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.
Protein tyrosine nitration in hydrophilic and hydrophobic environments   总被引:1,自引:1,他引:0  
Summary. In this review we address current concepts on the biological occurrence, levels and consequences of protein tyrosine nitration in biological systems. We focused on mechanistic aspects, emphasizing on the free radical mechanisms of protein 3-nitrotyrosine formation and critically analyzed the restrictions for obtaining large tyrosine nitration yields in vivo, mainly due to the presence of strong reducing systems (e.g. glutathione) that can potently inhibit at different levels the nitration process. Evidence is provided to show that the existence of metal-catalyzed processes, the assistance of nitric oxide-dependent nitration steps and the facilitation by hydrophobic environments, provide individually and/or in combination, feasible scenarios for nitration in complex biological milieux. Recent studies using hydrophobic tyrosine analogs and tyrosine-containing peptides have revealed that factors controlling nitration in hydrophobic environments such as biomembranes and lipoproteins can differ to those in aqueous compartments. In particular, exclusion of key soluble reductants from the lipid phase will more easily allow nitration and lipid-derived radicals are suggested as important mediators of the one-electron oxidation of tyrosine to tyrosyl radical in proteins associated to hydrophobic environments. Development and testing of hydrophilic and hydrophobic probes that can compete with endogenous constituents for the nitrating intermediates provide tools to unravel nitration mechanisms in vitro and in vivo; additionally, they could also serve to play cellular and tissue protective functions against the toxic effects of protein tyrosine nitration.  相似文献   

2.
Peroxynitrite has been receiving increasing attention as the pathogenic mediator of nitric oxide cytotoxicity. In most cases, the contribution of peroxynitrite to diseases has been inferred from detection of 3-nitrotyrosine in injured tissues. However, presently it is known that other nitric oxide-derived species can also promote protein nitration. Mechanistic details of protein nitration remain under discussion even in the case of peroxynitrite, although recent literature data strongly suggest a free radical mechanism. Here, we confirm the free radical mechanism of tyrosine modification by peroxynitrite in the presence and in the absence of the bicarbonate-carbon dioxide pair by analyzing the stable tyrosine products and the formation of the tyrosyl radical at pH 5.4 and 7.4. Stable products, 3-nitrotyrosine, 3-hydroxytyrosine, and 3, 3-dityrosine, were identified by high performance liquid chromatography and UV spectroscopy. The tyrosyl radical was detected by continuous-flow and spin-trapping electron paramagnetic resonance (EPR). 3-Hydroxytyrosine was detected at pH 5.4 and its yield decreased in the presence of the bicarbonate-carbon dioxide pair. In contrast, the yields of the tyrosyl radical increased in the presence of the bicarbonate-carbon dioxide pair and correlated with the yields of 3-nitrotyrosine under all tested experimental conditions. Taken together, the results demonstrate that the promoting effects of carbon dioxide on peroxynitrite-mediated tyrosine nitration is due to the selective reactivity of the carbonate radical anion as compared with that of the hydroxyl radical. Colocalization of 3-hydroxytyrosine and 3-nitrotyrosine residues in proteins may be useful to discriminate between peroxynitrite and other nitrating species.  相似文献   

3.
The mechanism of the reactions of myoglobin and hemoglobin with *OH and CO3*- in the presence of oxygen was studied using pulse and gamma-radiolysis. Unlike *NO2, which adds to the porphyrin iron, *OH and CO3*- form globin radicals. These secondary radicals oxidize the Fe(II) center through both intra- and intermolecular processes. The intermolecular pathway was further demonstrated when BSA radicals derived from *OH or CO3*- oxidized oxyhemoglobin and oxymyoglobin to their respective ferric states. The oxidation yields obtained by pulse radiolysis were lower compared to gamma-radiolysis, where the contribution of radical-radical reactions is negligible. Full oxidation yields by *OH-derived globin radicals could be achieved only at relatively high concentrations of the heme protein mainly via an intermolecular pathway. It is suggested that CO3*- reaction with the protein yields Tyr and/or Trp-derived phenoxyl radicals, which solely oxidize the porphyrin iron under gamma-radiolysis conditions. The *OH particularly adds to aromatic residues, which can undergo elimination of H2O forming the phenoxyl radical, and/or react rapidly with O2 yielding peroxyl radicals. The peroxyl radical can oxidize a neighboring porphyrin iron and/or give rise to superoxide, which neither oxidize nor reduce the porphyrin iron. The potential physiological implications of this chemistry are that hemoglobin and myoglobin, being present at relatively high concentrations, can detoxify highly oxidizing radicals yielding the respective ferric states, which are not toxic.  相似文献   

