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.
Acetylglutamate and ATP accelerate the oxidative inactivation of carbamoyl phosphate synthetase I by mixtures of Fe3+, ascorbate, and O2, but the mechanism of the inactivation differs with each ligand. In the presence of acetylglutamate, MgATP prevents, Mg2+, Mn2+, and catalase have no effect, and EDTA increases the inactivation, and the two phosphorylation steps of the enzyme reaction are lost simultaneously. The inactivation appears to be mediated by dehydroascorbate and is associated with the reversible oxidation of the highly reactive cysteines 1327 and 1337 and with oxidation of non-thiolic groups in the second 40-kDa domain (the enzyme consists of 4 domains of 40, 40, 60, and 20 kDa, from the amino terminus). The data are consistent with oxidation of groups at or near the site for ATPA (ATPA yields Pi; ATPB yields carbamoyl phosphate), and with the location of this site at the interphase between the second 40-kDa and the COOH-terminal domains. The oxidative inactivation promoted by ATP is inhibited by Mg2+, Mn2+, catalase, and EDTA, is not mediated by dehydroascorbate, and is not associated with oxidation of cysteines 1327 and 1337. Groups in the 60-kDa domain are oxidized. The phosphorylation step involving ATPB is lost preferentially, and the inactivation and the binding of ATPB exhibit the same dependency on the concentration of ATP. The results indicate that the oxidation is catalyzed by FeATP bound at the site for ATPB and support the binding of ATPB in the 60-kDa domain. We also demonstrate that mercaptoethanol, reducing impurities in glycerol, and dithioerythritol, in the presence of EDTA, replace ascorbate in the oxidative system. In addition, we study the influence of the oxidation on the degradation of the enzyme by rat liver lysosomes, mitochondria, and cytosol.  相似文献   

2.
The biosynthesis of C-terminal alpha-amidated peptides from their corresponding C-terminal glycine-extended precursors is catalyzed by peptidylglycine alpha-amidating enzyme (alpha-AE) in a reaction that requires copper, ascorbate, and molecular oxygen. Using bifunctional type A rat alpha-AE, we have shown that O2 is the source of the alpha-carbonyl oxygen of pyruvate produced during the amidation of dansyl-Tyr-Val-[alpha-13C]-D-Ala, as demonstrated by the 18O isotopic shift in the 13C NMR spectrum of [alpha-13C]lactate generated from [alpha-13C]pyruvate in the presence of lactate dehydrogenase and NADH. In addition, one-to-one stoichiometries have been determined for glyoxylate formed/dansyl-Tyr-Val-Gly consumed, pyruvate formed/dansyl-Tyr-Val-D-Ala consumed, dansyl-Tyr-Val-NH2 formed/ascorbate oxidized, and dansyl-Tyr-Val-NH2 formed/O2 consumed. Quantitative coupling of NADH oxidation to dansyl-Tyr-Val-NH2 production using Neurospora crassa semidehydroascorbate reductase showed that two one-electron reductions by ascorbate occurred per alpha-AE turnover. The stoichiometry of approximately 1.0 dansyl-Tyr-Val-NH2 produced/ascorbate oxidized observed in the absence of a semidehydroascorbate trap resulted from the disproportionation of two semidehydroascorbate molecules to ascorbate and dehydroascorbate.  相似文献   

3.
Captopril, an inhibitor of angiotensin-converting enzyme, has been suggested to have additional cardioprotective action because of its ability to act as an antioxidant. The rates of reaction of captopril with several biologically-relevant reactive oxygen species were determined. Captopril reacts slowly, if at all, with superoxide (rate constant less than 10(3) M-1 s-1) or hydrogen peroxide (rate constant less than M-1 s-1). It does not inhibit peroxidation of lipids stimulated by iron ions and ascorbate or by the myoglobin/H2O2 system. Indeed, mixtures of ferric ion and captopril can stimulate lipid peroxidation. Captopril reacts rapidly with hydroxyl radical (rate constant greater than 10(9) M-1 s-1) but might be unlikely to compete with most biological molecules for OH because of the low concentration of captopril that can be achieved in vivo during therapeutic use. Captopril did not significantly inhibit iron ion-dependent generation of hydroxyl radicals from hydrogen peroxide. By contrast, captopril is a powerful scavenger of hypochlorous acid: it was able to protect alpha 1-antiproteinase (alpha 1 AP) against inactivation by this species and to prevent formation of chloramines from taurine. We suggest that the antioxidant action of captopril in vivo is likely to be limited, and may be restricted to protection against damage by hypochlorous acid derived from the action of neutrophil myeloperoxidase.  相似文献   

