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.
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   

2.
Purified rat liver phenylalanine hydroxylase is inactivated in vitro by ascorbate and thiol compounds, dithiothreitol being the most effective inhibitor, with a second order rate constant for the inactivation of 0.066 +/- 0.002 mM-1.min-1 at 20 degrees C and pH 7.2. Anaerobic conditions and catalase protected the enzyme from inactivation by dithiothreitol. This suggests that hydrogen peroxide, produced by oxidation of the thiol, is involved in the inactivation. The substrate, L-phenylalanine, also partially protected the enzyme from this inactivation. It is shown that incubation of the enzyme with dithiothreitol at aerobic conditions, followed by gel filtration, causes the release of iron from the active site. The inactivation by dithiothreitol was reversed by incubation of the iron-depleted enzyme with Fe(II).  相似文献   

3.
The interaction of protein serine/threonine phosphatase calcineurin (CaN) with superoxide and hydrogen peroxide was investigated. Superoxide specifically inhibited phosphatase activity of CaN toward RII (DLDVPIPGRFDRRVSVAAE) phosphopeptide in tissue and cell homogenates as well as the activity of the enzyme purified under reducing conditions. Hydrogen peroxide was an effective inhibitor of CaN at concentrations several orders of magnitude higher than superoxide. Inhibition by superoxide was calcium/calmodulin-dependent. Nitric oxide (NO) antagonized superoxide action on CaN. We provide kinetic and spectroscopic evidence that native, catalytically active CaN has a Fe(2+)-Zn(2+) binuclear center in its active site that is oxidized to Fe(3+)-Zn(2+) by superoxide and hydrogen peroxide. This oxidation is accompanied by a gain of manganese dependence of enzyme activity. CaN isolated by a conventional purification procedure was found in the oxidized, ferric enzyme form, and it became increasingly dependent on divalent cations. These results point to a complex redox regulation of CaN phosphatase activity by superoxide, which is modified by calcium, NO, and superoxide dismutase.  相似文献   

4.
A divalent metal ion, such as Mn2+, is required for the catalytic reaction and allosteric regulation of pig heart NAD-dependent isocitrate dehydrogenase. The enzyme is irreversibly inactivated and cleaved by Fe2+ in the presence of O2 and ascorbate at pH 7.0. Mn2+ prevents both inactivation and cleavage. Nucleotide ligands, such as NAD, NADPH, and ADP, neither prevent nor promote inactivation or cleavage of the enzyme by Fe2+. The NAD-specific isocitrate dehydrogenase is composed of three distinct subunits in the ratio 2alpha:1beta:1gamma. The results indicate that the oxidative inactivation and cleavage are specific and involve the 40 kDa alpha subunit of the enzyme. A pair of major peptides is generated during Fe2+ inactivation: 29.5 + 10.5 kDa, as determined by SDS-PAGE. Amino-terminal sequencing reveals that these peptides arise by cleavage of the Val262-His263 bond of the alpha subunit. No fragments are produced when enzyme is incubated with Fe2+ and ascorbate under denaturing conditions in the presence of 6 M urea, indicating that the native structure is required for the specific cleavage. These results suggest that His263 of the alpha subunit may be a ligand of the divalent metal ion needed for the reaction catalyzed by isocitrate dehydrogenase. Isocitrate enhances the inactivation of enzyme caused by Fe2+ in the presence of oxygen, but prevents the cleavage, suggesting that inactivation occurs by a different mechanism when metal ion is bound to the enzyme in the presence of isocitrate: oxidation of cysteine may be responsible for the rapid inactivation in this case. Affinity cleavage caused by Fe2+ implicates alpha as the catalytic subunit of the multisubunit porcine NAD-dependent isocitrate dehydrogenase.  相似文献   