4.
The flavanol (-)-epicatechin is known to protect against peroxynitrite-induced nitration and oxidation reactions. This study investigated the protection afforded by (-)-epicatechin against both these reaction types on one target molecule, the aminoacid tyrosine, in a hydrophilic milieu as well as with a lipophilic tyrosine derivative, N-t-BOC l-tyrosine tert-butyl ester (BTBE), bound to liposomes. The flavanol efficiently attenuated both tyrosine nitration and tyrosine dimerization (which is based on an initial oxidation reaction) and was active in the hydrophilic and hydrophobic systems at similar IC(50) values, approximately 0.02-0.05 mol (-)-epicatechin/mol peroxynitrite. Related procyanidin oligomers of different chain-length (dimer to octamer) were also tested for their protective properties, and exhibited protection that, on a monomer basis, was in the same order of magnitude as those for (-)-epicatechin.  相似文献   

5.
Disulfide bond reduction by the CO2.- radical was investigated in aponeocarzinostatin, aporiboflavin-binding protein, and bovine immunoglobulin. Protein-bound cysteine free thiols were formed under gamma-ray irradiation in the course of a pH-dependent and protein concentration dependent chain reaction. The chain efficiency increased upon acidification of the medium, with an apparent pKa around 5, and decreased abruptly below pH 3.6. It decreased also at neutral pH as cysteine accumulated. From pulse radiolysis analysis, CO2.- proved able to induce rapid one-electron oxidation of thiols and of tyrosine phenolic groups in addition to one-electron donation to exposed disulfide bonds. The bulk rate constant of CO2.- uptake by the native proteins was 5- to 10-fold faster at pH 3 than at pH 8, and the protonated form of the disulfide radical anion, [symbol: see text], appeared to be the major protein radical species formed under acidic conditions. The main decay path of [symbol: see text] consisted of the rapid formation of a thiyl radical intermediate [symbol: see text] in equilibrium with the closed, cyclic form. The thiyl radical was subsequently reduced to the sulfhydryl level [symbol: see text] on reaction with formate, generating 1 mol of the CO2.- radical, thus propagating the chain reaction. The disulfide radical anion [symbol: see text] at pH 8 decayed through competing intramolecular and/or intermolecular routes including disproportionation, protein-protein cross-linking, electron transfer with tyrosine residues, and reaction with sulfhydryl groups in prereduced systems. Disproportionation and cross-linking were observed with the riboflavin-binding protein solely. Formation of the disulfide radical cation [symbol: see text], phenoxyl radical Tyr-O. disproportionation, and phenoxyl radical induced oxidation of preformed thiol groups should also be taken into consideration to explain the fate of the oxygen-centered phenoxyl radical.  相似文献   