4.
We have compared the abilities of ascorbate and reduced glutathione (GSH) to act as intracellular free radical scavengers and protect cells against radical-mediated lipid peroxidation. Phenoxyl radicals were generated in HL60 cells, through the action of their myeloperoxidase, by adding H2O2 and phenol. Normally cultured cells, which contain no ascorbate; cells that had been preloaded with ascorbate; and those that had been depleted of GSH with buthionine sulfoximine were investigated. Generation of phenoxyl radicals resulted in the oxidation of ascorbate and GSH. Ascorbate loss was much greater in the absence of GSH, and adding glucose gave GSH-dependent protection against ascorbate loss. Ascorbate, or glucose metabolism, had little effect on the GSH loss. Glutathionyl radical formation was detected by spin trapping with DMPO in cells lacking ascorbate, and the signal was suppressed by ascorbate loading. Addition of phenol plus H2O2 to the cells caused lipid peroxidation, as measured with C11-BODIPY. Peroxidation was greatest in cells that lacked both ascorbate and GSH. Either scavenger alone gave substantial inhibition but optimal protection was seen with both present. These results indicate that GSH and ascorbate can each act as an intracellular radical scavenger and protect against lipid peroxidation. With both present, ascorbate is preferred and acts as the ultimate radical sink for phenoxyl or glutathionyl radicals. However, GSH is still consumed by metabolically recycling dehydroascorbate. Thus, recycling scavenging by ascorbate does not spare GSH, but it does enable the two antioxidants to provide more protection against lipid peroxidation than either alone.  相似文献   

5.
Traces of iron, when complexed with either EDTA or diethylenetriaminepentaacetic acid (DTPA), catalyze an OH.-producing reaction between H2O2 and paraquat radical (PQ+.): H2O2 + PQ+.----PQ++ + OH. + OH-.[1]. Kinetic studies show that oxidation of formate induced by this reaction occurs by a Fenton-type mechanism, analagous to that assumed in the metal-catalyzed Haber-Weiss reaction, in which the rate determining step is H2O2 + Fe2+ (chelator)----Fe3+(chelator) + OH. + OH-,[7]; with k7 = 7 X 10(3) M-1 s-1 for EDTA and 8 X 10(2) M-1 s-1 for DTPA at pH 7.4. PQ+. rapidly reduces both Fe3+ (EDTA) and Fe3+ (DTPA), and hence allows both agents to catalyze [1] with comparable efficiency, in contrast to the much lower efficiency reported for the latter as a catalyst for the Haber-Weiss reaction. The catalytic properties of these chelating agents is attributed to their lowering of E0 (Fe3+/Fe2+) by 0.65 V, thus making [7] thermodynamically possible at pH 7. Approximately 2.5% of the OH. produced is consumed by internal or "cage" reactions, which decompose the chelator and produce CO2; however, the majority (97%) diffuses into the bulk solution and participates in competitive reactions with OH. scavengers.  相似文献   