5.
Tewari RK  Kumar P  Sharma PN 《Planta》2006,223(6):1145-1153
The aim of the study was to implicate the generation of reactive oxygen species (ROS) and altered cellular redox environment with the effects of Cu-deficiency or Cu-excess in mulberry (Morus alba L.) cv. Kanva 2 plants. A study of antioxidative responses, indicators of oxidative damage and cellular redox environment in Cu-deficient or Cu-excess mulberry plants was undertaken. While the young leaves of plants supplied with nil Cu showed chlorosis and necrotic scorching of laminae, the older and middle leaves of plants supplied with nil or 0.1 μM Cu showed purplish-brown pigmented interveinal areas that later turned necrotic along the apices and margins of leaves. The Cu-excess plants showed accelerated senescence of the older leaves. The Cu-deficient plants showed accumulation of hydrogen peroxide and superoxide anion radical. The accumulation of hydrogen peroxide was strikingly intense in the middle portion of trichomes on Cu-deficient leaves. Though the concentration of total ascorbate increased with the increasing supply of Cu, the ratio of the redox couple (DHA/ascorbic acid) increased in Cu-deficient or Cu-excess plants. The activities of superoxide dismutase (EC 1.15.1.1), catalase (EC 1.11.1.6), peroxidase (EC 1.11.1.7), ascorbate peroxidase (EC 1.11.1.11) and glutathione reductase (EC 1.6.4.2) increased in both Cu-deficient and Cu-excess plants. The results suggest that deficiency of Cu aggravates oxidative stress through enhanced generation of ROS and disturbed redox couple. Excess of Cu damaged roots, accelerated the rate of senescence in the older leaves, induced antioxidant responses and disturbed the cellular redox environment in the young leaves of mulberry plants.  相似文献   

6.
A NADP+-specific isocitrate dehydrogenase (EC 1.1.1.42) was isolated and purified over 400-fold from Anacystis nidulans. The enzyme activity responded slowly to rapid changes in ligand (NADP+, isocitrate, Mg2+-ions) or enzyme concentration as well as to rapid changes in temperature. These are properties characteristic of the hysteretic enzymes. In addition, the enzyme activity was subject to product (-ketoglutarate) inhibition. ATP, ADP and CDP also inhibited the enzyme. Unlike several other cyanobacterial enzymes, the isocitrate dehydrogenase of Anacystis is not under redox control.  相似文献   

7.
The inactivation of glutamine synthetase (GS; EC 6.3.1.2) by metal-catalyzed oxidation (MCO) systems was studied in several Prochlorococcus strains, including the axenic PCC 9511. GS was inactivated in the presence of various oxidative systems, either enzymatic (as NAD(P)H+NAD(P)H-oxidase+Fe(3+)+O(2)) or non-enzymatic (as ascorbate+Fe(3+)+O(2)). This process required the presence of oxygen and a metal cation, and is prevented under anaerobic conditions. Catalase and peroxidase, but not superoxide dismutase, effectively protected the enzyme against inactivation, suggesting that hydrogen peroxide mediates this mechanism, although it is not directly responsible for the reaction. Addition of azide (an inhibitor of both catalase and peroxidase) to the MCO systems enhanced the inactivation. Different thiols induced the inactivation of the enzyme, even in the absence of added metals. However, this inactivation could not be reverted by addition of strong oxidants, as hydrogen peroxide or oxidized glutathione. After studying the effect of addition of the physiological substrates and products of GS on the inactivation mechanism, we could detect a protective effect in the case of inorganic phosphate and glutamine. Immunochemical determinations showed that the concentration of GS protein significantly decreased by effect of the MCO systems, indicating that inactivation precedes the degradation of the enzyme.  相似文献   

8.
Summary The redox interconversion of Escherichia coli glutathione reductase has been studied both in situ, with permeabilized cells treated with different reductants, and in vivo, with intact cells incubated with compounds known to alter their intracellular redox state.The enzyme from toulene-permeabilized cells was inactivated in situ by NADPH, NADH, dithionite, dithiothreitol, or GSH. The enzyme remained, however, fully active upon incubation with the oxidized forms of such compounds. The inactivation was time-, temperature-, and concentration-dependent; a 50% inactivation was promoted by just 2 M NADPH, while 700 M NADH was required for a similar effect. The enzyme from permeabilized cells was completely protected against redox inactivation by GSSG, and to a lesser extent by dithiothreitol, GSH, and NAD(P)+. The inactive enzyme was efficiently reactivated in situ by physiological GSSG concentrations. A significant reactivation was promoted also by GSH, although at concentrations two orders of magnitude below its physiological concentrations. The glutathione reductase from intact E. coli cells was inactivated in vivo by incubation with DL-malate, DL-isocitrate, or higher L-lactate concentrations. The enzyme was protected against redox inactivation and fully reactivated by diamide in a concentration-dependent fashion. Diamide reactivation was not dependent on the synthesis of new protein, thus suggesting that the effect was really a true reactivation and not due to de novo synthesis of active enzyme. The glutathione reductase activity increased significantly after incubation of intact cells with tert-butyl or cumene hydroperoxides, suggesting that the enzyme was partially inactive within such cells. In conclusion, the above results show that both in situ and in vivo the glutathione reductase of Escherichia coli is subjected to a redox interconversion mechanism probably controlled by the intracellular NADPH and GSSG concentrations.  相似文献   

