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 0 毫秒
1.
2.
Complex I (NADH-ubiquinone reductase) and Complex III (ubiquinol-cytochrome c reductase) supplemented with NADH generated O2? at maximum rates of 9.8 and 6.5 nmol/min/mg of protein, respectively, while, in the presence of superoxide dismutase, the same systems generated H2O2 at maximum rates of 5.1 and 4.2 nmol/min/mg of protein, respectively. H2O2 was essentially produced by disproportionation of O2?, which constitutes the precursor of H2O2. The effectiveness of the generation of oxygen intermediates by Complex I in the absence of other specific electron acceptors was 0.95 mol of O2? and 0.63 mol of H2O2/mol of NADH. A reduced form of ubiquinone appeared to be responsible for the reduction of O2 to O2?, since (a) ubiquinone constituted the sole common major component of Complexes I and III, (b) H2O2 generation by Complex I was inhibited by rotenone, and (c) supplementation of Complex I with exogenous ubiquinones increased the rate of H2O2 generation. The efficiency of added quinones as peroxide generators decreased in the order Q1 > Q0 > Q2 > Q6 = Q10, in agreement with the quinone capacity of acting as electron acceptor for Complex I. In the supplemented systems, the exogenous quinone was reduced by Complex I and oxidized nonenzymatically by molecular oxygen. Additional evidence for the role of ubiquinone as peroxide generator is provided by the generation of O2? and H2O2 during autoxidation of quinols. In oxygenated buffers, ubiquinol (Q0H2), benzoquinol, duroquinol and menadiol generated O2? with k3 values of 0.1 to 1.4 m? · s?1 and H2O2 with k4 values of 0.009 to 4.3 m?1 · s?1.  相似文献   

3.
[14C]Dihydroxyacetone phosphate labeled in either the C-1 or C-3 position was enzymatically synthesized, isolated, and utilized as a substrate for crystalline methylglyoxal synthase purified from Proteus vulgaris. After reaction with the enzyme, the methyl carbon of methylglyoxal3 was identified as CHI3 by the iodoform reaction. The labeling pattern revealed that C-1 is dephosphorylated and reduced to the methyl group, while C-3 is oxidized to the aldehyde. Methylglyoxal was found to be noncompetitive with respect to dihydroxyacetone phosphate, while inorganic phosphate was competitive and transformed the dihydroxyacetone phosphate saturation kinetics from hyperbolic to sigmoidal. The enzyme was inactivated by freezing, and phosphate stabilized the enzyme toward both cold- and heat-induced denaturation. The phosphate moiety of the substrate appears to be required for binding, since the synthase is competitively inhibited by a variety of phosphorylated compounds but not by their nonphosphorylated counterparts. Based on these observations, and the ability of bromo- and iodoacetol phosphates to act as active-site reagents, a mechanism is proposed in which the enzyme first catalyzes the keto-enol tautomerization to the hydrogen-bonded enol which facilitates the internal oxidation-reduction and phosphoester cleavage with CO bond breakage.  相似文献   

4.
The chemiluminescence associated with peroxidation of luminol in buffered aqueous solution is a complex process involving several intermediates. It can be inhibited by removal of oxygen from the incubation medium. Superoxide radical is both an intermediate in this reaction and an essential component in light-producing steps. The importance of O2? in propagating this reaction was shown by the inhibition of luminescence by superoxide dismutase. A mechanism was proposed which is consistent with the data. It appears likely that the diverse biological effects of peroxidases are largely due to the reactivities of these intermediates and products.  相似文献   