6.
By application of pulse radiolysis it was demonstrated that nitrogen dioxide (NO2.) oxidizes Gly-Tyr in aqueous solution with a strongly pH-dependent rate constant (k6 = 3.2 X 10(5) M-1 S-1 at pH 7.5 and k6 = 2.0 X 10(7) M-1 S-1 at pH 11.3), primarily generating phenoxyl radicals. The phenoxyl can react further with NO2. (k7 approximately 3 X 10(9) M-1 S-1) to form nitrotyrosine, which is the predominant final product in neutral solution and at low tyrosyl concentrations under gamma-radiolysis conditions. Tyrosine nitration is less efficient in acidic solution, due to the natural disproportionation of NO2., and in alkaline solutions and at high tyrosyl concentrations due to enhanced tyrosyl dimerization. Selective tyrosine nitration by interaction of NO2. with proteins (at pH 7 to 9) was demonstrated in the case of histone, lysozyme, ribonuclease A, and subtilisin Carlsberg. Nitrotyrosine developed slowly also under incubation of Gly-Tyr with nitrite at pH 4 to 5, where NO2. is formed by acid decomposition of HONO. It is recalled in this context that NO2.-induced oxidations, by regenerating NO2-, can propagate NO2./NO2- redox cycling under acidic conditions. Even faster than with tyrosine is the NO2.-induced oxidation of cysteine-thiolate (k9 = 2.4 X 10(8) M-1 S-1 at pH 9.2), involving the transient formation of cystinyl radical anions. The interaction of NO2. with Gly-Trp was comparably slow (k approximately 10(6) M-1 S-1), and no reaction was detectable by pulse radiolysis with Met-Gly and (Cys-Gly)2, or with DNA. Slow reactions of NO2. were observed with arachidonic acid (k approximately 10(6) M-1 S-1 at pH 9.0) and with linoleate (k approximately 2 X 10(5) M-1 S-1 at pH 9.4), indicating that NO2. is capable of initiating lipid peroxidation even in an aqueous environment. NO2.-Induced tyrosine nitration, using 50 microM Gly-Tyr at pH 8.2, was hardly inhibited, however, in the presence of 1 mM linoleate, and was not affected at all in the presence of 5 mM dimethylamine (a nitrosamine precursor). It is concluded that protein modifications, and particularly phenol and thiol oxidation, may be an important mechanism, as well as initiation of lipid peroxidation, of action of NO2. in biological systems.  相似文献   

7.
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   

8.
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   

9.
The iron chelating agent desferrioxamine inhibits peroxynitrite-mediated oxidations and attenuates nitric oxide and oxygen radical-dependent oxidative damage both in vitro and in vivo. The mechanism of protection is independent of iron chelation and has remained elusive over the past decade. Herein, stopped-flow studies revealed that desferrioxamine does not react directly with peroxynitrite. However, addition of peroxynitrite to desferrioxamine in both the absence and the presence of physiological concentrations of CO2 and under excess nitrite led to the formation of a one-electron oxidation product, the desferrioxamine nitroxide radical, consistent with desferrioxamine reacting with the peroxynitrite-derived species carbonate (CO3*-) and nitrogen dioxide (*NO2) radicals. Desferrioxamine inhibited peroxynitrite-dependent free radical-mediated processes, including tyrosine dimerization and nitration, oxyhemoglobin oxidation in the presence of CO2, and peroxynitrite plus carbonate-dependent chemiluminescence. The direct two-electron oxidation of glutathione by peroxynitrite was unaffected by desferrioxamine. The reactions of desferrioxamine with CO3*- and *NO2 were unambiguously confirmed by pulse radiolysis studies, which yielded second-order rate constants of 1.7 x 10(9) and 7.6 x 10(6) M(-1) s(-1), respectively. Desferrioxamine also reacts with tyrosyl radicals with k = 6.3 x 10(6) M(-1) s(-1). However, radical/radical combination reactions between tyrosyl radicals or of tyrosyl radical with *NO2 outcompete the reaction with desferrioxamine and computer-assisted simulations indicate that the inhibition of tyrosine oxidation can be fully explained by scavenging of the peroxynitrite-derived radicals. The results shown herein provide an alternative mechanism to account for some of the biochemical and pharmacological actions of desferrioxamine via reactions with CO3*- and *NO2 radicals.  相似文献   

10.
In blood, peroxynitrite (ONOO- ) and CO2 react to form NO2* and CO3* through the short-lived adduct ONOOCO2-, leading to protein-bound tyrosine nitration. ONOO(-) -modified LDL is atherogenic. Oxidatively modified LDL generally appears in the vessel wall surrounded by antioxidants. Human serum albumin (HSA) is one of them, partly associated to LDL as a LDL-albumin complex (LAC). The purpose was to study the effect of a mild nitration on LAC and whether albumin may interfere with LDL nitration. To do so, SIN-1 was used as ONOO- generator in the presence or absence of 25 mM HCO3-. The human serum albumin (HSA)-bound tyrosine nitration rate was found to be 4.4 x 10(-3) mol/mol in the presence of HCO3-. HSA decreased the LAC-tyrosine nitration rate from 2.5 x 10(-3) to 0.6 x 10(-3) mol/mol. It was concluded that HSA impaired the apoB-bound tyrosine nitration. These findings raise the question of the patho-physiological significance of these nitrations and their interactions which may potentially prevent both atheromatous plaque formation and endothelium dysfunction in particular and appear to be beneficial in terms of atherogenic risk.  相似文献   