6.
The objective of the present study is to delineate the mechanism of oxidative damage in human gastric ulcerated mucosa despite the presence of some antioxidant enzymes. We report for the first time the critical role of an endogenous peroxidase, a major H(2)O(2) metabolizing enzyme, in controlling oxidative damage in gastric mucosa. Human gastric mucosa contains a highly active peroxidase in addition to the myeloperoxidase contributed by neutrophil. In both non-Helicobacter pylori (H. pylori)- and H. pylori-mediated gastric ulcer, when myeloperoxidase level increases due to neutrophil accumulation, gastric peroxidase (GPO) level decreases significantly. Moreover, gastric ulcer is associated with oxidative damage of the mucosa as evidenced by significant increase in lipid peroxidation, protein oxidation, and thiol depletion indicating accumulation of reactive oxygen metabolites (ROM). Mucosal total superoxide dismutase (Mn and Cu-Zn SOD) level also decreases significantly leading to increased accumulation of O(2)(*-). To investigate the plausible ROM-mediated inactivation of the GPO during ulceration, the enzyme was partially purified from the mucosa. When exposed to an in vitro ROM generating system, using Cu(2+), ascorbate, and H(2)O(2,) the enzyme gets inactivated, which is dependent on Cu(2+), ascorbate, or H(2)O(2). Insensitivity to SOD excludes inactivation by O(2)(*-). However, complete protection by catalase indicates that H(2)O(2) is essential for inactivation. Sensitivity to EDTA and hydroxyl radical *OH) scavengers indicates that GPO is inactivated most probably by *OH generated from H(2)O(2). We propose that GPO is inactivated in vivo by ROM generated by activated neutrophil. This leads to further accumulation of endogenous H(2)O(2) to cause more oxidative damage to aggravate the ulcer.  相似文献   

7.
P Pasta  G Mazzola  G Carrea 《Biochemistry》1987,26(5):1247-1251
Diethyl pyrocarbonate inactivated the tetrameric 3 alpha,20 beta-hydroxysteroid dehydrogenase with second-order rate constants of 1.63 M-1 s-1 at pH 6 and 25 degrees C or 190 M-1 s-1 at pH 9.4 and 25 degrees C. The activity was slowly and partially restored by incubation with hydroxylamine (81% reactivation after 28 h with 0.1 M hydroxylamine, pH 9, 25 degrees C). NADH protected the enzyme against inactivation with a Kd (10 microM) very close to the Km (7 microM) for the coenzyme. The ultraviolet difference spectrum of inactivated vs. native enzyme indicated that a single histidyl residue per enzyme subunit was modified by diethyl pyrocarbonate, with a second-order rate constant of 1.8 M-1 s-1 at pH 6 and 25 degrees C. The histidyl residue, however, was not essential for activity because in the presence of NADH it was modified without enzyme inactivation and modification of inactivated enzyme was rapidly reversed by hydroxylamine without concomitant reactivation. Progesterone, in the presence of NAD+, protected the histidyl residue against modification, and this suggests that the residue is located in or near the steroid binding site of the enzyme. Diethyl pyrocarbonate also modified, with unusually high reaction rate, one lysyl residue per enzyme subunit, as demonstrated by dinitrophenylation experiments carried out on the treated enzyme. The correlation between inactivation and modification of lysyl residues at different pHs and the protection by NADH against both inactivation and modification of lysyl residues indicate that this residue is essential for activity and is located in or near the NADH binding site of the enzyme.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

8.
Caspases have an active site cysteine whose oxidation blocks catalytic activity. Caspase activity, measured in lysates of apoptotic cells, was inhibited by H2O2 with an IC50 of 7 microM. Recombinant caspase-3 was directly inhibited by H2O2, with an estimated second-order rate constant of 750 M-1 s-1. These values were determined when H2O2 was added while the caspases were cleaving a peptide substrate. There was a 40-fold decrease in sensitivity to inactivation if the substrate was absent at the time of H2O2 addition. These results rationalise conflicting reports of the sensitivity of caspase-3 to H2O2, and identify a novel mechanism for sensitising a thiol enzyme to oxidative inactivation.  相似文献   

9.
Treatment of the Cu(II)-Fe(III) derivative of pig allantoic fluid acid phosphatase with hydrogen peroxide caused irreversible inactivation of the enzyme and loss of half of the intensity of the visible absorption spectrum. Phosphate, a competitive inhibitor, protected against this inactivation, suggesting that it occurred as a result of a reaction at the active site. The native Fe(II)-Fe(III) enzyme was irreversibly inactivated by H2O2 to a much smaller extent than the Cu(II)-Fe(III) derivative, whereas the Zn(II)-Fe(III) derivative was stable to H2O2 treatment. The rates of inactivation of the Cu(II)-Fe(III) and Fe(II)-Fe(III) enzymes in the presence of H2O2 were increased by addition of ascorbate. These results suggest involvement of a Fenton-type reaction, generating hydroxyl radicals which react with essential active site groups. Experiments carried out on the Fe(II)-Fe(III) enzyme showed that irreversible inactivation by H2O2 in the presence of ascorbate obeyed pseudo first-order kinetics. A plot of kobs for this reaction against H2O2 concentration (at saturating ascorbate) was hyperbolic, giving kobs(max) = 0.41 +/- 0.025 min-1 and S0.5(H2O2) = 1.16 +/- 0.18 mM. A kinetic scheme is presented to describe the irreversible inactivation, involving hydroxyl radical generation by reaction of H2O2 with Fe(II)-Fe(III) enzyme, reduction of the product Fe(III)-Fe(III) enzyme by ascorbate and reaction of hydroxyl radical with an essential group in the enzyme.  相似文献   

