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.
Anaerobically induced NAD-linked glycerol dehydrogenase of Klebsiella pneumoniae for fermentative glycerol utilization was reported previously to be inactivated in the cell during oxidative metabolism. In vitro inactivation was observed in this study by incubating the purified enzyme in the presence of O2, Fe2+, and ascorbate or dihydroxyfumarate. It appears that O2 and the reducing agent formed H2O2 and that H2O2 reacted with Fe2+ to generate an activated species of oxygen which attacked the enzyme. The in vitro-oxidized enzyme, like the in vivo-inactivated enzyme, showed an increased Km for NAD (but not glycerol) and could no longer be activated by Mn2+ which increased the Vmax of the native enzyme but decreased its apparent affinity for NAD. Ethanol dehydrogenase and 1,3-propanediol oxidoreductase, two enzymes with anaerobic function, also lost activity when the cells were incubated aerobically with glucose. However, glucose 6-phosphate dehydrogenase (NADP-linked), isocitrate dehydrogenase, and malate dehydrogenase, expected to function both aerobically and anaerobically, were not inactivated. Thus, oxidative modification of proteins in vivo might provide a mechanism for regulating the activities of some anaerobic enzymes.  相似文献   

2.
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   

3.
N J Silman  N G Carr    N H Mann 《Journal of bacteriology》1995,177(12):3527-3533
Glutamine synthetase (GS) inactivation was observed in crude cell extracts and in the high-speed supernatant fraction from the cyanobacterium Synechocystis sp. strain PCC 6803 following the addition of ammonium ions, glutamine, or glutamate. Dialysis of the high-speed supernatant resulted in loss of inactivation activity, but this could be restored by the addition of NADH, NADPH, or NADP+ and, to a lesser extent, NAD+, suggesting that inactivation of GS involved ADP-ribosylation. This form of modification was confirmed both by labelling experiments using [32P]NAD+ and by chemical analysis of the hydrolyzed enzyme. Three different forms of GS, exhibiting no activity, biosynthetic activity only, or transferase activity only, could be resolved by chromatography, and the differences in activity were correlated with the extent of the modification. Both biosynthetic and transferase activities were restored to the completely inactive form of GS by treatment with phosphodiesterase.  相似文献   

4.
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+Fe3++O2) or non-enzymatic (as ascorbate+Fe3++O2). 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   

5.
Calcineurin (CaN) is a Ca2+-and calmodulin (CaM)-dependent serine/threonine phosphatase containing a dinuclear Fe-Zn center in the active site. Recent studies have indicated that CaN is a possible candidate for redox regulation. The inactivation of bovine brain CaN and of the catalytic CaN A-subunit from Dictyostelium by the vicinal dithiol reagents phenylarsine oxide (PAO) and melarsen oxide (MEL) and by H2O2 was investigated. PAO and MEL inhibited CaN with an IC50 of 3-8 microM and the inactivation was reversed by 2, 3-dimercapto-1-propane sulfonic acid. The treatment of isolated CaN with hydrogen peroxide resulted in a concentration-dependent inactivation. Analysis of the free thiol content performed on the H2O2 inactivated enzyme demonstrated that only two or three of the 14 Cys residues in CaN are modified. The inactivation of CaN by H2O2 could be reversed with 1,4-dithiothreitol and with the dithiol oxidoreductase thioredoxin. We propose that a bridging of two closely spaced Cys residues in the catalytic CaN A-subunit by PAO/MEL or the oxidative formation of a disulfide bridge by H2O2 involving the same Cys residues causes the inactivation. Our data implicate a possible involvement of thioredoxin in the redox control of CaN activity under physiological conditions. The low temperature EPR spectrum of the native enzyme was consistent with a Fe3+-Zn2+ dinuclear centre. Upon H2O2-mediated inactivation of the enzyme no significant changes in the EPR spectrum were observed ruling out that Fe2+ is present in the active enzyme and that the dinuclear metal centre is the target for the oxidative inactivation of CaN.  相似文献   

