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 375 毫秒
1.
During autoxidation of 1,4-hydroquinone (H2Q, less than 1 mM) at pH 7.4 and 37 degrees C, stoichiometric amounts of 1,4-benzoquinone (Q) and hydrogen peroxide were formed during the initial reaction. The reaction kinetics showed a significant induction period which was abolished by minute amounts of Q. Hydrogen peroxide and catalase were without effect on the autoxidation process. Transition metals apparently were not involved, since chelators like EDTA, DETAPAC, and desferrioxamine or FeSO4 had no influence on the autoxidation kinetics. Superoxide dismutase (SOD) did not abolish the induction period but dramatically enhanced the autoxidation rate by more than two orders of magnitude. The stimulatory effect was first-order in SOD concentration but showed saturation kinetics. The dependence of Q and hydrogen peroxide formation rates on H2Q concentration shows a biphasic behaviour: dependence on the square at low H2Q, but on the square root at high H2Q concentration. As revealed by calculatory simulations the results can be adequately described by the known reaction rate constants. The reaction starts with the comproportionation of H2Q and Q to yield two semiquinone molecules which autoxidize to give two superoxide radicals and two molecules of Q which enter into a new cycle of comproportionation. Because of unfavourable equilibria the autocatalytic reaction soon comes to steady state, and the further reaction is governed by the rate of superoxide removal. At excess SOD, the comproportionation reaction is rate-limiting, thus explaining the saturation effects of SOD. The experiments do not allow a decision between the two functions of SOD; the conventional action as a superoxide:superoxide oxidoreductase or as a semiquinone:superoxide oxidoreductase. In the latter reaction SOD is thought to be reduced by semiquinone with Q formation. In the second step the reduced enzyme would be re-oxidized by a superoxide radical which is formed during autoxidation of the second semiquinone molecule generated in the comproportionation reaction. From thermodynamic considerations, the latter function of SOD appears to be plausible.  相似文献   

2.
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   

3.
The Na+,K(+)-ATPase is a membrane-bound, sulfhydryl-containing protein whose activity is critical to maintenance of cell viability. The susceptibility of the enzyme to radical-induced membrane lipid peroxidation was determined following incorporation of a purified Na+,K(+)-ATPase into soybean phosphatidylcholine liposomes. Treatment of liposomes with Fenton's reagent (Fe2+/H2O2) resulted in malondialdehyde formation and total loss of Na+,K(+)-ATPase activity. At 150 microM Fe2+/75 microM H2O2, vitamin E (5 mol%) totally prevented lipid peroxidation but not the loss of enzyme activity. Lipid peroxidation initiated by 25 microM Fe2+/12.5 microM H2O2 led to a loss of Na+,K(+)-ATPase activity, however, vitamin E (1.2 mol%) prevented both malondialdehyde formation and loss of enzyme activity. In the absence of liposomes, there was complete loss of Na+,K(+)-ATPase activity in the presence of 150 microM Fe2+/75 microM H2O2, but little effect by 25 microM Fe2+/12.5 microM H2O2. The activity of the enzyme was also highly sensitive to radicals generated by the reaction of Fe2+ with cumene hydroperoxide, t-butylhydroperoxide, and linoleic acid hydroperoxide. Lipid peroxidation initiated by 150 microM Fe2+/150 microM Fe3+, an oxidant which may be generated by the Fenton's reaction, inactivated the enzyme. In this system, inhibition of malondialdehyde formation by vitamin E prevented loss of Na+,K(+)-ATPase activity. These data demonstrate the susceptibility of the Na+,K(+)-ATPase to radicals produced during lipid peroxidation and indicate that the ability of vitamin E to prevent loss of enzyme activity is highly dependent upon both the nature and the concentration of the initiating and propagating radical species.  相似文献   