10.
Sheep liver 6-phosphogluconate dehydrogenase is shown to be inactivated by diethylpyrocarbonate in a biphasic manner at pH 6.0, 25 degrees C. After allowing for the hydrolysis of the reagent, rate constants of 56 M-1 s-1 and 11.0 M-1 s-1 were estimated for the two processes. The complete reactivation of partially inactivated enzyme by neutral hydroxylamine, the elimination of the possibility that modification of cysteine or tyrosine residues are responsible for inactivation, and the magnitudes of the rate constants for inactivation relative to the experimentally determined value for the reaction of diethylpyrocarbonate with N alpha-acetylhistidine (2.2 M-1 s-1), all suggested that enzyme inactivation occurs solely by modification of histidine residues. Comparison of the experimental plot of residual fractional activity versus the number of modified histidine residues per subunit with simulated plots for three hypothetical models, each predicting biphasic kinetics, indicated that inactivation results from the modification of at most one essential histidine residue per subunit, although it appears that other (non-essential) histidines react independently. This histidine is thought to be His-242 and is present in the active site. Evidence in support of its role in catalysis is briefly discussed. Both 6-phosphogluconate and organic phosphate protect against inactivation, and a kinetic analysis of the protection indicated a dissociation constant of 2.1 X 10(-6) M for the enzyme--6-phosphogluconate complex. NADP+ also protected, but this might be due, at least in part, to a reduction in the effective concentration of diethylpyrocarbonate.  相似文献   

11.
The reaction of H2O2 with reduced cytochrome c oxidase was investigated with rapid-scan/stopped-flow techniques. The results show that the oxidation rate of cytochrome a3 was dependent upon the peroxide concentration (k = 2 X 10(4) M-1 X s-1). Cytochrome a and CuA were oxidised with a maximal rate of approx. 20 s-1, indicating that the rate of internal electron transfer was much slower with H2O2 as the electron acceptor than with O2 (k greater than or equal to 700 s-1). Although other explanations are possible, this result strongly suggests that in the catalytic cycle with oxygen as a substrate the internal electron-transfer rate is enhanced by the formation of a peroxo-intermediate at the cytochrome a3-CuB site. It is shown that H2O2 took up two electrons per molecule. The reaction of H2O2 with oxidised cytochrome c oxidase was also studied. It is shown that pulsed oxidase readily reacted with H2O2 (k approximately 700 M-1 X s-1). Peroxide binding is followed by an H2O2-independent conformational change (k = 0.9 s-1). Resting oxidase partially bound H2O2 with a rate similar to that of pulsed oxidase; after H2O2 binding the resting enzyme was converted into the pulsed conformation in a peroxide-independent step (k = 0.2 s-1). Within 5 min, 55% of the resting enzyme reacted in a slower process. We conclude from the results that oxygenated cytochrome c oxidase probably is an enzyme-peroxide complex.  相似文献   

