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 15 毫秒
1.
Various cuticular regions from the desert locust, Schistocerca gregaria, were quantitatively analyzed for two cross-linking amino acids, dityrosine and trityrosine, characteristic constituents of the rubberlike cuticular protein, resilin. These amino acids were found in all regions of cuticle investigated, but in widely varying amounts. In fully mature adult locusts the largest amounts of di- and trityrosine were obtained from the prealar arms and wing-hinges, structures possessing long-range elasticity and being involved in energy storage in the flight system. In structures where deformations tend to occur more slowly, such as the clypeo-labral springs and tracheae, di- and trityrosine are less abundant. In sclerotized cuticle from femur and tibia, as well as in cornea and in the highly stretchable intersegmental membranes of mature females, they are only found in trace amounts and are probably unrelated to elasticity. The trityrosine-to-dityrosine ratio in the various cuticular regions vary from nearly equal amounts of the two amino acids to about ten times more dityrosine than trityrosine, indicating that the regions differ in degree of cross-linking; the tracheal wall is the material with the highest trityrosine-to-dityrosine ratio. In some cuticular regions the ratio increases during maturation from newly moulted (teneral) adults to reproductively active locusts; the most pronounced increase was observed for the wing-hinges, and only a small increase was observed for the abdominal tergal plates. In most cuticular regions in fifth instar locust nymphs the contents of di- and trityrosine corresponded to the contents measured for the adult cuticular regions, but only trace amounts of the two amino acids were obtained from the region of the nymphal wing base which corresponds to the wing-hinge containing cuticular region in adult locusts.  相似文献   

2.
B Gmeiner  C Seelos 《FEBS letters》1989,255(2):395-397
Treatment of L-tyrosine in a peroxidase/H2O2 system results in the formation of dityrosine. However, the phosphoester derivative of tyrosine, O-phospho-L-tyrosine, was unable to form dityrosine even in mixtures with free L-tyrosine. Dephosphorylation of O-phospho-L-tyrosine by alkaline phosphatase followed by horseradish peroxidase/H2O2 treatment resulted in the formation of dityrosine. Our in vitro results indicate that phosphorylation/dephosphorylation of L-tyrosine may regulate dityrosine formation, and is supposed to play an important role in protein-protein interactions, i.e. cross-linking.  相似文献   

3.
Han Q  Li G  Li J 《Biochimica et biophysica acta》2000,1523(2-3):246-253
A specific chorion peroxidase is present in Aedes aegypti and this enzyme is responsible for catalyzing chorion protein cross-linking through dityrosine formation during chorion hardening. Peroxidase-mediated dityrosine cross-linking requires H(2)O(2), and this study discusses the possible involvement of the chorion peroxidase in H(2)O(2) formation by mediating NADH/O(2) oxidoreduction during chorion hardening in A. aegypti eggs. Our data show that mosquito chorion peroxidase is able to catalyze pH-dependent NADH oxidation, which is enhanced in the presence of Mn(2+). Molecular oxygen is the electron acceptor during peroxidase-catalyzed NADH oxidation, and reduction of O(2) leads to the production of H(2)O(2), demonstrated by the formation of dityrosine in a NADH/peroxidase reaction mixture following addition of tyrosine. An oxidoreductase capable of catalyzing malate/NAD(+) oxidoreduction is also present in the egg chorion of A. aegypti. The cooperative roles of chorion malate/NAD(+)oxidoreductase and chorion peroxidase on generating H(2)O(2) with NAD(+) and malate as initial substrates were demonstrated by the production of dityrosine after addition of tyrosine to a reaction mixture containing NAD(+) and malate in the presence of both malate dehydrogenase fractions and purified chorion peroxidase. Data suggest that chorion peroxidase-mediated NADH/O(2) oxidoreduction may contribute to the formation of the H(2)O(2) required for chorion protein cross-linking mediated by the same peroxidase, and that the chorion associated malate dehydrogenase may be responsible for the supply of NADH for the H(2)O(2) production.  相似文献   

4.
Using pure tyrosine and phosphotyrosine we have recently shown that phosphotyrosine is unable to form peroxidase catalyzed dimers (1989, FEBS Lett. 255, 395-397). In the present report, the effect of phosphotyrosine residues within a protein structure on dityrosine formation was studied using casein as a model protein. Dephosphorylation of casein resulted in a dose and time dependent increased synthesis of dityrosines following treatment with peroxidase/H2O2. The extent of crosslink formation was inversely related to the amount of phosphorylated tyrosine residues as quantitated by immunoblotting. Thus, phosphorylation of tyrosine residues could play a regulatory role in protein-crosslinking where dityrosine bonds are involved.  相似文献   