4.
N R Matheson  J Travis 《Biochemistry》1985,24(8):1941-1945
Human alpha 1-proteinase inhibitor is easily susceptible to inactivation because of the presence of a methionyl residue at its reactive site. Thus, oxidizing species derived from the myeloperoxidase system (enzyme, H2O2, and C1-), as well as hypochlorous acid, can inactivate this inhibitor, although H2O2 alone has no effect. Butylated hydroxytoluene, a radical scavenger, partially protects alpha 1-proteinase inhibitor from the myeloperoxidase system and completely protects it from hypochlorous acid. Each oxidant also reacts differently with the inhibitor, in that the myeloperoxidase system and hypochlorous acid can each oxidize as many as six methionyl residues, but hypochlorous acid can also oxidize a single tyrosine residue. Myeloperoxidase can be inactivated by hypochlorous acid, by autoxidation in the presence of H2O2 and C1-, as well as by H2O2 alone. Butylated hydroxytoluene completely protects this enzyme from hypochlorous acid inactivation, does not affect the action of H2O2, and enhances autoinactivation. As many as six methionyl residues and two tyrosine residues could be oxidized during autoxidation and six methionine residues by H2O2 alone. Eight methionine residues and one tyrosine residue could be oxidized by hypochlorous acid. The tyrosine residue in myeloperoxidase was oxidized only at a relatively high concentration (600 microM) of hypochlorous acid at which point the enzyme simultaneously and completely lost its enzymatic activity. Loss of activity of myeloperoxidase could also be correlated with the loss of the heme groups present in the enzyme when a relatively high concentration of hypochlorous acid (600 microM) was used and also during autoxidation. It appears that once there is sufficient oxidant to modify one of the tyrosine residues, the heme group itself becomes susceptible.  相似文献   

5.
Role of hydrogen peroxide in sperm capacitation and acrosome reaction   总被引:12,自引:0,他引:12  
The generation of reactive oxygen species (ROS) has been implicated in the regulation of sperm capacitation and acrosome reaction; however, the mechanisms underlying this regulation remain unclear. To examine the cellular processes involved, we studied the effect of different concentrations of hydrogen peroxide (H(2)O(2)) on protein tyrosine phosphorylation under various conditions. Treatment of spermatozoa with H(2)O(2) in medium without heparin caused a time- and dose-dependent increase in protein tyrosine phosphorylation of at least six proteins in which maximal effect was seen after 2 h of incubation with 50 microM H(2)O(2). At much higher concentrations of H(2)O(2) (0.5 mM), there is significant reduction in the phosphorylation level, and no protein tyrosine phosphorylation is observed at 5 mM H(2)O(2) after 4 h of incubation. Exogenous NADPH enhanced protein tyrosine phosphorylation similarly to H(2)O(2). These two agents, but not heparin, induced Ca(2+)-dependent tyrosine phosphorylation of an 80-kDa protein. Treatment with H(2)O(2) (50 microM) caused approximately a twofold increase in cAMP, which is comparable to the effect of bicarbonate, a known activator of soluble adenylyl cyclase in sperm. This report suggests that relatively low concentrations of H(2)O(2) are beneficial for sperm capacitation, but that too high a concentration inhibits this process. We also conclude that H(2)O(2) activates adenylyl cyclase to produce cAMP, leading to protein kinase A-dependent protein tyrosine phosphorylation.  相似文献   

6.
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   

7.
Hydrogen peroxide (H(2)O(2)) triggers a redox cycle between ferric and ferryl hemoglobin (Hb) leading to the formation of a transient protein radical and a covalent hemeprotein cross-link. Addition of H(2)O(2) to highly purified human hemoglobin (HbA(0)) induced structural changes that primarily resided within beta subunits followed by the internalization of the heme moiety within alpha subunits. These modifications were observed when an equal molar concentration of H(2)O(2) was added to HbA(0) yet became more abundant with greater concentrations of H(2)O(2). Mass spectrometric and amino acid analysis revealed for the first time that betaCys-93 and betaCys-112 were oxidized extensively and irreversibly to cysteic acid when HbA(0) was treated with H(2)O(2). Oxidation of further amino acids in HbA(0) exclusive to the beta-globin chain included modification of betaTrp-15 to oxyindolyl and kynureninyl products as well as betaMet-55 to methionine sulfoxide. These findings may therefore explain the premature collapse of the beta subunits as a result of the H(2)O(2) attack. Analysis of a tryptic digest of the main reversed phase-high pressure liquid chromatography fraction revealed two alpha-peptide fragments (alpha128-alpha139) and a heme moiety with the loss of iron, cross-linked between alphaSer-138 and the porphyrin ring. The novel oxidative pathway of HbA(0) modification detailed here may explain the diverse oxidative, toxic, and potentially immunogenic effects associated with the release of hemoglobin from red blood cells during hemolytic diseases and/or when cell-free Hb is used as a blood substitute.  相似文献   