11.
Electrolysis or horseradish peroxidase (HRP)-catalyzed oxidation of tyrosine and bityrosine in aqueous solution at pH 7.4 resulted in light emission in the visible region. Electrolysis of tyrosine emitted light which peaked at 490 nm and was almost completely quenched by superoxide dismutase (SOD), while emission by bityrosine peaked at 530 nm. In the HRP-H(2)O(2)-tyrosine system the oxidation-reduction of tyrosine emitted light with two prominent peaks, 490 and 530 nm, and was not quenched by SOD. The phenoxyl neutral radical of the tyrosine in HRP-H(2)O(2)-tyrosine system was detected by electron spin resonance (ESR) spectrometry using tert-nitrosobutane as a spin trap; the spin adduct was found to adhere to the HRP molecule during the enzymatic reaction. Further, bityrosine was detected in the HRP-H(2)O(2)-tyrosine reaction system. Changes in absorption spectra of HRP and chemiluminescence intensities during HRP-catalyzed oxidation of tyrosine suggest that for photon emission compound III is a candidate superoxide donor to the phenoxyl cation radical of tyrosine on the enzyme molecule. The luminescence observed in this study might be originated from at least two exciplexes involved with the tyrosine cation radical (Tyr(*+)) and the bityrosine cation radical (BT(*+))  相似文献   

12.
Alpha-lipoic acid (LA) and dihydrolipoic acid (DHLA) may have a role as antioxidants against nitric oxide-derived oxidants. We previously reported that peroxynitrite reacts with LA and DHLA with second-order rate constants of 1400 and 500 M(-1) s(-1), respectively, but indicated that these direct reactions are not fast enough to protect against peroxynitrite-mediated damage in vivo. Moreover, the mechanism of the reaction of peroxynitrite with LA has been recently challenged (J. Biol. Chem.279:9693-9697; 2004). Pulse radiolysis studies indicate that LA and DHLA react with peroxynitrite-derived nitrogen dioxide (*NO2) (k2 = 1.3 x 10(6) and 2.9 x 10(7) M(-1) s(-1), respectively) and carbonate radicals (CO(3-)) (k2 = 1.6 x 10(9) and 1.7 x 10(8) M(-1) s(-1), respectively). Carbonate radical-mediated oxidation of LA led to the formation of the potent one-electron oxidant LA radical cation. LA inhibited peroxynitrite-mediated nitration of tyrosine and of a hydrophobic tyrosine analog, N-t-BOC L-tyrosine tert-butyl ester (BTBE), incorporated into liposomes but enhanced tyrosine dimerization. Moreover, while LA competitively inhibited the direct oxidation of glutathione by peroxynitrite, it was poorly effective against the radical-mediated thiol oxidation. The mechanisms of reaction defined herein allow to rationalize the biochemistry of peroxynitrite based on direct and free radical-mediated processes and contribute to the understanding of the antioxidant actions of LA and DHLA.  相似文献   

13.
Light-induced nitration pathways of phenols are important processes for the transformation of pesticide-derived secondary pollutants into toxic derivatives in surface waters and for the formation of phytotoxic compounds in the atmosphere. Moreover, phenols can be used as ˙NO(2) probes in irradiated aqueous solutions. This paper shows that the nitration of 4-chlorophenol (4CP) into 2-nitro-4-chlorophenol (NCP) in the presence of irradiated nitrate and nitrite in aqueous solution involves the radical ˙NO(2). The experimental data allow exclusion of an alternative nitration pathway by ˙OH + ˙NO(2). Quantum mechanical calculations suggest that the nitration of both phenol and 4CP involves, as a first pathway, the abstraction of the phenolic hydrogen by ˙NO(2), which yields HNO(2) and the corresponding phenoxy radical. Reaction of phenoxyl with another ˙NO(2) follows to finally produce the corresponding nitrated phenol. Such a pathway also correctly predicts that 4CP undergoes nitration more easily than phenol, because the ring Cl atom increases the acidity of the phenolic hydrogen of 4CP. This favours the H-abstraction process to give the corresponding phenoxy radical. In contrast, an alternative nitration pathway that involves ˙NO(2) addition to the ring followed by H-abstraction by oxygen (or by ˙NO(2) or ˙OH) is energetically unfavoured and erroneously predicts faster nitration for phenol than for 4CP.  相似文献   