9.
Reduction potentials for the catalytic compound I/compound II and compound II/Fe3+ redox couples, and for the two-electron compound I/Fe3+ redox couple, have been determined for ascorbate peroxidase (APX) and for a number of site-directed variants. For the wild type enzyme, the values are E degrees '(compound I/compound II) = 1156 mV, E degrees '(compound II/Fe3+) = 752 mV, and E degrees '(compound I/Fe3+) = 954 mV. For the variants, the analysis also includes determination of Fe3+/Fe2+ potentials which were used to calculate (experimentally inaccessible) E degrees '(compound II/Fe3+) potentials. The data provide a number of new insights into APX catalysis. The measured values for E degrees '(compound I/compound II) and E degrees '(compound II/Fe3+) for the wild type protein account for the much higher oxidative reactivity of compound I compared to compound II, and this correlation holds for a number of other active site and substrate binding variants of APX. The high reduction potential for compound I also accounts for the known thermodynamic instability of this intermediate, and it is proposed that this instability can account for the deviations from standard Michaelis kinetics observed for most APX enzymes during steady-state oxidation of ascorbate. This study provides the first systematic evaluation of the redox properties of any ascorbate peroxidase using a number of methods, and the data provide an experimental and theoretical framework for accurate determination of the redox properties of Fe3+, compound I, and compound II species in related enzymes.  相似文献   

10.
Inactivation of Ascorbate Peroxidase by Thiols Requires Hydrogen Peroxide   总被引:2,自引:0,他引:2  
The hydrogen peroxide-dependent oxidation of ascorbate by ascorbateperoxidase from tea leaves was inhibited by thiols, such asdithiothreitol, glutathione, mercaptoethanol and cysteine. Thesethiols themselves did not inactivate the enzyme. However, theyinactivated the enzyme when hydrogen peroxide was produced bythe metal-catalyzed oxidation of thiols or when exogenous hydrogenperoxide was added. Thiols were oxidized by ascorbate peroxidaseand hydrogen peroxide to thiyl radicals, as detected by theESR spectra of the thiyl radical-5,5'-dimethyll- pyrroline-N-oxidieadducts. Inactivation of ascorbate peroxidase by thiols andhydrogen peroxide is caused by the interaction of the enzymewith the thiyl radicals produced at its reaction center. (Received September 10, 1991; Accepted December 9, 1991)  相似文献   

11.
The purification and in vitro inactivation of AMP-deaminase from white muscle of carp Cyprinus carpio were conducted in the Fe2+/H2O2 and Fe2+/ascorbate oxidation systems. The enzyme activity decreases by 50% within 30 minutes of incubation in the presence of 100 microM of hydrogen peroxide and 5 microM of ferrous sulfate. Inactivation depended on incubation time and concentrations of FeSO4 and H2O2. In the system Fe2+/ascorbate the enzyme activity decreased by 50% at concentration of ascorbate 1 mM and 5 ferrous sulfate microM. Sodium nitrite did not affect the activity. S(0.5) and n(H) of both native and partially inactivated enzymes were virtually the same, while maximal activity of the inactivated enzyme was 2-3-fold lower than that of the native one.  相似文献   

12.
Pig heart NAD-specific isocitrate dehydrogenase is inactivated by reaction with iodoacetate at pH 6.0. Loss of activity can be attributed to the formation of 1-2 mol of carboxymethyl-cysteine per peptide chain. The rate of inactivation is markedly decreased by the combined addition of Mn2+ and isocitrate, but not by alpha-ketoglutarate, the coenzyme NAD or the allosteric activator ADP. The substrate concentration dependence of the decreased rate of inactivation yields a dissociation constant of 1.6 mM for the enzyme-manganous-dibasic isocitrate complex, a value that is 50 times higher than the Km for this substrate. This result suggests that in protecting the enzyme against iodoacetate, isocitrate may bind to a region distinct from the catalytic site. Isocitrate and Mn2+ also prevent thermal denaturation, with an affinity for the enzyme close to that observed for the iodoacetate-sensitive site. The alkylatable cysteine residues may contribute to a manganous-isocitrate binding site which is responsible for stabilizing an active conformation of the enzyme.  相似文献   