8.
Hydrogen peroxide (H2O2) (10-100 microM) induced rapid and transient accumulation of phosphatidic acid (PA) in suspension-cultured rice cells. When phospholipase activity in the cellular extract fraction prepared from rice cells treated with H2O2 was assayed in the presence of 1-butanol (0.1%), rapid and transient phosphatidylbutanol (PtdBut) formation was observed. Thus, the H2O2-activated phospholipase was concluded to be phospholipase D (PLD). Furthermore, H2O2 directly induced in vitro PLD activation in the cytosolic fraction without H2O2 treatment. In vitro and in vivo activation of PLD were completely suppressed in the presence of lavendustin A (0.05 mM), a potent inhibitor of protein tyrosine kinase. Phytoalexin biosynthesis induced by N-acetylchitooligosaccharide elicitor was enhanced in the presence of H2O2 (10-100 microM), whereas it was suppressed in the presence of tiron, a potent scavenger of O2-, 1-butanol (0.1%) and lavendustin A (0.05 mM). These results indicate that H2O2-inducible PLD activation enhances signal transduction leading to phytoalexin biosynthesis in rice cells.  相似文献   

9.
Ferrylmyoglobin-catalyzed linoleic acid peroxidation   总被引:1,自引:0,他引:1  
The addition of linoleic acid (18:2) to a solution containing oxymyoglobin (MbIIO2), metmyoglobin (MbIII), or metmyoglobin-azide complex (MbIII-N3-) resulted in the formation of a common complex with identical absorption spectral properties. The addition of H2O2 to a MbIII/linoleic acid mixture revealed a spectral profile with lambda max at 530 nm and different from that observed in the reaction of MbIII with H2O2 and identical to that of ferrylmyoglobin. This was accompanied by a progressive decrease in the absorption in the visible region, indicating heme degradation during the lipid peroxidation process. The oxidation products of linoleic acid during the MbIII/18:2/H2O2 interaction were assessed by HPLC under anaerobic and aerobic conditions. In both instances, the chromatograms at lambda 234 nm revealed the formation of a main peak with a retention time of 11.1 min, which cochromatographed with a standard of 9-hydroperoxide of linoleic acid. The latter adduct was not degraded by the oxoferryl complex of myoglobin. The conclusions originating from this research are two-fold. On the one hand, the identical spectral properties exhibited by the product originating from the reaction of either MbIIO2 or MbIII with linoleic acid bridge the apparent discrepancy between the different reactivities of MbIIO2 and MbIII toward H2O2 and their ability to promote lipid peroxidation. On the other hand, the pattern of oxidation products of linoleic acid observed during the MbIII/H2O2 interaction, i.e., the formation of a 9-hydroperoxide adduct as a major product, points to a specific binding character and a regioselectivity of the oxoferryl complex in the oxidation of unsaturated fatty acids or a catalytic preference for decomposition of the various isomeric hydroperoxides over that of the 9-hydroperoxide.  相似文献   

10.
To clarify the effect of superoxide dismutase (SOD) on the formation of hydroxyl radical in a standard reaction mixture containing 15 microM of xanthone, 0.1 M of 5,5-dimethyl-1-pyrroline N-oxide (DMPO), and 45 mM of phosphate buffer (pH 7.4) under UVA irradiation, electron paramagnetic resonance (EPR) measurements were performed. SOD enhanced the formation of hydroxyl radicals. The formation of hydroxyl radicals was inhibited on the addition of catalase. The rate of hydroxyl radical formation also slowed down under a reduced oxygen concentration, whereas it was stimulated by disodium ethylenediaminetetraacetate (EDTA) and diethyleneaminepentaacetic acid (DETAPAC). Above findings suggest that O(2), H(2)O(2), and iron ions participate in the reaction. SOD possibly enhances the formation of the hydroxyl radical in reaction mixtures of photosensitizers that can produce O(2)(-.).  相似文献   

11.
Tyrosine hydroxylation was studied in intact cells of mouse neuroblastoma clone N1E-115 which have high levels of tyrosine 3-monooxygenase (EC 1.14.16.2) and which have been fully characterized for tyrosine transport. Measurement of [3H]OH formed from L-[3,5(-3)H]tyrosine in the medium was the method of assay and [3H]OH formed was stoichiometric with the formation of L-[3H]3,4-dihydroxyphenylalanine. Tyrosine hydroxylation was dependent on time of incubation, cell number, and the concentration of [3H]tyrosine in the medium. From velocity vs. [3H]tyrosine concentration experiments, two apparent Km values were obtained: Km1 = 10 +/- 2 microM; Km2 = 140 +/- 10 microM. Substrate inhibition occurred with tyrosine concentrations between 20 and 50 microM. The reaction was twice as fast at pH 5.5 as at pH 7.4. alpha,alpha'-Dipyridyl (1 mM) caused major inhibition (75%) when [3H]tyrosine concentration was 10 microM. L-3-Iodotyrosine was a competitive inhibitor with Ki = 0.3 microM. Dopamine was a non-competitive inhibitor with Ki = 500 microM. 1-Norepinephrine had no effect. These results show that the hydroxylation of tyrosine by living N1E-115 cells has many of the properties of the reaction catalyzed by purified tyrosine 3-monooxygenase from normal tissue.  相似文献   