6.
As a variety of eukaryotic cells age, the specific activity of glucose-6-phosphate dehydrogenase (Glu-6-PDH) declines as much as 50%. Because of the central role of this enzyme in metabolism, it is important to define factors responsible for this loss in enzyme activity. We report that Glu-6-PDH from Leuconostoc mesenteroides is rapidly inactivated by micromolar concentrations of Fe2+ and H2O2. Inactivation correlated with the formation of one carbonyl functionality/enzyme subunit, indicating that inactivation is the result of site-specific oxidative modification. Our results suggest that Fe2+ binds to the glucose 6-phosphate binding site and that interaction of the enzyme-bound Fe2+ with H2O2 leads to the oxidative modification of amino acids essential for enzyme activity. Partially inactivated enzyme remained predominantly in the dimeric form, and no change in the apparent affinity of the remaining active subunits for substrate was observed. Partial inactivation did, however, lead to a decrease in the thermal stability of the remaining activity. This decrease in thermal stability could be largely overcome by the addition of glucose 6-phosphate. Thus, although exposure to H2O2 and Fe2+ results in the irreversible inactivation of Glu-6-PDH, the resulting modification is selective, leads to the formation of heterodimers of both active and inactive subunits, and does not appear to cause large scale structural changes. Our results demonstrate the inherent susceptibility of Glu-6-PDH from L. mesenteroides to modification by an oxidation system known to exist in vivo. An assessment of the physiological significance of Fe(2+)-catalyzed oxidation of Glu-6-PDH awaits extension of these studies to mammalian sources known to accumulate less active or inactive forms of the enzyme as a function of age.  相似文献   

7.
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   

8.
The activity of acetyl-CoA hydrolase (dimeric form) purified from the supernatant fraction of rat liver was shown to have a half-life (t1/2) of 3 min at 0 degree C, but to stable at 37 degrees C (t1/2 = 34 h) [Isohashi, F., Nakanishi, Y. & Sakamoto, Y. (1983) Biochemistry 22, 584-590]. Incubation of the purified enzyme with L-ascorbic acid (AsA) at 37 degrees C resulted in inactivation of the enzyme (t1/2 = 90 min at 2 mM AsA). The extent of inactivation was greatly enhanced by addition of transition metal ions (Cu2+, Fe2+, and Fe3+). Thiol reducing agents, such as reduced glutathione and DL-dithiothreitol, protected the hydrolase from inactivation by AsA. However, these materials did not restore the catalytic activity of the enzyme inactivated by AsA. When AsA solution containing Cu2+ was preincubated under aerobic conditions at 37 degrees C for various times in the absence of enzyme, and then aliquots were incubated with the enzyme solution for 20 min, remaining activity was found to decrease with increase in the preincubation time, reaching a minimum at 60 min. However, further preincubation reduced the potential for inactivation. Catalase, a hydrogen peroxide (H2O2) scavenger, almost completely prevented inactivation of the enzyme by AsA plus Cu2+. Superoxide dismutase and tiron, which are both superoxide (O2-) scavengers, also prevented inactivation of the enzyme. A high concentration of mannitol, a hydroxyl radical (OH) scavenger, partially protected the enzyme from inactivation. These results suggest that inactivation of the enzyme by AsA in the presence of Cu2+ was due to the effect of active oxygen species (H2O2, O2-, OH) that are known to be autoxidation products of AsA. Valeryl-CoA, a competitive inhibitor of acetyl-CoA hydrolase, greatly protected the enzyme from inactivation by AsA plus Cu2+, but ATP and ADP, which are both effectors of this enzyme, had only slight protective effects. These results suggest that inactivation of this enzyme by addition of AsA plus Cu2+ was mainly due to attack on its active site.  相似文献   

9.
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   

10.
对三种固氮蓝藻:固氮鱼腥藻(水生686)、柱孢鱼腥藻和鱼腥藻7120的整细胞及无细胞抽提液的固氮酶活性,进行了比较研究。水生686的整细胞酶活虽然不低(51.9毫米乙烯峰高/光密度/30分),仅次于柱孢鱼腥藻,但其无细胞抽提液的酶活却最低。这可能与它含有大量藻胶有关。研究了Mn++、Fe++对蓝藻固氮酶的作用,以及测定其在不同酶浓度下的反应动力学表明:柱孢鱼腥藻中不存在象深红螺菌中所看到的那种激活因子。用甲苯-乙醇溶液处理藻细胞,对固氮酶作原位测定,探索了它的氧损伤及氧保护机理。    相似文献   