5.
Sperm whale metmyoglobin, which has tyrosine residues at positions 103, 146, and 151, dimerizes in the presence of H2O2. Equine metmyoglobin, which lacks Tyr-151, and red kangaroo metmyoglobin, which lacks Tyr-103 and Tyr-151, do not dimerize in the presence of H2O2. The dityrosine content of the sperm whale myoglobin dimer shows that it is primarily held together by dityrosine cross-links, although more tyrosine residues are lost than are accounted for by dityrosine formation. Digestion of the myoglobin dimer with chymotrypsin yields a peptide with the fluorescence spectrum of dityrosine. The amino acid composition, amino acid sequence, and mass spectrum of the peptide show that cross-linking involves covalent bond formation between Tyr-103 of one myoglobin chain and Tyr-151 of the other. Replacement of the prosthetic group of sperm whale myoglobin with zinc protoporphyrin IX prevents H2O2-induced dimerization even when intact horse metmyoglobin is present in the incubation. This suggests that the tyrosine radicals required for the dimerization reaction are generated by intra- rather than intermolecular electron transfer to the ferryl heme. Rapid electron transfer from Tyr-103 to the ferryl heme followed by slower electron transfer from Tyr-151 to Tyr-103 is most consistent with the present results.  相似文献   

6.
Eosinophil peroxidase (EPO) has been implicated in promoting oxidative tissue injury in conditions ranging from asthma and other allergic inflammatory disorders to cancer and parasitic/helminthic infections. Studies thus far on this unique peroxidase have primarily focused on its unusual substrate preference for bromide (Br(-)) and the pseudohalide thiocyanate (SCN(-)) forming potent hypohalous acids as cytotoxic oxidants. However, the ability of EPO to generate reactive nitrogen species has not yet been reported. We now demonstrate that EPO readily uses nitrite (NO(2)(-)), a major end-product of nitric oxide ((.)NO) metabolism, as substrate to generate a reactive intermediate that nitrates protein tyrosyl residues in high yield. EPO-catalyzed nitration of tyrosine occurred more readily than bromination at neutral pH, plasma levels of halides, and pathophysiologically relevant concentrations of NO(2)(-). Furthermore, EPO was significantly more effective than MPO at promoting tyrosine nitration in the presence of plasma levels of halides. Whereas recent studies suggest that MPO can also promote protein nitration through indirect oxidation of NO(2)(-) with HOCl, we found no evidence that EPO can indirectly mediate protein nitration by a similar reaction between HOBr and NO(2)(-). EPO-dependent nitration of tyrosine was modulated over a physiologically relevant range of SCN(-) concentrations and was accompanied by formation of tyrosyl radical addition products (e.g. o,o'-dityrosine, pulcherosine, trityrosine). The potential role of specific antioxidants and nucleophilic scavengers on yields of tyrosine nitration and bromination by EPO are examined. Thus, EPO may contribute to nitrotyrosine formation in inflammatory conditions characterized by recruitment and activation of eosinophils.  相似文献   

7.
Protoporphyrin-sensitized photooxidation in human red blood cell membranes leads to severe deterioration of membrane structure and function. The membrane damage is caused by direct oxidation of amino acid residues, with subsequent cross-linking of membrane proteins. The chemical nature of these cross-links was studied in model systems, isolated spectrin and red cell ghosts. Cysteine and methionine are not involved in the cross-linking reaction. Further it could be shown that dityrosine formation, the crucial mechanism in oxidative cross-linking of proteins by peroxidase-H2O2 treatment, plays no role in photodynamic cross-linking. Experimental evidence indicated that a secondary reaction between free amino groups and a photooxidation product of histidine, tyrosine or tryptophan is involved in photodynamic cross-linking. This was deduced from the reaction observed between compounds containing a free amino group and photooxidation products of these amino acids, both in model systems, isolated spectrin and erythrocyte ghosts. In accordance, succinylation of free amino groups of membrane proteins or addition of compounds with free amino groups protected against cross-linking. Quantitative data and consideration of the reaction mechanisms of photodynamic oxidation of amino acids make it highly probable that an oxidation product of histidine rather than of tyrosine or tryptophan is involved in the cross-linking reaction, via a nucleophilic addition by free amino groups.  相似文献   