5.
Superoxide generation, assessed as the rate of acetylated cytochrome c reduction inhibited by superoxide dismutase, by purified NADPH cytochrome P-450 reductase or intact rat liver microsomes was found to account for only a small fraction of their respective NADPH oxidase activities. DTPA-Fe3+ and EDTA-FE3+ greatly stimulated NADPH oxidation, acetylated cytochrome c reduction, and O(2) production by the reductase and intact microsomes. In contrast, all ferric chelates tested caused modest inhibition of acetylated cytochrome c reduction and O(2) generation by xanthine oxidase. Although both EDTA-Fe3+ and DTPA-Fe3+ were directly reduced by the reductase under anaerobic conditions, ADP-Fe3+ was not reduced by the reductase under aerobic or anaerobic conditions. Desferrioxamine-Fe3+ was unique among the chelates tested in that it was a relatively inert iron chelate in these assays, having only minor effects on NADPH oxidation and/or O(2) generation by the purified reductase, intact microsomes, or xanthine oxidase. Desferrioxamine inhibited microsomal lipid peroxidation promoted by ADP-Fe3+ in a concentration-dependent fashion, with complete inhibition occurring at a concentration equal to that of exogenously added ferric iron. The participation of O(2) generated by the reductase in NADPH-dependent lipid peroxidation was also investigated and compared with results obtained with a xanthine oxidase-dependent lipid peroxidation system. NADPH-dependent peroxidation of either phospholipid liposomes or rat liver microsomes in the presence of ADP-Fe3+ was demonstrated to be independent of O(2) generation by the reductase.  相似文献   

6.
Effects of monensin, a carboxylic ionophore, on intracellular transport of albumin were studied in primary cultured rat hepatocytes. The lag time after which newly synthesized albumin first appeared in medium was 10 min in the control cells, while it was prolonged to 40 min in the monensin-treated cells. In addition, this inhibition of secretion by monensin was accompanied by an intracellular accumulation of proalbumin. The results strongly suggest that monensin arrests the intracellular transport of proalbumin before the site where its conversion takes place.  相似文献   

7.
The scavenging of superoxide radical by manganous complexes: in vitro   总被引:22,自引:0,他引:22  
Dialyzable manganese has been shown to be present in millimolar concentrations within cells of Lactobacillus plantarum and related lactic acid bacteria. This unusual accumulation of Mn appears to serve the same function as Superoxide dismutase (SOD), conferring hyperbaric oxygen and Superoxide tolerance on these SOD-free organisms. The form of the Mn in the lactic acid bacteria and the mechanisms whereby it protects the cell from oxygen damage are unknown. This report examines the mechanisms by which Mn catalytically scavenges O2?, both in the xanthine oxidase/cytochrome c SOD assay and in a number of in vitro systems relevant to the in vivo situation. In all the reaction mixtures examined, Mn(II) is first oxidized by O2? to Mn(III), and H2O2 is formed. In pyrophosphate buffer the Mn(III) thus formed is re-reduced to Mn(II) by a second O2?, making the reaction a true metal-catalyzed dismutation like that catalyzed by SOD. Alternatively, if the reaction takes place in orthophosphate or a number of other buffers, the Mn(III) is preferentially reduced largely by reductants other than O2?, such as thiols, urate, hydroquinone, or H2O2. H2O2, a common product of the lactic acid bacteria, reacted rapidly with Mn(III) to form O2, apparently without intermediate O2 release. Free hexaquo Mn(II) ions were shown by electron spin resonance spectroscopy and activity assays in noncomplexing buffers to be poorly reactive with O2?. In contrast, Mn(II) formed complexes having a high catalytic activity in scavenging O2? with a number of organic acids, including malate, pyruvate, propionate, succinate, and lactate, with the Mn-lactate complex showing the greatest activity.  相似文献   

8.
Paraquat (10–30 μm) exerted a dose-dependent and light-dependent toxicity on Chlorella sorokiniana. Paraquat was also seen to increase the superoxide dismutase content of these cells and to cause the appearance of a new electrophoretically distinct isozyme. Cells grown in the absence of paraquat contained one manganese-superoxide dismutase and two iron-superoxide dismutases, while the paraquat-grown cells contained an additional manganese-superoxide dismutase. Cells which were grown in the presence of 25 μm Paraquat, and which therefore possessed elevated levels of superoxide dismutase, were resistant to 30 μm Paraquat, whereas control cells were bleached and killed by this level of Paraquat. Electron micrography and chemical analysis revealed that Paraquat decreased the starch content of the cells and caused a failure of dividing cells to separate. It appears that Paraquat increases the photoproduction of O2? in C. sorokiniana and that an increase in the cell content of superoxide dismutase is an adaptive response which provides protection against this herbicide.  相似文献   