11.
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   

12.
3-Hydroxyanthranilate-3,4-dioxygenase (HAD) is a non-heme Fe(II) dependent enzyme that catalyzes the oxidative ring-opening of 3-hydroxyanthranilate to 2-amino-3-carboxymuconic semialdehyde. The enzymatic product subsequently cyclizes to quinolinate, an intermediate in the biosynthesis of nicotinamide adenine dinucleotide. Quinolinate has also been implicated in important neurological disorders. Here, we describe the mechanism by which 4-chloro-3-hydroxyanthranilate inhibits the HAD catalyzed reaction. Using overexpressed and purified bacterial HAD, we demonstrate that 4-chloro-3-hydroxyanthranilate functions as a mechanism-based inactivating agent. The inactivation results in the consumption of 2 +/- 0.8 equiv of oxygen and the production of superoxide. EPR analysis of the inactivation reaction demonstrated that the inhibitor stimulated the oxidation of the active site Fe(II) to the catalytically inactive Fe(III) oxidation state. The inactivated enzyme can be reactivated by treatment with DTT and Fe(II). High resolution ESI-FTMS analysis of the inactivated enzyme demonstrated that the inhibitor did not form an adduct with the enzyme and that four conserved cysteines were oxidized to two disulfides (Cys125-Cys128 and Cys162-Cys165) during the inactivation reaction. These results are consistent with a mechanism in which the enzyme, complexed to the inhibitor and O2, generates superoxide which subsequently dissociates, leaving the inhibitor and the oxidized iron center at the active site.  相似文献   

13.
One strain of bacteria which showed high H2O2-generating activity was isolated from soil and characterized as Arthrobacter globiformis based on its morphological, nutritional, and physiological characteristics. The activities of H2O2 generation, NAD reduction and oxygen consumption in the bacterial cells were examined using choline, betaine aldehyde or betaine as substrate. Choline was oxidized to betaine aldehyde under aerobic conditions in a reaction coupled with H2O2 generation and oxygen consumption. On the other hand, betaine aldehyde seemed to be oxidized to betaine through two distinct oxidative reactions, H2O2 generation (oxygen consumption) under aerobic conditions and NAD reduction under either aerobic or anaerobic conditions. These enzyme activities were found in the supernatant fraction of the sonicated cell preparation.  相似文献   

14.
Exposure of nitrogen-fixing cultures of Anabaena spp. to 100% oxygen resulted in the rapid decline of nitrogenase activity. When oxygen-treated cells were transferred to 100% argon, nitrogenase activity was quickly restored in a process that required protein synthesis. Anaerobiosis was not essential for the recovery process; in fact, cells of Anabaena sp. strains CA and 1F will recover nitrogenase activity after prolonged incubation in 100% oxygen. Oxygen treatment acted directly on the intracellular nitrogenase and did not affect other metabolic processes. Examination of crude extracts of oxygen-treated Anabaena sp. strain CA indicated that both components of nitrogenase are inactivated. However, several lines of evidence suggest that oxygen treatment does not result in irreversible denaturation of nitrogenase, but rather results in a reversible inactivation which may serve as a protection mechanism. Nitrogenase present in crude extracts from cells of Anabaena sp. strain 1F which had been incubated for a prolonged period in 100% oxygen was less sensitive to oxygen in vitro than was nitrogenase of a crude extract of untreated cells.  相似文献   

15.
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   

16.
Koningic acid, a sesquiterpene antibiotic, is a specific inhibitor of the enzyme glyceraldehyde-3-phosphate dehydrogenase (D-glyceraldehyde-3-phosphate:NAD+ oxidoreductase (phosphorylating), EC 1.2.1.12). In the presence of 3 mM of NAD+, koningic acid irreversibly inactivated the enzyme in a time-dependent manner. The pseudo-first-order rate constant for inactivation (kapp) was dependent on koningic acid concentration in saturate manner, indicating koningic acid and enzyme formed a reversible complex prior to the formation of an inactive, irreversible complex; the inactivation rate (k 3) was 5.5.10(-2) s-1, with a dissociation constant for inactivation (Kinact) of 1.6 microM. The inhibition was competitive against glyceraldehyde 3-phosphate with a Ki of 1.1 microM, where the Km for glyceraldehyde 3-phosphate was 90 microM. Koningic acid inhibition was uncompetitive with respect to NAD+. The presence of NAD+ accelerated the inactivation. In its absence, the charcoal-treated NAD+-free enzyme showed a 220-fold decrease in apparent rate constant for inactivation, indicating that koningic acid sequentially binds to the enzyme next to NAD+. The enzyme, a tetramer, was inactivated when maximum two sulfhydryl groups, possibly cysteine residues at the active sites of the enzyme, were modified by the binding of koningic acid. These observations demonstrate that koningic acid is an active-site-directed inhibitor which reacts predominantly with the NAD+-enzyme complex.  相似文献   