13.
Three Solanum genotypes with various polygenic resistance levels to the oomycete pathogen Phytophthora infestans (Mont.) De Bary were studied for their antioxidant response to the pathogen culture filtrate (CF). Detached plant leaves were treated with CF for 6, 18 and 30 h, and assayed for changes in hydrogen peroxide content, total ascorbate and glutathione pools and redox ratios (reduced form to total pool), as well as for changes in activities of ascorbate peroxidase, glutathione reductase and glutathione-S-transferase. In CF treated leaves of non-host resistant S. nigrum var. gigantea and field resistant S. tuberosum cv Bzura, the H(2)O(2) content did not change in comparison to water treated control leaves, whereas in the susceptible S. tuberosum clone H-8105 it decreased below the control level. In CF treated leaves of all genotypes, the total ascorbate pools were relatively unaltered and their redox ratio changed only transiently. In Bzura leaves the total glutathione content increased earlier than in the two other genotypes. The glutathione redox ratio remained rather stable, except for the susceptible clone H-8105, where it decreased transiently by about 42%. The relative increases in activity of all the studied enzymes were the highest in the susceptible clone H-8105. The results are discussed in the light of oxidative processes occurring in CF treated leaves. We conclude that stringent control of pro- and anti-oxidant reactions bringing the H(2)O(2) and/or cellular redox state to the threshold level is decisive for deployment of an effective defense strategy.  相似文献   

14.
Xanthine oxidase inactivation by reagents that modify thiol groups   总被引:1,自引:1,他引:0  
1. The presence of xanthine was required for the inhibition of bovine milk xanthine oxidase by o-iodosobenzoate, iodoacetamide, hydrogen peroxide or p-chloromercuribenzoate. 2. Inactivation by p-chloromercuribenzoate was very rapid, was reversed by cysteine and was less in the presence of FAD. Lineweaver-Burk plots showed that the inactivation by p-chloromercuribenzoate was competitive with substrate. 3. Inactivation by o-iodosobenzoate, iodoacetamide or hydrogen peroxide could not be reversed by cysteine or xanthine. However, the presence of xanthine during the incubation with inhibitor protected the enzyme against o-iodosobenzoate but not against iodoacetamide or hydrogen peroxide. 4. p-Chloromercuribenzoate protected the enzyme against inactivation by hydrogen peroxide.  相似文献   

15.
Dark addition of hydrogen peroxide to intact spinach chloroplastsresulted in the inactivation of ascorbate peroxidase accompaniedby a decrease in ascorbate contents. This was also the casein reconstituted chloroplasts containing ascorbate, NADP+, NAD+and ferredoxin. The addition of hydrogen peroxide during light,however, showed little effect on ascorbate contents and ascorbateperoxidase activity in either the intact or reconstituted chloroplasts.In contrast to ascorbate peroxidase, the enzymes participatingin the regeneration of ascorbate in chloroplasts (monodehydroascorbatereductase, dehydroascorbate reductase and glutathione reductase)were not affected by the dark addition of hydrogen peroxide.Ascorbate contents increased again by illumination of the chloroplastsafter the dark addition of hydrogen peroxide. These resultsshow that the inactivation of the hydrogen peroxide scavengingsystem on dark addition of hydrogen peroxide [Anderson et al.(1983) Biochim. Biophys. Acta 724: 69, Asada and Badger (1984)Plant & Cell Physiol. 25: 1169] is caused by the loss ofascorbate peroxidase activity. Ascorbate peroxidase activitywas rapidly lost in ascorbate-depleted medium, and protectedby its electron donors, ascorbate, isoascorbate, guaiacol andpyrogallol, but not by GSH, NAD(P)H and ferredoxin. (Received June 14, 1984; Accepted August 15, 1984)  相似文献   