9.
Deoxymannojirimycin (dMM) was tested as an inhibitor of the processing of the oligosaccharide portion of viral and cellular N-linked glycoproteins. The NWS strain of influenza virus was grown in MDCK cells in the presence of various amounts of dMM, and the glycoproteins were labeled by the addition of 2-[3H]mannose to the medium. At levels of 10 micrograms/ml dMM or higher, most of the viral glycopeptides became susceptible to digestion by endoglucosaminidase H, and the liberated oligosaccharide migrated mostly like a Hexose9GlcNAc on a calibrated column of Bio-Gel P-4. This oligosaccharide was characterized as a typical Man9GlcNAc by a variety of chemical and enzymatic procedures. Deoxymannojirimycin gave rise to similar oligosaccharide structures in the cellular glycoproteins. In both the viral and the cellular glycoproteins, this inhibitor caused a significant increase in the amount of [3H]mannose present in the glycoproteins. Deoxymannojirimycin did not inhibit the incorporation of [3H]leucine into protein in MDCK cells, nor did it affect the yield or infectivity of NWS virus particles. However, its effect on mannose incorporation into lipid-linked saccharides depended on the incubation time, the virus strain, and the cell line. Thus, high concentrations of dMM showed some inhibition of mannose incorporation into lipid-linked oligosaccharides with the NWS strain in a 3-h incubation, but no inhibition was observed after 48 h of incubation. On the other hand, the PR8 strain was much more sensitive to dMM inhibition, and mannose incorporation into lipid-linked oligosaccharides was strongly inhibited when the virus was raised in chick embryo cells, but less inhibition was observed when this virus was grown in MDCK cells. Nevertheless, in these cases also, the major oligosaccharide structure in the glycoproteins was the Man9GlcNAc2 species.  相似文献   

10.
Vanadate or molybdate strongly accelerate the cooxidation of NADH, or of reduced nicotinamide mononucleotide, by the xanthine oxidase plus xanthine reaction. Superoxide dismutase eliminated the effect of vanadate or molybdate, while catalase was without effect. It follows that vanadate or molybdate accelerate the oxidation of dihydropyridines by O-2. A stoichiometry of 4 NADH oxidized per O-2 introduced suggests a chain reaction for which a mechanism is proposed. These results provide an explanation for the reported stimulation, by vanadate, of NADH oxidation by biological membranes.  相似文献   

11.
A model lipid peroxidation system dependent upon the hydroxyl radical, generated by Fenton's reagent, was compared to another model system dependent upon the enzymatic generation of superoxide by xanthine oxidase. Peroxidation was studied in detergent-dispersed linoleic acid and in phospholipid liposomes. Hydroxyl radical generation by Fenton's reagent (FeCl2 + H2O2) in the presence of phospholipid liposomes resulted in lipid peroxidation as evidenced by malondialdehyde and lipid hydroperoxide formation. Catalase, mannitol, and Tris-Cl were capable of inhibiting activity. The addition of EDTA resulted in complete inhibition of activity when the concentration of EDTA exceeded the concentration of Fe2+. The addition of ADP resulted in slight inhibition of activity, however, the activity was less sensitive to inhibition by mannitol. At an ADP to Fe2+ molar ratio of 10 to 1, 10 mm mannitol caused 25% inhibition of activity. Lipid peroxidation dependent on the enzymatic generation of superoxide by xanthine oxidase was studied in liposomes and in detergent-dispersed linoleate. No activity was observed in the absence of added iron. Activity and the apparent mechanism of initiation was dependent upon iron chelation. The addition of EDTA-chelated iron to the detergent-dispersed linoleate system resulted in lipid peroxidation as evidenced by diene conjugation. This activity was inhibited by catalase and hydroxyl radical trapping agents. In contrast, no activity was observed with phospholipid liposomes when iron was chelated with EDTA. The peroxidation of liposomes required ADP-chelated iron and activity was stimulated upon the addition of EDTA-chelated iron. The peroxidation of detergent-dispersed linoleate was also enhanced by ADP-chelated iron. Again, this peroxidation in the presence of ADP-chelated iron was not sensitive to catalase or hydroxyl radical trapping agents. It is proposed that initiation of superoxide-dependent lipid peroxidation in the presence of EDTA-chelated iron occurs via the hydroxyl radical. However, in the presence of ADP-chelated iron, the participation of the free hydroxyl radical is minimal.  相似文献   