12.
The reaction of choline acetyltransferase with methoxycarbonyl alkyl disulfides leads to a progressive loss in enzyme activity as the size of the alkyl group increases from methyl to n-butyl. Reaction with 5,5'-dithiobis(2-nitrobenzoic acid) (DTNB) or methoxycarbonyl coenzyme A (CoA) disulfide, leads to a total loss of enzyme activity. DTNB inactivation is biphasic (k1 = approximately 9 x 10(2) M-1 s-1, k2 = approximately 6 x 10(1) M-1 s-1) with the slow phase being diminished by acetyl-CoA. Methoxycarbonyl-CoA disulfide inactivation is also biphasic (k1 = approximately 2.1 x 10(3) M-1 s-1, k2 = approximately 6 x 10(1) M-1 s-1), with the rapid phase being diminished in the presence of acetyl-CoA. Inactivation by methoxycarbonyl methyl disulfide, ethyl disulfide, or hydroxyethyl disulfide, or by methyl methanethiosulfonate is not biphasic. Pretreatment of the enzyme with methyl methanethiosulfonate, which leads to a 25% loss in enzyme activity, abolishes the fast phase of DTNB inactivation, the slow phase of methoxycarbonyl-CoA disulfide inactivation, and any further inactivation by methoxycarbonyl ethyl disulfide. These results are interpreted to suggest that choline acetyltransferase contains two classes of reactive sulfhydryl groups, neither of which are required for enzyme activity.  相似文献   

13.
D Y Cai  M Tien 《Biochemistry》1990,29(8):2085-2091
The oxycomplexes (compound III, oxyperoxidase) of two lignin peroxidase isozymes, H1 (pI = 4.7) and H8 (pI = 3.5), were characterized in the present study. After generation of the ferroperoxidase by photochemical reduction with deazoflavin in the presence of EDTA, the oxycomplex is formed by mixing ferroperoxidase with O2. The oxycomplex of isozyme H8 is very stable, with an autoxidation rate at 25 degrees C too slow to measure at pH 3.5 or 7.0. In contrast, the oxycomplex of isozyme H1 has a half-life of 52 min at pH 4.5 and 29 min at pH 7.5 at 25 degrees C. The decay of isozyme H1 oxycomplex follows a single exponential. The half-lives of lignin peroxidase oxycomplexes are much longer than those observed with other peroxidases. The binding of O2 to ferroperoxidase to form the oxycomplex was studied by stopped-flow methods. At 20 degrees C, the second-order rate constants for O2 binding are 2.3 X 10(5) and 8.9 X 10(5) M-1 s-1 for isozyme H1 and 6.2 X 10(4) and 3.5 X 10(5) M-1 s-1 for isozyme H8 at pH 3.6 and pH 6.8, respectively. The dissociation rate constants for the oxycomplex of isozyme H1 (3.8 Z 10(-3) s-1) and isozyme H8 (1.0 X 10(-3) s-1) were measured at pH 3.6 by CO trapping. Thus, the equilibrium constants (K, calculated from kon/koff) for both isozymes H1 (7.0 X 10(7) M-1) and H8 (6.2 X 10(7) M-1) are higher than that of myoglobin (1.9 Z 10(6) M-1).(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

14.
The mechanism of lignin peroxidase (LiP) was examined using bovine pancreatic ribonuclease A (RNase) as a polymeric lignin model substrate. SDS/PAGE analysis demonstrates that an RNase dimer is the major product of the LiP-catalyzed oxidation of this protein. Fluorescence spectroscopy and amino acid analyses indicate that RNase dimer formation is due to the LiP-catalyzed oxidation of Tyr residues to Tyr radicals, followed by intermolecular radical coupling. The LiP-catalyzed polymerization of RNase in strictly dependent on the presence of veratryl alcohol (VA). In the presence of 100 microM H2O2, relatively low concentrations of RNase and VA, together but not individually, can protect LiP from H2O2 inactivation. The presence of RNase strongly inhibits VA oxidation to veratraldehyde by LiP; whereas the presence of VA does not inhibit RNase oxidation by LiP. Stopped-flow and rapid-scan spectroscopy demonstrate that the reduction of LiP compound I (LiPI) to the native enzyme by RNase occurs via two single-electron steps. At pH 3.0, the reduction of LiPI by RNase obeys second-order kinetics with a rate constant of 4.7 x 10(4) M-1.s-1, compared to the second-order VA oxidation rate constant of 3.7 x 10(5) M-1.s-1. The reduction of LiP compound II (LiPII) by RNase also follows second-order kinetics with a rate constant of 1.1 x 10(4) M-1.s-1, compared to the first-order rate constant for LiPII reduction by VA. When the reductions of LiPI and LiPIi are conducted in the presence of both VA and RNase, the rate constants are essentially identical to those obtained with VA alone. These results suggest that VA is oxidized by LiP to its cation radical which, while still in its binding site, oxidizes RNase.  相似文献   