16.
Redox interconversion of glutathione reductase was studiedin situ withS. cerevisiae. The enzyme was more sensitive to redox inactivation in 24 hour-starved cells than in freshly-grown ones. While 5 μM NADPH or 100 μM NADH caused 50% inactivation in normal cells in 30 min, 0.75 μM NADPH or 50 μM NADH promoted a similar effect in starved cells. GSSG reactivated the enzyme previously inactivated by NADPH, ascertaining that the enzyme was subjected to redox interconversion. Low EDTA concentrations fully protected the enzyme from NADPH inactivation, thus confirming the participation of metals in such a process. Extensive inactivation was obtained in permeabilized cells incubated with glucose-6-phosphate or 6-phosphogluconate, in agreement with the very high specific activities of the corresponding dehydrogenases. Some inactivation was also observed with malate, L-lactate, gluconate or isocitrate in the presence of low NADP+ concentrations. The inactivation of yeast glutathione reductase has also been studiedin vivo. The activity decreased to 75% after 2 hours of growth with glucono-δ-lactone as carbon source, while NADPH rose to 144% and NADP+ fell to 86% of their initial values. Greater changes were observed in the presence of 1.5 μM rotenone: enzymatic activity descended to 23% of the control value, while the NADH/NAD+ and NADPH/NADP+ ratios rose to 171% and 262% of their initial values, respectively. Such results indicate that the lowered redox potential of the pyridine nucleotide pool existing when glucono-δ-lactone is oxidized promotesin vivo inactivation of glutathione reductase.  相似文献   

17.
The inhibitory effect on PTP1B caused by the addition of pyridazine analogues has been investigated. Biophysical techniques, that is, mass spectrometry (MS), nuclear magnetic resonance (NMR), and isothermal titration calorimetry (ITC) were used for the characterization. Pyridazine analogues cause catalytic oxidation of the reducing agent, generating hydrogen peroxide that oxidizes the active site cysteine on the enzyme, leading to enzyme inactivation. Two additional compound classes show the same effect. We found one common structural feature in these molecules that allows the reaction with triplet molecular oxygen to be less endothermic. A proposed mechanism for the catalytic redox cycle is described.  相似文献   

18.
Hydrogen peroxide inactivates the purified human angiotensin-converting enzyme (ACE) in vitro; the inactivating effect of H2O2 is eliminated by an addition of catalase. The lung and kidney ACE are equally sensitive to the effect of hydrogen peroxide. After addition of oxidants (H2O2 alone or H2O2 + ascorbate or H2O2 + Fe2+ mixtures) to the membranes or homogenates of the lung, the inactivation of membrane-bound ACE is far less pronounced despite the large-scale accumulation of lipid peroxidation products. The marked inactivation of ACE in the membrane fraction (up to 55% of original activity) was observed during ACE incubation with a glucose:glucose oxidase:Fe2+ mixture. Presumably the oxidative potential of H2O2 in tissues in consumed, predominantly, for the oxidation of other components of the membrane (e.g., lipids) rather than for ACE inactivation.  相似文献   

19.
Methionine sulfoxide reductase A is an essential enzyme in the antioxidant system which scavenges reactive oxygen species through cyclic oxidation and reduction of methionine and methionine sulfoxide. Recently it has also been shown to catalyze the reverse reaction, oxidizing methionine residues to methionine sulfoxide. A cysteine at the active site of the enzyme is essential for both reductase and oxidase activities. This cysteine has been reported to have a pK(a) of 9.5 in the absence of substrate, decreasing to 5.7 upon binding of substrate. Using three independent methods, we show that the pK(a) of the active site cysteine of mouse methionine sulfoxide reductase is 7.2 even in the absence of substrate. The primary mechanism by which the pK(a) is lowered is hydrogen bonding of the active site Cys-72 to protonated Glu-115. The low pK(a) renders the active site cysteine susceptible to oxidation to sulfenic acid by micromolar concentrations of hydrogen peroxide. This characteristic supports a role for methionine sulfoxide reductase in redox signaling.  相似文献   

20.
Iron and aconitase activity   总被引:5,自引:2,他引:3       下载免费PDF全文
Aconitase activated with Fe(2+), cysteine and ascorbate incorporates 1 g-atom of Fe(2+)/mol. Loss of this Fe(2+) by transfer to ferrozine, a Fe(2+) chelator, results in loss of activity. Ascorbate increases the rate of transfer of the essential Fe(2+) whereas citrate retards the rate of transfer. Transfer of Fe(2+) from inactive aconitase, 2 g-atoms of Fe/mol, can be accomplished in the presence of urea and ascorbate. The correlation of activity with the presence of an added g-atom of Fe(2+)/mol leads to the conclusion that active aconitase has only one active site per mol.  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