12.
Y Zick  R Sagi-Eisenberg 《Biochemistry》1990,29(44):10240-10245
Treatment of four cell lines [rat hepatoma (Fao), murine muscle (BC3H-1), Chinese hamster ovary (CHO), and rat basophilic leukemia (RBL)] with a combination of 3 mM H2O2 and 1 mM sodium orthovanadate markedly stimulates protein tyrosine phosphorylation, which is accompanied by a dramatic increase (5-15-fold) in inositol phosphate (InsP) formation. H2O2/vanadate stimulate best formation of inositol triphosphate while their effects on the mono and di derivatives are more moderate. In the presence of 3 mM H2O2, both protein tyrosine phosphorylation and InsP formation are highly correlated and manifest an identical dose-response relationship for vanadate. Half-maximal and maximal effects are obtained at 30 and 100 microM, respectively. This stimulatory effect of H2O2/vanadate is not mimicked by other oxidants such as spermine, spermidine, KMnO4, and vitamin K3. In RBL cells, the kinetics of inositol triphosphate formation correlate with tyrosine phosphorylation of a 67-kDa protein, while tyrosine phosphorylation of a 55-kDa protein is closely correlated with both inositol monophosphate formation and serotonin secretion from these cells. Taken together, these results suggest a causal relationship between tyrosine phosphorylation triggered in a nonhormonal manner and polyphosphoinositide breakdown. Furthermore, these results implicate protein tyrosine phosphorylation in playing a role in the stimulus-secretion coupling in RBL cells.  相似文献   

13.
Spectral scans in both the visible (650-450 nm) and the Soret (450-380 nm) regions were recorded for the native enzyme, Compound II, and Compound III of lactoperoxidase and thyroid peroxidase. Compound II for each enzyme (1.7 microM) was prepared by adding a slight excess of H2O2 (6 microM), whereas Compound III was prepared by adding a large excess of H2O2 (200 microM). After these compounds had been formed it was observed that they were slowly reconverted to the native enzyme in the absence of exogenous donors. The pathway of Compound III back to the native enzyme involved Compound II as an intermediate. Reconversion of Compound III to native enzyme was accompanied by the disappearance of H2O2 and generation of O2, with approximately 1 mol of O2 formed for each 2 mol of H2O2 that disappeared. A scheme is proposed to explain these observations, involving intermediate formation of the ferrous enzyme. According to the scheme, Compound III participates in a reaction cycle that effectively converts H2O2 to O2. Iodide markedly affected the interconversions between native enzyme, Compound II, and Compound III for lactoperoxidase and thyroid peroxidase. A low concentration of iodide (4 microM) completely blocked the formation of Compound II when lactoperoxidase or thyroid peroxidase was treated with 6 microM H2O2. When the enzymes were treated with 200 microM H2O2, the same low concentration of iodide completely blocked the formation of Compound III and largely prevented the enzyme degradation that otherwise occurred in the absence of iodide. These effects of iodide are readily explained by (i) the two-electron oxidation of iodide to hypoiodite by Compound I, which bypasses Compound II as an intermediate, and (ii) the rapid oxidation of H2O2 to O2 by the hypoiodite formed in the reaction between Compound I and iodide.  相似文献   

14.
Li G  Zhang H  Sader F  Vadhavkar N  Njus D 《Biochemistry》2007,46(23):6978-6983
At alkaline pH, 4-methylcatechol oxidizes more rapidly than the related catecholamines: dopamine, norepinephrine, and epinephrine. This oxidation is not inhibited by superoxide dismutase or catalase, indicating that O2 itself is the oxidant, but the reduction potential of O2/O2-* is too low for it to oxidize 4-methylcatechol directly. Instead, O2 oxidizes the 4-methylcatechol semiquinone, which is formed by comproportionation of 4-methylcatechol and its o-quinone. Aniline reacts very quickly with the o-quinone and thus stops the comproportionation reaction that oxidizes 4-methylcatechol to the semiquinone. Oxidation of 4-methylcatechol then requires superoxide, and in the presence of aniline, oxidation of 4-methylcatechol by O2 is inhibited by superoxide dismutase. When catecholamines oxidize, the side chain amine inserts into the catechol o-quinone, forming a bicyclic compound. By eliminating the quinone, this ring closure prevents comproportionation and the consequent oxidation of catecholamines by O2. It also prevents reaction of the quinone with other compounds and the formation of potentially toxic products.  相似文献   