14.
Oxidation of tyrosine moieties by radicals involved in lipid peroxidation is of current interest; while a rate constant has been reported for reaction of lipid peroxyl radicals with a tyrosine model, little is known about the reaction between tyrosine and alkoxyl radicals (also intermediates in the lipid peroxidation chain reaction). In this study, the reaction between a model alkoxyl radical, the tert-butoxyl radical and tyrosine was followed using steady-state and pulse radiolysis. Acetone, a product of the β-fragmentation of the tert-butoxyl radical, was measured; the yield was reduced by the presence of tyrosine in a concentration- and pH-dependent manner. From these data, a rate constant for the reaction between tert-butoxyl and tyrosine was estimated as 6 ± 1 × 10(7) M(-1) s(-1) at pH 10. Tyrosine phenoxyl radicals were also monitored directly by kinetic spectrophotometry following generation of tert-butoxyl radicals by pulse radiolysis of solutions containing tyrosine. From the yield of tyrosyl radicals (measured before they decayed) as a function of tyrosine concentration, a rate constant for the reaction between tert-butoxyl and tyrosine was estimated as 7 ± 3 × 10(7) M(-1) s(-1) at pH 10 (the reaction was not observable at pH 7). We conclude that reaction involves oxidation of tyrosine phenolate rather than undissociated phenol; since the pK(a) of phenolic hydroxyl dissociation in tyrosine is ≈ 10.3, this infers a much lower rate constant, about 3 × 10(5) M(-1) s(-1), for the reaction between this alkoxyl radical and tyrosine at pH 7.4.  相似文献   

15.
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   

16.
Shi WQ  Cai H  Xu DD  Su XY  Lei P  Zhao YF  Li YM 《Regulatory peptides》2007,144(1-3):1-5
Proteins are targets of reactive nitrogen species such as peroxynitrite and nitrogen dioxide. Among the various amino acids in proteins, tyrosine and tryptophan residues are especially susceptible to attack by reactive nitrogen species. On the other hand, protein tyrosine phosphorylation has gained much attention in respect to cellular regulatory events and signal transduction. Peroxynitrite-mediated nitration of peptide YPPPPPW and phosphopeptide pYPPPPPW were studied at pH 7.4. The predominant nitrated products were separated and identified by reverse phase high performance liquid chromatography coupled with electrospray ionization mass spectrometry (LC-MS). The nitration sites were established by tandem electrospray ionization-mass spectrometry (LC-MS/MS). A regulatory effect of tyrosine phosphorylation/dephosphorylation on peptide nitration was observed. YPPPPPW was predominantly nitrated at tyrosine residue while pYPPPPPW was nitrated at tryptophan one. Our results can help in understanding the biochemical significance of the relationship of tyrosine phosphorylation and nitration in proteins.  相似文献   

17.
A A Shaw  A M Falick  M D Shetlar 《Biochemistry》1992,31(45):10976-10983
We report here the photoinduced formation of a thymine-N-acetyltyrosine adduct. Irradiation of dilute solutions of thymine in the presence of N-acetyltyrosine (NAT) leads to the formation of N-acetyl-4-hydroxy-3-(6-hydrothymin-5-yl)phenylalanine (I), isolated as a mixture of the 5R and 5S diastereoisomers; the photoreaction occurs when irradiation is done either at lambda = 254 nm or at wavelengths of lambda > 290 nm. Irradiation of thymidine in the presence of NAT and of thymine in the presence of tyrosine leads to analogous photoadducts. The photoreaction of thymine with NAT is completely quenched by oxygen and cannot be sensitized by acetone. The likely mechanism involves initial photoionization of the amino acid and deprotonation to form the phenoxyl radical. Thymine then probably captures the released aqueous electron, leading to protonation at C6 of the resulting radical anion. Combination of the phenoxyl and 5,6-dihydrothymin-5-yl radicals would then lead to formation of the final products. The quantum yield for production of the thymine-NAT adduct at pH 7.8 was estimated to be about 5.5 x 10(-4), while a value of 2.3 x 10(-3) was estimated for production of corresponding thymidine adduct at pH 8.1. The dependence of the quantum yield for adduct formation on pH has been determined for both the thymine and thymidine reactions with NAT; the maxima in the quantum yield profiles occur at pH 8-8.5, while appreciable values were measured at pH 7.5. We have also demonstrated that a similar reaction occurs when tyrosine is located within a peptide.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