12.
Rats fed ethanol (1.74 +/- 0.12 g/day/100 g body wt for 12 weeks) showed a 45% increased microsomal production of O-2 (2.23 +/- 0.14 nmol/min/mg protein) and a 28% increased content of endoplasmic reticulum protein (26.8 +/- 1.4 mg/g liver). This could lead, at substrate saturation, to a 86% increased cytosolic production of O-2 which is not compensated by cytosolic superoxide dismutase levels that remain normal. It is claimed that this unbalance between O-2 production and superoxide dismutase leads to a peroxidative stress in agreement with the 54% increased spontaneous liver chemiluminescence (37 +/- 2 cps/cm2) measured in the ethanol-treated rats. Hydroperoxide-induced chemiluminescence was 57, 43, and 28% higher, respectively, in homogenates, mitochondria, and microsomes isolated from ethanol-treated rats as compared with controls. Vitamins E and A were more effective inhibitors of the hydroperoxide-stimulated chemiluminescence in the liver homogenates from ethanol-treated rats as compared with the effect on the homogenates from control animals. The results are consistent with a peroxidative stress in chronic alcoholism leading to increased lipoperoxidation and decreased levels of antioxidants.  相似文献   

13.
The rate constants for hydrolysis of the enantiomers of amino acid p-nitrophenyl esters catalyzed by bifunctional comicellar catalysts containing the imidazolyl and hydroxyl groups have been determined at pH 7.30, 0.02 m phosphate buffer, and 25°C. The kinetic analysis suggests a reaction scheme which involves acylation followed by deacylation at the imidazolyl group. Although no appreciable cooperative catalytic efficiencies are observed between the bifunctional groups in the acylation step, it is found that the deacylation rates are thus accelerated by surfactant hydroxyl groups, and some of the stereoselective acyl transfer reaction occurs from the imidazolyl to the hydroxyl group in optically active comicellar systems.  相似文献   

14.
Rabbit antibodies to bovine superoxide dismutase have been produced and used to develop a double-antibody solid phase radioimmunoassay for the enzyme. The assay is sensitive and highly specific for the bovine enzyme, showing no cross-reactivity with the murine or human superoxide dismutases. It has been applied to the quantitation of exogenous enzyme in serum and extracts of mouse cells and tissues.  相似文献   

15.
16.
For the pork heart, extramitochondrial aspartate aminotransferase (EC 2.6.1.1), the “half-reaction” equilibrium, amino acid + phosphopyridoxal enzyme ? keto acid + phosphopyridoxamine enzyme, is displaced in favor of the phosphopyridoxamine enzyme by the addition of halide ions. The order of effectiveness is I? > Br? > Cl? > F?. A kinetic analysis of this equilibrium with alanine and pyruvate as substrates showed that halide ions (0.01–0.1 m) both increase the rate of the forward reaction and decrease the rate of the reverse reaction. Chloride ions decrease the rate of the reverse reaction by competitively inhibiting the formation of an intermediate enzyme-pyruvate complex. The rate of the forward reaction is proportional to the alanine concentration up to 0.5 m alanine, indicating that the initial combination of alanine with the enzyme is the rate-limiting step in this direction. The activation by anions must therefore involve the initial binding of the substrates to the enzyme. Chloride ions also cause a marked activation of the enzyme in the presence of glutarate by displacing the inhibitory glutarate from the enzyme. These results indicate that some enzyme activations may be due to relieving a preexisting inhibition by ligand substitution reactions. The finding that aspartate aminotransferase has an anion-sensitive “half-reaction” equilibrium, or redox potential, suggests that transaminases may function in both active and passive transport of anions across membranes.  相似文献   