15.
The antioxidant nordihydroguaiaretic acid (NDGA) has recently become well known as a putative anticancer drug. In this paper, it was evaluated the in vitro peroxynitrite (ONOO(-)), singlet oxygen ((1)O(2)), hydroxyl radical (OH(v)), hydrogen peroxide (H(2)O(2)), superoxide anion and hypochlorous acid (HOCl) scavenging capacity of NDGA. It was found that NDGA scavenges: (a) ONOO(-) (IC(50) = 4 +/- 0.94 microM) as efficiently as uric acid; (b) (1)O(2) (IC(50) = 151 +/- 20 microM) more efficiently than dimethyl thiourea, lipoic acid, N-acetyl-cysteine and glutathione; (c) OH(v) (IC(50) = 0.15 +/- 0.02 microM) more efficiently than dimethyl thiourea, uric acid, trolox, dimethyl sulfoxide and mannitol, (d) (IC(50) = 15 +/- 1 microM) more efficiently than N-acetyl-cysteine, glutathione, tempol and deferoxamine and (e) HOCl (IC(50) = 622 +/- 42 microM) as efficiently as lipoic acid and N-acetyl-cysteine. NDGA was unable to scavenge H(2)O(2). In an in vivo study in rats, NDGA was able to prevent ozone-induced tyrosine nitration in lungs. It is concluded that NDGA is a potent in vitro scavenger of ONOO(-), (1)O(2), OH(v), and HOCl and is able to prevent lung tyrosine nitration in vivo.  相似文献   

16.
H2O2 and vanadate are known insulinomimetic agents. Together they induce insulin's bioeffects with a potency which exceeds that seen with insulin, vanadate, or H2O2 alone. Employing Western blotting with anti-P-Tyr antibodies, we have identified in Fao cells at least four proteins (pp180, 150, 114, and 100) whose P-Tyr content is rapidly increased upon treatment of the cells with 3 mM H2O2. Tyrosine phosphorylation of these and additional proteins was markedly potentiated (6-10-fold) when 100 microM sodium orthovanadate was added together with H2O2. The effects of H2O2 and vanadate on protein tyrosine phosphorylation were rapid and specific. The enhanced tyrosine phosphorylation was accompanied by a concomitant inhibition of a cytosolic protein tyrosine phosphatase activity. The latter was inhibited by 50% in 3 mM H2O2-treated cells. The inhibitory effect was augmented in the combined presence of H2O2 and vanadate. Half- and maximal effects of vanadate were obtained at 15 microM and 1 mM, respectively. Vanadate (1 mM) alone, added to the cells, had only a trivial effect on protein tyrosine phosphatase activity. A 45-s challenge with insulin (10(-7) M) of cells pretreated with H2O2 largely mimicked the potentiating effects of vanadate on protein tyrosine phosphorylation but not on protein tyrosine phosphatase activity. Our results suggest the involvement of multiple tyrosine-phosphorylation proteins in mediating the biological effects of H2O2/vanadate. Their enhanced phosphorylation can be attributed at least in part, to the inhibitory effects exerted by H2O2 alone, or in combination with vanadate, on protein tyrosine phosphatase activity. The similarity between proteins phosphorylated in Fao cells in response to H2O2/vanadate or H2O2/insulin, suggests that either treatment stimulates protein tyrosine kinases having similar substrate specificities. The insulin receptor kinase is a likely candidate as its activity is markedly enhanced either by insulin (plus H2O2) or by H2O2/vanadate.  相似文献   