18.
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   

19.
Peroxynitrite is a strong oxidant involved in cell injury. In tissues, most of peroxynitrite reacts preferentially with CO(2) or hemoproteins, and these reactions affect its fate and toxicity. CO(2) promotes tyrosine nitration but reduces the lifetime of peroxynitrite, preventing, at least in part, membrane crossing. The role of hemoproteins is not easily predictable, because the heme intercepts peroxynitrite, but its oxidation to ferryl species and tyrosyl radical(s) may catalyze tyrosine nitration. The modifications induced by peroxynitrite/CO(2) on oxyhemoglobin were determined by mass spectrometry, and we found that alphaTyr42, betaTyr130, and, to a lesser extent, alphaTyr24 were nitrated. The suggested nitration mechanism is tyrosyl radical formation by long-range electron transfer to ferrylhemoglobin followed by a reaction with (*)NO(2). Dityrosine (alpha24-alpha42) and disulfides (beta93-beta93 and alpha104-alpha104) were also detected, but these cross-linkings were largely due to modifications occurring under the denaturing conditions employed for mass spectrometry. Moreover, immunoelectrophoretic techniques showed that the 3-nitrotyrosine content of oxyhemoglobin sharply increased only in molar excess of peroxynitrite, thus suggesting that this hemoprotein is not a catalyst of nitration. The noncatalytic role may be due to the formation of the nitrating species (*)NO(2) mainly in molar excess of peroxynitrite. In agreement with this hypothesis, oxyhemoglobin strongly inhibited tyrosine nitration of a target dipeptide (Ala-Tyr) and of membrane proteins from ghosts resealed with oxyhemoglobin. Erythrocytes were poor inhibitors of Ala-Tyr nitration on account of the membrane barrier. However, at the physiologic hematocrit, Ala-Tyr nitration was reduced by 65%. This "sink" function was facilitated by the huge amount of band 3 anion exchanger on the cell membrane. We conclude that in blood oxyhemoglobin is a peroxynitrite scavenger of physiologic relevance.  相似文献   

20.
We have shown previously that peroxynitrite-induced nitration of a hydrophobic tyrosyl probe is greater than that of tyrosine in the aqueous phase (Zhang, H., Joseph, J., Feix, J., Hogg, N., and Kalyanaraman, B. (2001) Biochemistry 40, 7675-7686). In this study, we have tested the hypothesis that the extent of tyrosine nitration depends on the intramembrane location of tyrosyl probes and on the nitrating species. To this end, we have synthesized membrane spanning 23-mer containing a single tyrosyl residue at positions 4, 8, and 12. The location of the tyrosine residues in the phospholipid membrane was determined by fluorescence and electron spin resonance techniques. Nitration was initiated by slow infusion of peroxynitrite, co-generated superoxide and nitric oxide ((.)NO), or a myeloperoxidase/hydrogen peroxide/nitrite anion (MPO/H(2)O(2)/NO(2)(-)) system. Results indicate that with slow infusion of peroxynitrite, nitration of transmembrane tyrosyl peptides was much higher (10-fold or more) than tyrosine nitration in aqueous phase. Peroxynitrite-dependent nitration of tyrosyl-containing peptides increased with increasing depth of the tyrosyl residue in the bilayer. In contrast, MPO/H(2)O(2)/ NO(2)(-)-induced tyrosyl nitration decreased with increasing depth of tyrosyl residues in the membrane. Transmembrane nitrations of tyrosyl-containing peptides induced by both peroxynitrite and MPO/H(2)O(2)/NO(2)(-) were totally inhibited by (.)NO that was slowly released from spermine NONOate. Nitration of peptides in both systems was concentration-dependently inhibited by unsaturated fatty acid. Concomitantly, an increase in lipid oxidation was detected. A mechanism involving (.)NO(2) radical is proposed for peroxynitrite and MPO/H(2)O(2)/NO(2)(-)-dependent transmembrane nitration reactions.  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