8.
The mechanisms by which ligand-stimulated generation of reactive oxygen species in nonphagocytic cells mediate biologic effects are largely unknown. The profibrotic cytokine, transforming growth factor-beta1 (TGF-beta1), generates extracellular hydrogen peroxide (H2O2) in contrast to intracellular reactive oxygen species production by certain mitogenic growth factors in human lung fibroblasts. To determine whether tyrosine residues in fibroblast-derived extracellular matrix (ECM) proteins may be targets of H2O2-mediated dityrosine-dependent cross-linking reactions in response to TGF-beta1, we utilized fluorophore-labeled tyramide, a structurally related phenolic compound that forms dimers with tyrosine, as a probe to detect such reactions under dynamic cell culture conditions. With this approach, a distinct pattern of fluorescent labeling that seems to target ECM proteins preferentially was observed in TGF-beta1-treated cells but not in control cells. This reaction required the presence of a heme peroxidase and was inhibited by catalase or diphenyliodonium (a flavoenzyme inhibitor), similar to the effect on TGF-beta1-induced dityrosine formation. Exogenous addition of H2O2 to control cells that do not release extracellular H2O2 produced a similar fluorescent labeling reaction. These results support the concept that, in the presence of heme peroxidases in vivo, TGF-beta1-induced H2O2 production by fibroblasts may mediate oxidative dityrosine-dependent cross-linking of ECM protein(s). This effect may be important in the pathogenesis of human fibrotic diseases characterized by overexpression/activation of TGF-beta1.  相似文献   

9.
The sheath or second-molt cuticle (2M) was isolated from in vitro exsheathed Haemonchus contortus infective larvae (L3[2M]). Acid hydrolysates of 2-mercaptoethanol (2ME)-soluble and 2ME-insoluble cuticular proteins were analyzed by high performance liquid chromatography for tyrosine-derived cross-linking amino acids. Dityrosine and isotrityrosine were identified by their chromatographic behavior, absorbance spectra, and other chemical characteristics in both the 2ME-soluble and 2ME-insoluble fractions. Dityrosine and isotrityrosine were found in greater amounts in the 2ME-insoluble proteins. When intact 2M cuticles were labeled with 125I prior to acid hydrolysis, radiolabel was recovered in tyrosine but not dityrosine or isotrityrosine indicating that the tyrosine cross-links are not susceptible to iodination in the intact protein. The results are consistent with a hypothesis that tyrosine-derived cross-links are important components of H. contortus 2M cuticular proteins.  相似文献   

10.
We have previously reported that amyloid Abeta, the major component of senile plaques in Alzheimer's disease (AD), binds Cu with high affinity via histidine and tyrosine residues [Atwood, C. S., et al. (1998) J. Biol. Chem. 273, 12817-12826; Atwood, C. S., et al. (2000) J. Neurochem. 75, 1219-1233] and produces H(2)O(2) by catalyzing the reduction of Cu(II) or Fe(III) [Huang, X., et al. (1999) Biochemistry 38, 7609-7616; Huang, X., et al. (1999) J. Biol. Chem. 274, 37111-37116]. Incubation with Cu induces the SDS-resistant oligomerization of Abeta [Atwood, C. S., et al. (2000) J. Neurochem. 75, 1219-1233], a feature characteristic of neurotoxic soluble Abeta extracted from the AD brain. Since residues coordinating Cu are most vulnerable to oxidation, we investigated whether modifications of these residues were responsible for Abeta cross-linking. SDS-resistant oligomerization of Abeta caused by incubation with Cu was found to induce a fluorescence signal characteristic of tyrosine cross-linking. Using ESI-MS and a dityrosine specific antibody, we confirmed that Cu(II) (at concentrations lower than that associated with amyloid plaques) induces the generation of dityrosine-cross-linked, SDS-resistant oligomers of human, but not rat, Abeta peptides. The addition of H2O2 strongly promoted Cu-induced dityrosine cross-linking of Abeta1-28, Abeta1-40, and Abeta1-42, suggesting that the oxidative coupling is initiated by interaction of H2O2 with a Cu(II) tyrosinate. The dityrosine modification is significant since it is highly resistant to proteolysis and is known to play a role in increasing structural strength. Given the elevated concentration of Cu in senile plaques, our results suggest that Cu interactions with Abeta could be responsible for causing the covalent cross-linking of Abeta in these structures.  相似文献   