15.
Dehydroascorbate and traces of ascorbate were present apoplastically in living spruce (Picea abies) twigs. Since the proposed apoplastic ascorbate degradation pathway contains several steps that possibly generate H(2)O(2), the effects of ascorbate and some of its degradation products were tested on apoplastic H(2)O(2) concentrations in a cell culture of P. abies as a model and on non-enzymic H(2)O(2) production in vitro. Ascorbate scavenged H(2)O(2) in the culture medium of lignin-producing Picea cells and in spent and boiled spent medium; in the presence of Cu(2+) or fresh medium, ascorbate led to the non-enzymic generation of H(2)O(2). Preparations of dehydroascorbate (the initial oxidation-product of ascorbate), and diketogulonate (the hydrolysis-product of dehydroascorbate) induced H(2)O(2) accumulation both non-enzymically and enzymically in Picea cell-suspensions. Paper electrophoresis showed that the dehydroascorbate and diketogulonate preparations contained several degradation products; some of these probably contributed to H(2)O(2) production and/or scavenging in these experiments, and would also do so in vivo. These results indicate a complex ability of apoplastic ascorbate, dehydroascorbate, diketogulonate, and further products to modulate H(2)O(2) concentrations, with potential consequences for the control of growth, development and lignification.  相似文献   

16.
Maize malic enzyme was rapidly inactivated by micromolar concentrations of cupric nitrate in the presence of ascorbate at pH, 5.0. Ascorbate or Cu2+ alone had no effect on enzyme activity. The substrate L-malate or NADP individually provided almost total protection against Cu2+-ascorbate inactivation. The loss of enzyme activity was accompanied by cleavage of the enzyme. The cleaved peptides showed molecular mass of 55 kDa, 48 kDa, 38 kDa, and 14 kDa. Addition of EDTA, histidine and imidazole provided protection. The results of protection experiments with sodium azide, DABCO and catalase suggested that reactive oxygen species were generated resulting in loss of enzyme activity. This was further supported by experiments showing that the rate of enzyme inactivation was higher in D2O than in water. It is suggested that maize malic enzyme is modified by reactive oxygen species like singlet oxygen and H2O2 generated by Cu2+-ascorbate system and the modified amino acid residue(s) may be located at or near the substrate-binding site of the enzyme.  相似文献   

17.
The reaction of hydrogen peroxide with the copper-zinc bovine-liver superoxide dismutase at low molar ratios (0.2-20.0) of H2O2/active site between pH 7.3-10.0 leads to the loss of native enzyme as a distinct form monitored by electrophoresis. The pH dependence of the loss of native enzyme between 7.3 and 9.0 indicates the involvement of a conjugate base on the enzyme of pKa of 8.7 +/- 0.1. The rate of loss of the native enzyme is first order with respect to the concentration of both enzyme and hydrogen peroxide between pH 7.3 and 9.0 with no evidence for binding of peroxide. A second-order rate constant of 3.0 +/- 1.0 M-1 s-1 is obtained from these data. At pH 10.0 the reaction is first order with respect to enzyme concentration but saturable in H2O2. All data are consistent with the interpretation that H2O2 reacts with the enzyme at the lower pH where the reaction is dependent upon the conjugate base of a functional group on the enzyme. At the higher pH, the data are consistent with the reaction of HO2- and H2O2 with the dismutase. The dissociation constant for HO2- calculated from the kinetic data at pH 10.0 is between 25-50 microM and the rate constant for the breakdown of the HO2- dismutase complex is 1.10 + 0.05 x 10(-2) s-1. The change in the electrophoretic pattern at all pH values is accompanied by the loss of the ability of the enzyme to bind copper. Weakly bound or free copper can be detected using bathocuproine disulfonate. Furthermore copper-defficient forms of the enzyme can be detected by staining gels of the peroxide-treated dismutase with diethyldithiocarbamate.  相似文献   