17.
DuBois JL  Klinman JP 《Biochemistry》2005,44(34):11381-11388
The copper amine oxidases (CAOs) catalyze the O(2)-dependent, two-electron oxidation of amines to aldehydes at an active site that contains Cu(II) and topaquinone (TPQ) cofactor. TPQ arises from the autocatalytic, post-translational oxidation of a tyrosine side chain in the active site. Monooxygenation within the ring of tyrosine at a single Cu(II) site is unique in biology and occurs as an early step in the formation of TPQ. The mechanism of this reaction has been further examined in the CAO from Hansenula polymorpha (HPAO). When a Clark electrode fitted to a custom-made, gastight apparatus over a range of initial concentrations of O(2) was used, rates of O(2) consumption at levels greater than air are seen to be reduced relative to earlier results, yielding K(D)(apparent) = 216 microM for O(2). This is consistent with a mechanism in which O(2) binds reversibly to the active site, triggering a conformational change that promotes ligation of tyrosinate to Cu(II). The activated Cu(II)-tyrosinate species has been proposed to react with O(2) in a rate-limiting step, although it was also possible that breakdown of a putative peroxy-intermediate controlled TPQ formation. To test the latter hypothesis, Cu(II)-free HPAO was prepared with 3,5-ring-[(2)H(2)]-tyrosine incorporated throughout the primary sequence. The absence of an isotope effect on the rate of TPQ formation eliminates cleavage of this C-H bond in a proposed Cu(II)-aryl-peroxide intermediate as a rate limiting step. The role of methionine 634, previously found to moderate O(2) binding during the catalytic cycle, is shown here to serve a similar function in TPQ formation. As with catalysis, the rate of TPQ formation correlates with the volume of the hydrophobic side chain at position 634, implicating similar binding sites for O(2) during catalysis and cofactor biogenesis.  相似文献   

18.
The comparative changes in the amino acid composition of calf skin collagen after gamma-irradiation (doses from 100 to 1,000 Gy) in aqueous solutions under different gas atmospheres (O2, N2O, H2, vacuum) were investigated. The radiochemical yields of collagen amino acid residues destruction were determined. Under O2 (OH X, O2-) most of amino acids are destroyed with higher yields than under N2O. Leucine, valine, isoleucine, phenylalanine, arginine were the exception because of their high reaction rate constants with OH X and hydroxylation reactions. Under H2 (e-aq, H) and in vacuum (e-aq, OH X) the mechanism of collagen radiolysis changed due to its aggregation; the destruction of those amino acids which have high reaction rate constants with water radiolysis products was mainly observed (phenylalanine, tyrosine, histidine).  相似文献   

19.
Upon exposure to hypotonic medium, skate red blood cells swell and then reduce their volume by releasing organic osmolytes and associated water. The regulatory volume decrease is inhibited by stilbenes and anion exchange inhibitors, suggesting involvement of the red blood cell anion exchanger skAE1. To determine the role of tyrosine phosphorylation, red blood cells were volume expanded with and without prior treatment with the tyrosine kinase inhibitor piceatannol. At the concentration used, 130 microM, piceatannol nearly completely inhibits p72(syk), a tyrosine kinase previously shown to phosphorylate skAE1 (M. W. Musch, E. H. Hubert, and L. Goldstein. J Biol Chem 274: 7923-7928, 1999). Hyposmotic-induced volume expansion stimulated association of p72(syk) with a light membrane fraction of skate red blood cells. Piceatannol did not inhibit this association but decreased hyposmotically stimulated increased skAE1 tyrosine phophorylation. Movement of skAE1 from an intracellular to a surface detergent-resistant membrane domain and tetramer formation were not inhibited by piceatannol treatment. Two effects of hyposmotic-induced volume expansion, decreased band 4.1 binding and increased ankyrin, were both inhibited by piceatannol. These results suggest that at least one event requiring p72(syk) activation is pivotal for hyposmotic-induced increased transport; however, steps that do not require tyrosine phosphorylation may also play a role.  相似文献   

20.
Ruggiero CE  Dooley DM 《Biochemistry》1999,38(10):2892-2898
The stoichiometry of the topa quinone biogenesis reaction in phenylethylamine oxidase from Arthrobacter globiformis (AGAO) has been determined. We have shown that the 6e- oxidation of tyrosine to topa quinone (TPQ) consumes 2 mol of O2 and produces 1 mol of H2O2/mol of TPQ formed. The rate of H2O2 production is first-order (kobs = 1.0 +/- 0.2 min-1), a rate only slightly lower than the rate of TPQ formation directly determined previously (kobs = 1.5 +/- 0.2 min-1). This gives the following net reaction stoichiometry for TPQ biogenesis: E-Tyr + 2O2 --> E-TPQ + H2O2. This stoichiometry is in agreement with recently proposed mechanisms for TPQ biogenesis, and rules out several possible alternatives.  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