11.
High molecular weight homologues of gp91phox, the superoxide-generating subunit of phagocyte nicotinamide adenine dinucleotide phosphate (NADPH)-oxidase, have been identified in human (h) and Caenorhabditis elegans (Ce), and are termed Duox for "dual oxidase" because they have both a peroxidase homology domain and a gp91phox domain. A topology model predicts that the enzyme will utilize cytosolic NADPH to generate reactive oxygen, but the function of the ecto peroxidase domain was unknown. Ce-Duox1 is expressed in hypodermal cells underlying the cuticle of larval animals. To investigate function, RNA interference (RNAi) was carried out in C. elegans. RNAi animals showed complex phenotypes similar to those described previously in mutations in collagen biosynthesis that are known to affect the cuticle, an extracellular matrix. Electron micrographs showed gross abnormalities in the cuticle of RNAi animals. In cuticle, collagen and other proteins are cross-linked via di- and trityrosine linkages, and these linkages were absent in RNAi animals. The expressed peroxidase domains of both Ce-Duox1 and h-Duox showed peroxidase activity and catalyzed cross-linking of free tyrosine ethyl ester. Thus, Ce-Duox catalyzes the cross-linking of tyrosine residues involved in the stabilization of cuticular extracellular matrix.  相似文献   

12.
The genetic addition of hexahistidine (H(6)) tags is widely used to isolate recombinant proteins by immobilized metal-affinity chromatography (IMAC). Addition of a tyrosine residue to H(6) tags enabled proteins to be covalently cross-linked under mild conditions in a manner similar to the natural, site-specific cross-linking of tyrosines into dityrosine. A series of seven hexahistidine tags with tyrosines placed in various positions (H(6)Y tags) were added to the amino terminus of the I28 immunoglobulin domain of the human cardiac titin. The H(6)Y-tagged I28 dimerized in the presence of excess Ni(2+) with a K(D) of 200 microM. Treatment of Ni(2+)-dimerized H(6)Y-I28 with an oxidant, monoperoxyphthalic acid (MMPP) or sodium sulfite, resulted in covalent protein multimerization through chelated Ni(2+)-catalyzed cross-linking of the Y residues engineered into the H(6) tag. The protein oligomerization was observed by sodium dodecyl sulfate polyacrylamide gel electrophoresis (SDS PAGE). The presence of dityrosine in the cross-linked proteins was confirmed by fluorescence emission at 410 nm. Proteins lacking the Y residue in the H(6) tag treated with the same oxidative conditions did not cross-link or exhibit dityrosine fluorescence, despite the presence of an endogenous Y residue. The method may have potential uses in other protein conjugation applications such as protein labeling and interfacial immobilization of proteins on artificial surfaces.  相似文献   

13.
The proteins in the distensible alloscutal cuticle of the blood-feeding tick, Ixodes ricinus, have been characterized by electrophoresis and chromatography, two of the proteins were purified and their total amino acid sequence determined. They show sequence similarity to cuticular proteins from the spider, Araneus diadematus, and the horseshoe crab, Limulus polyphemus, and to a lesser extent to insect cuticular proteins. They contain a conserved sequence region, which is closely related to the chitin-binding Rebers-Riddiford consensus sequence present in many insect cuticular proteins. Only a fraction of the alloscutal proteins can be readily dissolved, and the dissolved proteins are difficult to separate by electrophoresis and column chromatography. The insoluble fraction can only be dissolved after degradation to smaller peptides. The mixture of extractable proteins as well as hydrolysates of the insoluble fraction are fluorescent when exposed to ultraviolet light, and the fluorescence corresponds in excitation and emission maxima to the fluorescence of the rubber-like arthropodan protein, resilin, and to the amino acid dityrosine. Small amounts of dityrosine were obtained from ticks in the early phase of a blood meal when the cuticle weighs less than 4 mg; increasing amounts were obtained from animals in the initial period of feeding, during which the cuticular weight increases from 4 to 11 mg, whereas little increase in dityrosine content was observed during the final period of engorgement. Cuticle from fully distended ticks contains about 60-80 nmole dityrosine per tick, corresponding to 2-3 microg/mg cuticle. It is suggested that the major part of the cuticular proteins is made inextractable by cross-linking by dityrosine residues, and that dityrosine plays a role in stabilizing the cuticular structure during the extensive distension occurring during a blood meal. Small amounts of 3-monochlorotyrosine and 3,5-dichlorotyrosine were obtained from the distended tick cuticle, corresponding to chlorination of between 0.5% and 1.5% of the tyrosine residues. It is suggested that the chlorotyrosines are a side-product of oxidative processes in the cuticle.  相似文献   