17.
The relationship between the delta 9-desaturase activity of the psychrophilic bacterium Micrococcus cryophilus grown at different temperatures and the physical state of its membrane lipids as measured by ESR spectroscopy has been studied. Arrhenius plots of desaturase activity were biphasic with a discontinuity at a temperature which depended upon the bacterial growth temperature. Changes in the desaturase activation energy, which increased as the growth temperature was lowered, are discussed in the context of membrane lipid fluidity adaptation to changing environmental temperature. The fluidity of membranes and isolated lipids was measured using nitroxide-labeled fatty acids. The spectra of 2-(10-carboxydecyl)-2-hexyl-4,4-dimethyl-3-oxazolidinoxyl in membranes indicated that there were two lipid environments within the membrane whose relative proportions were dependent both on temperature of measurement and on bacterial growth temperature. In contrast, 2-(3-carboxypropyl)-4,4-dimethyl-2-tridecyl-3-oxazolidinoxyl spectra showed a single lipid environment and plots of log order parameter (S3) vs 1/T were biphasic with inflexion temperatures which were closely related to the bacterial growth temperature. As with membranes, plots of log S3 vs 1/T for total lipids, phosphatidylglycerol and cardiolipin, but not phosphatidylethanolamine, were biphasic and showed inflexions which correlated well with bacterial growth temperature. These results are interpreted as being consistent with a location for the desaturase within the bulk lipid of the membrane rather than in association with specific lipid types.  相似文献   

18.
The effect of hypolipidemic drugs, WY14643 and DH990, on plant lipid metabolism has been studied. The total incorporation of [14C]acetate into lipids was inhibited by addition of both drugs to aged potato (Solanum tuberosum) tuber discs, spinach (Spinacia oleracea) leaves, and spinach chloroplasts, while the incorporation in Chlorella vulgaris cells was affected only by DH990. Moreover, DH990 inhibited the incorporation of 14C-labeled fatty acids into phosphatidylcholine and phosphatidylethanolamine of potato discs, and decreased the incorporation into phosphatidylglycerol of Chlorella cells. DH990 inhibited the formation of polyunsaturated fatty acids in potato discs, Chlorella cells, and spinach leaves, whereas WY14643 had no effect on the formation of these fatty acids. Stearoyl-ACP desaturase from safflower (Carthamus tinctorius) seeds was very sensitive to both drugs, especially DH990, which completely blocked the activity at 2 mM levels. When safflower lysophospholipid acyltransferases were solubilized by detergent treatment, only DH990 inhibited the incorporation of [14C]oleoyl-CoA into lysophosphatidylcholine or lysophosphatidylethanolamine. Both drugs inhibited fatty acid synthesis from [14C]malonyl-CoA in the microsomal fraction from safflower seeds, but only DH990 inhibited FAS activity in the soluble fraction; both drugs inhibited severely the formation of stearic acid. Both acetyl-CoA carboxylase and acetyl-CoA synthetase were sensitive to both drugs.  相似文献   

19.
20.
The dissociation of the extracellular hemoglobin of Tubifex tubifex at alkaline and acid pH, and its reassociation upon return to neutral pH, was investigated using gel filtration, ultracentrifugation, and polyacrylamide gel electrophoresis in sodium dodecyl sulfate (SDS-PAGE). Tubifex hemoglobin dissociated at pH above 8 and below 6; both dissociations appeared to be equilibrium processes. The extent of dissociation increased as the pH moved away from neutrality; although dissociation was virtually complete at pH 11, its extent at acid pH did not exceed 50–60% at pH 4. Ca(II), Mg(II), and Sr(II) cations over the range 1–100 mm decreased the extent of the dissociation only at alkaline pH. The visible absorption spectrum of the oxyhemoglobin remained unaltered in the pH range 4–9. At more extreme pH, it changed with time, altering irreversibly to that of the aquo ferri form. Gel filtration of the hemoglobin at both extremes of pH showed that it dissociated into two heme-containing fragments; one consisting of subunit 1 (Mr ~ 17,000) and the other containing subunits 2, 3, and 4 of the hemoglobin (Mr ~ 60,000). Upon return to neutral pH, the dissociated fragment reassociated to the extent of 50 to 80% to whole hemoglobin molecules. The reassociation decreased with increase in alkaline pH, and with decrease in acid pH to which the hemoglobin had been exposed; it increased in the presence of Ca(II), Sr(II), and Mg(II) only subsequent to dissociation at alkaline pH. The SDS-PAGE patterns, gel-filtration elution volumes, and α-helical contents, determined from circular dichroism at 222 nm, of the reassociated whole molecules were identical to those of the native hemoglobin.  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