18.
Oxidation of Neurospora crassa glutamine synthetase.   总被引:3,自引:2,他引:1       下载免费PDF全文
The glutamine synthetase of Neurospora crassa, either purified or in cell extracts, was inactivated by ascorbate plus FeCl3 and by H2O2 plus FeSO4. The inactivation reaction was oxygen dependent, inhibited by MnCl2 and EDTA, and stimulated in cell extracts by sodium azide. This inactivation could also be brought about by adding NADPH to the cell extract. The alpha and beta polypeptides of the active glutamine synthetase were modified by these inactivating reactions, giving rise to two novel acidic polypeptides. These modifications were observed with the purified enzyme, with cell extracts, and under in vivo conditions in which glutamine synthetase is degraded. The modified glutamine synthetase was more susceptible to endogenous phenylmethylsulfonyl fluoride-insensitive proteolytic activity, which was inhibited by MnCl2 and stimulated by EDTA. The possible physiological relevance of enzyme oxidation is discussed.  相似文献   

19.
E K Hodgson  I Fridovich 《Biochemistry》1975,14(24):5294-5299
Bovine erythrocyte superoxide dismutase was slowly and irreversibly inactivated by hydrogen peroxide. The rate of this inactivation was directly dependent upon the concentrations of both H2O2 and of enzyme, and its second-order rate constant at pH 10.0 and 25 degrees was 6.7 M-1 sec-1. Inactivation was preceded by a bleaching due to rapid reduction of Cu2+ on the enzyme, and following this there was a gradual reappearance of a new absorption in the visible region, which was coincident with the loss of catalytic activity. Inactivation of the enzyme was pH-dependent and indicated an essential ionization whose pKa was approximately 10.2. Replacement of H2O by D2O raised this pKa but did not diminish the catalytic activity of superoxide dismutase, measured at pH 10.0. Several compounds, including xanthine, urate, formate, and azide, protected the enzyme against inactivation by H2O2. Alcohols and benzoate, which scavenge hydroxyl radical, did not protect. Compounds with special affinity for singlet oxygen were similarly ineffective. The data were interpreted in terms of the reduction of the enzyme-bound Cu2+ to Cu+, by H2O2, followed by a Fenton's type reaction of the Cu+ with additional H2O2. This would generate Cu2+-OH- or its ionized equivalent, Cu2+-O--, which could then oxidatively attack an adjacent histidine and thus inactivate the enzyme. Compounds which protected the enzyme could have done so by reacting with the bound oxidant, in competition with the adjacent histidine.  相似文献   

20.
The mechanism of action of the flavoprotein D-aspartate oxidase (EC 1.4.3.1) has been investigated by steady-state and stopped flow kinetic studies using D-aspartate and O2 as substrates in 50 mM KPi, 0.3 mM EDTA, pH 7.4, 4 degrees C. Steady-state results indicate that a ternary complex containing enzyme, O2, and substrate (or product) is an obligatory intermediate in catalysis. The kinetic parameters are turnover number = 11.1 s-1, Km(D-Asp) = 2.2 x 10(-3) M, Km(O2) = 1.7 x 10(-4) M. Rapid reaction studies show that 1) the reductive half reaction is essentially irreversible with a maximum rate of reduction of 180 s-1; 2) the free reduced enzyme cannot be the species which is reoxidized during turnover since its reoxidation by oxygen (second order rate constant equal to 5.3 x 10(2) M-1 s-1) is too slow to be of relevance in catalysis; 3) reduced enzyme can bind a ligand rapidly and be reoxidized as a complex at a rate faster than that observed for the free reduced enzyme; 4) the rate of reoxidation of reduced enzyme by oxygen during turnover is dependent on both O2 and D-aspartate concentrations (second order rate constant of reaction between O2 and reduced enzyme-substrate complex equal to 6.2 x 10(4) M-1 s-1); and 5) the rate-limiting step in catalysis occurs after reoxidation of the enzyme and before its reduction in the following turnover. A mechanism involving reduction of enzyme by substrate, dissociation of product from reduced enzyme, binding of a second molecule of substrate to the reduced enzyme, and reoxidation of the reduced enzyme-substrate complex is proposed for the enzyme-catalyzed oxidation of D-aspartate.  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