14.
1. Fourteen peptides were isolated from the egg jelly of sea urchins, Pseudocentrotus depressus, Strongylocentrotus purpuratus, Hemicentrotus pulcherrimus and Anthocidaris crassispina and their amino acid sequences were determined. 2. The peptides stimulated H. pulcherrimus sperm respiration one half-maximally at about 8-60 pM. 3. Addition of speract to intact spermatozoa of P. depressus, H. pulcherrimus and A. crassispina resulted in the appearance of a newly stained protein (Mr 128,000 for P. depressus, Mr 128,000 for H. pulcherrimus and Mr 131,000 for A. crassispina) on sodium dodecyl sulfate-polyacrylamide gels.  相似文献   

15.
Dityrosine formation leads to the cross-linking of proteins intra- or intermolecularly. The formation of dityrosine in lens proteins oxidized by metal-catalyzed oxidation (MCO) systems was estimated by chemical and immunochemical methods. Among the four MCO systems examined (H(2)O(2)/Cu, H(2)O(2)/Fe-ethylenediaminetetraacetic acid (Fe-EDTA), ascorbate/Cu, ascorbate/Fe-EDTA), the treatment with H(2)O(2)/Cu preferentially caused dityrosine formation in the lens proteins. The success of oxidative protein modification with all the MCO systems was confirmed by carbonyl formation estimated using 2,4-dinitrophenylhydrazine. The loss of tyrosine by the MCO systems was partly due to the formation of protein-bound 3,4-dihydroxyphenylalanine. The formation of dityrosine specific to H(2)O(2)/Cu was confirmed by using poly-(Glu, Ala, Tyr) and N-acetyl-tyrosine as a substrate. The dissolved oxygen concentration in the H(2)O(2)/Cu system hardly affected the amount of dityrosine formation, suggesting that dityrosine generation by the H(2)O(2)/Cu system is independent of oxygen concentration. Moreover, the combination of copper ion with H(2)O(2) is the most effective system for dityrosine formation among various metal ions examined. The addition of reducing agents, glutathione or ascorbic acid, into the H(2)O(2)/Cu system suppressed the generation of the dityrosine moiety, suggesting effective quench of tyrosyl radicals by the reducing agents.  相似文献   

16.
Ovoperoxidase, an enzyme secreted by the eggs of the sea urchin Stronglycocentrotus purpuratus upon activation, catalyzes the formation of dityrosine residues in the fertilization envelope. This cross-linking reaction requires extracellular H2O2, which is produced by the egg during the cyanide-insensitive "respiratory burst" of fertilization. While investigating the possibility that the sea urchin oxidase might generate O2- as a precursor to H2O2, we discovered that ovoperoxidase possessed O2- degrading activity. Ovoperoxidase catalyzed the breakdown of O2- in a reaction that was sensitive to inhibition by catalase, indicating a requirement for H2O2. High concentrations of either O2- or H2O2 inhibited the O2- degrading activity of ovoperoxidase, as did the peroxidase inhibitors aminotriazole, azide, and phenylhydrazine. When ovoperoxidase was heated at 56 degrees C, it lost O2- degrading activity in parallel with peroxidase activity. In contrast, the copper-chelating agent diethyldithiocarbamate, which completely inactivated CuZn superoxide dismutase, failed to affect ovoperoxidase. The requirement for H2O2 and the inhibition by aminotriazole, azide, and phenylhydrazine support the hypothesis that ovoperoxidase catalyzes the breakdown of O2- by a peroxidative mechanism. Ovoperoxidase may play a role in protecting the developing embryo from oxidants derived from O2-.  相似文献   

17.
In vitro experiments are reported showing that the dipeptides Tyr-L-Arg (kyotorphin) and Tyr-D-Arg (D-Arg-kyotorphin) can be oxidized by H2O2-horseradish peroxidase system: the products formed are characterized by absorption spectra with two peaks at 290 nm and 315 nm. The effects of substrate and enzyme concentration on the oxidation rate are described. Amino acid analysis of hydrolysates of peroxidase-treated kyotorphins provides evidence for the presence of dityrosine. The data suggest that the oxidation leads to the production of dimers with an o,o-linkage between the tyrosine residues.  相似文献   

