partially (<25%) reversible by dithiothreitol, and exposure of GSTs to hydrogen peroxide or S-nitrosoglutathione was only partially inhibitory (<25%) and did not result in protein nitration. Thus, irreversible modifications such as tyrosine nitration may have contributed to GST inactivation by RNS. Since all GSTs contain a critical, highly conserved, active-site tyrosine residue, we postulated that this Tyr residue might present a primary target for nitration by RNS, thus leading to enzyme inactivation. To directly investigate this possibility, we analyzed purified mouse liver GST-mu, following nitration by several RNS, by trypsin digestion, HPLC separation, and matrix-assisted laser desorption/ionization-time of flight analysis, to determine the degree of tyrosine nitration of individual Tyr residues. Indeed, nitration was found to occur preferentially on several tyrosine residues located in and around the GST active site. However, RNS concentrations that resulted in near complete GST inactivation only caused up to 25% nitration of even preferentially targeted tyrosine residues. Hence, nitration of active-site tyrosine residues may contribute to GST inactivation by RNS, but is unlikely to fully account for enzyme inactivation. Overall, our studies illustrate a potential mechanism by which RNS may promote (oxidative) injury by environmental pollutants in association with inflammation.  相似文献   

16.
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   

17.
Earlier findings on electronitration of hen egg-white lysozyme demonstrated a product which was mononitrated at Tyr23, by ion-exchange chromatography, absorbance at 430 nm, dithionite reduction, and Edman sequencing of a nitrated proteolytic peptide. However, the whole protein was not sequenced; therefore, although the enzyme remained active upon nitration, reaction at other residues could not be completely eliminated. This study has now been extended to the redox protein myoglobin. We demonstrate the novel electronitration (electrooxidation in the presence of nitrite) of a specific tyrosine residue in horse heart myoglobin and also in apomyoglobin. Production of the yellow chromophore, 3-nitrotyrosine (3-NT), was apparent in apomyoglobin from A430 but was masked in holomyoglobin by the Soret band. In both cases, the presence of 3-NT in the electronitrated samples was further indicated by the binding of antibody to 3-NT in Western blots. High-resolution electrospray ionization (ESI) Fourier transform ion cyclotron resonance (FTICR) mass spectrometry revealed a reaction product at [M + 45] (consistent with substitution of NO2 for H), indicating that the nitration reaction is the only reaction occurring which gives rise to a change in mass in the electrooxidation. Fragmentation mass spectrometry identified the nitration site as Tyr103, with no nitration at Tyr146. The procedure may be useful in preparing model nitrated proteins for the study of disease mechanisms.  相似文献   

18.
The extracellular alkaline proteinase subtilisin DY was nitrated with increasing amounts of tetranitromethane. At 2-fold molar excess of the reagent with respect to the tyrosine residues in the enzyme, when 1.3 residues were modified, a peak of the caseinolytic activity (13% increase) was observed. Evidence is provided that the diminishing of the pK of the phenolic hydroxyl group in Tyr(3NO2)104 causes this phenomenon. The products obtained after nitration of the enzyme with 5-fold and 200-fold molar excess of tetranitromethane were cleaved by trypsin and cyanogen bromide and the peptides obtained were studied by analysis with respect to the tyrosine and 3-nitrotyrosine residues. Their degree of substitution was established. Tyrosine-104 was the first modified residue, then follow the residues with numbers 57, 143, 206, 262 and somewhat later 21, 209, 263, all fully modified by 200-fold molar excess of the reagent. Partial modification was observed at numbers 91, 167, 214, 238 and no modification at numbers 6 and 171. It has been established that the nonmodified residues are buried inside the molecule and the partially modified residues are screened by the side chains of lysine, valine, leucine, and tryptophan as seen on a working video three-dimensional model of subtilisin Carlsberg. The approach for characterization of tyrosyl groups in proteins based on peptide sequencing and HPLC quantitation of the phenylthiohydantoin derivatives of tyrosine and 3-nitrotyrosine was further developed with respect to the quantitation of the HPLC-separated peptides using fragments of the protein studied.  相似文献   

19.
Eosinophil recruitment and enhanced production of NO are characteristic features of asthma. However, neither the ability of eosinophils to generate NO-derived oxidants nor their role in nitration of targets during asthma is established. Using gas chromatography-mass spectrometry we demonstrate a 10-fold increase in 3-nitrotyrosine (NO(2)Y) content, a global marker of protein modification by reactive nitrogen species, in proteins recovered from bronchoalveolar lavage of severe asthmatic patients (480 +/- 198 micromol/mol tyrosine; n = 11) compared with nonasthmatic subjects (52.5 +/- 40.7 micromol/mol tyrosine; n = 12). Parallel gas chromatography-mass spectrometry analyses of bronchoalveolar lavage proteins for 3-bromotyrosine (BrY) and 3-chlorotyrosine (ClY), selective markers of eosinophil peroxidase (EPO)- and myeloperoxidase-catalyzed oxidation, respectively, demonstrated a dramatic preferential formation of BrY in asthmatic (1093 +/- 457 micromol BrY/mol tyrosine; 161 +/- 88 micromol ClY/mol tyrosine; n = 11 each) compared with nonasthmatic subjects (13 +/- 14.5 micromol BrY/mol tyrosine; 65 +/- 69 micromol ClY/mol tyrosine; n = 12 each). Bronchial tissue from individuals who died of asthma demonstrated the most intense anti-NO(2)Y immunostaining in epitopes that colocalized with eosinophils. Although eosinophils from normal subjects failed to generate detectable levels of NO, NO(2-), NO(3-), or NO(2)Y, tyrosine nitration was promoted by eosinophils activated either in the presence of physiological levels of NO(2-) or an exogenous NO source. At low, but not high (e.g., >2 microM/min), rates of NO flux, EPO inhibitors and catalase markedly attenuated aromatic nitration. These results identify eosinophils as a major source of oxidants during asthma. They also demonstrate that eosinophils use distinct mechanisms for generating NO-derived oxidants and identify EPO as an enzymatic source of nitrating intermediates in eosinophils.  相似文献   

20.
Protein 3-nitrotyrosine is a post-translational modification that commonly arises from the nitration of tyrosine residues. This modification has been detected under a wide range of pathological conditions and has been shown to alter protein function. Whether 3-nitrotyrosine is important in normal cellular processes or is likely to affect specific biological pathways remains unclear. Using GPS-YNO2, a recently described 3-nitrotyrosine prediction algorithm, a set of predictions for nitrated residues in the human proteome was generated. In total, 9.27 per cent of the proteome was predicted to be nitratable (27 922/301 091). By matching the predictions against a set of curated and experimentally validated 3-nitrotyrosine sites in human proteins, it was found that GPS-YNO2 is able to predict 73.1 per cent (404/553) of these sites. Furthermore, of these sites, 42 have been shown to be nitrated endogenously, with 85.7 per cent (36/42) of these predicted to be nitrated. This demonstrates the feasibility of using the predicted dataset for a whole proteome analysis. A comprehensive bioinformatics analysis was subsequently performed on predicted and all experimentally validated nitrated tyrosine. This found mild but specific biophysical constraints that affect the susceptibility of tyrosine to nitration, and these may play a role in increasing the likelihood of 3-nitrotyrosine to affect processes, including phosphorylation and DNA binding. Furthermore, examining the evolutionary conservation of predicted 3-nitrotyrosine showed that, relative to non-nitrated tyrosine residues, 3-nitrotyrosine residues are generally less conserved. This suggests that, at least in the majority of cases, 3-nitrotyrosine is likely to have a deleterious effect on protein function and less likely to be important in normal cellular function.  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