17.
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   

18.
The enzyme rhodanese (thiosulfate sulfurtransferase; EC 2.8.1.1) is inactivated with a half-time of approximately 3 min when incubated with 50 mM NADH. NAD+, however, has virtually no effect on the activity. Inactivation can be prevented by the inclusion of the substrate thiosulfate. The concentration of thiosulfate giving half-protection is 0.038 mM. In addition, NADH, but not NAD+, is a competitive inhibitor with respect to thiosulfate in the catalyzed reaction (Ki = 8.3 mM). Fluorescence studies are consistent with a time-dependent oxidation of NADH in the presence of rhodanese. The sulfur-free form of rhodanese is more rapidly inactivated than the sulfur-containing form. Spectrophotometric titrations show that inactivation is accompanied by the loss of two free SH groups per enzyme molecule. Inactivation is prevented by the exclusion of air and the inclusion of EDTA (1 mM), and the enzyme activity can be largely protected by incubation with superoxide dismutase or catalase. Rhodanese, inactivated with NADH, can be reactivated by incubation with the substrate thiosulfate (75 mM) for 48 h or more rapidly, but only partially, by incubating with 180 mM dithiothreitol. It is concluded that, in the presence of rhodanese, NADH can be oxidized by molecular oxygen and produce intermediates of oxygen reduction, such as superoxide and/or hydrogen peroxide, that can inactivate the enzyme with consequent formation of an intraprotein disulfide. In addition, NADH, but not NAD+, can reversibly bind to the active site region in competition with thiosulfate. These data are of interest in view of x-ray studies that show structural similarities between rhodanese and nucleotide binding proteins.  相似文献   

19.
To elucidate mechanism of ganglioside neuroprotection, it is important to study their metabolic effects, specifically of action on Na+, K+ -ATPase. It has been shown that under effect of oxidative stress inductors and neurotoxins an oxidative inactivation of this enzyme takes place in PC12 cells and brain cortex synaptosomes, this inactivation being able to be prevented or decreased by ganglioside GM1. Thus, for instance, 24 h after action of 1 mM H2O2, activity of Na+, K+ -ATPase in PC12 cells decreased more than twice. However, in the case of preincubation of the cells with ganglioside GM1 prior to the H2O2 action this enzyme activity did not differ statistically significantly from control. Ganglioside GM1 also was able to increase significantly the enzyme activity decreased by action on the PC12 cells of amyloid beta-peptide (AP) causing lesion of neurons in Alzheimer's disease and at low H202 concentrations. Experiments on brain cortex synaptosomes have established that not only antioxidants--alpha-tocopherol and superoxide dismutase--but also ganglioside GM1 prevent the glutamateproduced Na+, K+ -ATPase oxidative inactivation. The obtained data agree with a suggestion that the ganglioside neuroprotective effect at action on nerve cells of such toxins as Abeta, glutamate or reactive oxygen species is due to their ability to inhibit the free-radical reactions.  相似文献   

20.
Evidence is presented which suggests that the NAD(P)H-cytochrome c reductase component of nitrate reductase is the main site of action of the inactivating enzyme. When tested on the nitrate reductase (NADH) from the maize root and scutella, the NADH-cytochrome c reductase was inactivated at a greater rate than was the FADH2-nitrate reductase component. With the Neurospora nitrate reductase (NADPH) only the NADPH-cytochrome c reductase was inactivated. p-Chloromercuribenzoate at 50 muM, which gave almost complete inhibition of the NADH-cytochrome c reductase fraction of the maize nitrate reductase, had no marked effect on the action of the inactivating enzyme. A reversible inactivation of the maize nitrate reductase has been shown to occur during incubation with NAD(P)H. In contrast to the action of the inactivating enzyme, it is the FADH2-nitrate reductase alone which is inactivated. No inactivation of the Neurospora nitrate reductase was produced by NAD(P)H alone and also in the presence of FAD. The lack of effect of the inactivating enzyme and NAD(P)H on the FADH2-nitrate reductase of Neurospora suggests some differences in its structure or conformation from that of the maize enzyme. A low level of cyanide (0.4 mu M) markedly enhanced the action of NAD(P)H on the maize enzyme; Cyanide at a higher level (6 mu M) did give inactivation of the Neurospora nitrate reductase in the presence of NADPH and FAD. The maize nitrate reductase, when partially inactivated by NADH and cyanide, was not altered as a substrate for the inactivating enzyme. The maize root inactivating enzyme was also shown to inactivate the nitrate reductase (NADH) in the pea leaf. It had no effect on the nitrate reductase from either Pseudomonas denitrificans or Nitrobacter agilis.  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