18.
Protein modification methods represent fundamental techniques that are applicable in many fields. In this study, a site-specific protein cross-linking based on the oxidative tyrosine coupling reaction was demonstrated. In the presence of horseradish peroxidase (HRP) and H(2)O(2), tyrosine residues undergo one-electron oxidation reactions and form radicals in their phenolic moieties, and these species subsequently react with each other to form dimers or further react to generate polymers. Here, a peptide-tag containing a tyrosine residue(s) (Y-tag, of which the amino acid sequences were either GGGGY or GGYYY) was genetically introduced at the C-terminus of a model protein, Escherichia coli alkaline phosphatase (BAP). Following the incubation of recombinant BAPs with HRP and H(2)O(2), Y-tagged BAPs were efficiently cross-linked with each other, whereas wild-type BAP did not undergo cross-linking, indicating that the tyrosine residues in the Y-tags were recognized by HRP as the substrates. To determine the site-specificity of the cross-linking reaction, the Y-tag was selectively removed by thrombin digestion. The resultant BAP without the Y-tag showed no reactivity in the presence of HRP and H(2)O(2). Conversely, Y-tagged BAPs cross-linked by HRP treatment were almost completely digested into monomeric BAP units following incubation with the protease. Moreover, cross-linked Y-tagged BAPs retained ~95% of their native enzymatic activity. These results show that HRP catalyzed the site-specific cross-linking of BAPs through tyrosine residues positioned in the C-terminal Y-tag. The site-selective enzymatic oxidative tyrosine coupling reaction should offer a practical option for site-specific and covalent protein modifications.  相似文献   

19.
The effect of bicarbonate anion (HCO(3)(-)) on the peroxidase activity of copper, zinc superoxide dismutase (SOD1) was investigated using three structurally different probes: 5, 5'-dimethyl-1-pyrroline N-oxide (DMPO), tyrosine, and 2, 2'-azino-bis-[3-ethylbenzothiazoline]-6-sulfonic acid (ABTS). Results indicate that HCO(3)(-) enhanced SOD/H(2)O(2)-dependent (i) hydroxylation of DMPO to DMPO-OH as measured by electron spin resonance, (ii) oxidation and nitration of tyrosine to dityrosine, nitrotyrosine, and nitrodityrosine as measured by high pressure liquid chromatography, and (iii) oxidation of ABTS to the ABTS cation radical as measured by UV-visible spectroscopy. Using oxygen-17-labeled water, it was determined that the oxygen atom present in the DMPO-OH adduct originated from H(2)O and not from H(2)O(2). This result proves that neither free hydroxyl radical nor enzyme-bound hydroxyl radical was involved in the hydroxylation of DMPO. We postulate that HCO(3)(-) enhances SOD1 peroxidase activity via formation of a putative carbonate radical anion. This new and different perspective on HCO(3)(-)-mediated oxidative reactions of SOD1 may help us understand the free radical mechanism of SOD1 and related mutants linked to amyotrophic lateral sclerosis.  相似文献   

20.
Filamentous inclusions of alpha-synuclein protein are hallmarks of neurodegenerative diseases collectively known as synucleinopathies. Previous studies have shown that exposure to oxidative and nitrative species stabilizes alpha-synuclein filaments in vitro, and this stabilization may be due to dityrosine cross-linking. To test this hypothesis, we mutated tyrosine residues to phenylalanine and generated recombinant wild type and mutant alpha-synuclein proteins. alpha-Synuclein proteins lacking some or all tyrosine residues form fibrils to the same extent as the wild type protein. Tyrosine residues are not required for protein cross-linking or filament stabilization resulting from transition metal-mediated oxidation, because higher Mr SDS-resistant oligomers and filaments stable to chaotropic agents are detected using all Tyr --> Phe alpha-synuclein mutants. By contrast, cross-linking resulting from exposure to nitrating agents required the presence of one or more tyrosine residues. Furthermore, tyrosine cross-linking is involved in filament stabilization, because nitrating agent-exposed assembled wild type, but not mutant alpha-synuclein lacking all tyrosine residues, was stable to chaotropic treatment. In addition, the formation of stable alpha-synuclein inclusions in intact cells after exposure to oxidizing and nitrating species requires tyrosine residues. These findings demonstrate that nitrative and/or oxidative stress results in distinct mechanisms of alpha-synuclein protein modifications that can influence the formation of stable alpha-synuclein fibrils.  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