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.
The ability of paraquat radicals (PQ+.) generated by xanthine oxidase and glutathione reductase to give H2O2-dependent hydroxyl radical production was investigated. Under anaerobic conditions, paraquat radicals from each source caused chain oxidation of formate to CO2, and oxidation of deoxyribose to thiobarbituric acid-reactive products that was inhibited by hydroxyl radical scavengers. This is in accordance with the following mechanism derived for radicals generated by γ-irradiation [H. C. Sutton and C. C. Winterbourn (1984) Arch. Biochem. Biophys.235, 106–115] PQ+. + Fe3+ (chelate) → Fe2+ (chelate) + PQ++ H2O2 + Fe2+ (chelate) → Fe3+ (chelate) + OH? + OH.. Iron-(EDTA) and iron-(diethylenetriaminepentaacetic acid) (DTPA) were good catalysts of the reaction; iron complexed with desferrioxamine or transferrin was not. Extremely low concentrations of iron (0.03 μm) gave near-maximum yields of hydroxyl radicals. In the absence of added chelator, no formate oxidation occurred. Paraquat radicals generated from xanthine oxidase (but not by the other methods) caused H2O2-dependent deoxyribose oxidation. However, inhibition by scavengers was much less than expected for a reaction of hydroxyl radicals, and this deoxyribose oxidation with xanthine oxidase does not appear to be mediated by free hydroxyl radicals. With O2 present, no hydroxyl radical production from H2O2 and paraquat radicals generated by radiation was detected. However, with paraquat radicals continuously generated by either enzyme, oxidation of both formate and deoxyribose was measured. Product yields decreased with increasing O2 concentration and increased with increasing iron(DTPA). These results imply a major difference in reactivity between free and enzymatically generated paraquat radicals, and suggest that the latter could react as an enzyme-paraquat radical complex, for which the relative rate of reaction with Fe3+ (chelate) compared with O2 is greater than is the case with free paraquat radicals.  相似文献   

2.
Rebecca R. Mathis  Olen R. Brown   《BBA》1976,440(3):723-732
Escherichia coli, strain E-26, grown in defined salts medium with glucose as the sole carbon and energy source, contained 1.50 ± 0.16 · 106 molecules of ATP/cell. ATP was extracted with HClO4 and assayed with a Dupont Luminescence Biometer using the luciferin-luciferase assay. Exposure during exponential growth at 37°C to 4.2 atm of oxygen resulted in complete growth cessation within 5 min, and to cyclic changes in cellular ATP concentration over a 2 h period. However, significant decrease in cellular ATP concentration occurred after growth inhibition in hyperbaric oxygen; hence, lack of ATP was not the cause of growth inhibition from oxygen toxicity.  相似文献   

3.
The reduction of ferricytochrome c in the presence of 6-hydroxydopamine/O2 mixtures was examined under various reaction conditions. As the autoxidation of 6-hydroxy-dopamine progressed to completion, there were fluctuations in the net redox reactivity between reducing and oxidizing steady states. This was reflected in a sequence of damped oscillations in the redox state of cytochrome c. Corresponding to the time when 6-hydroxydopamine was 75–100% exhausted, reoxidation of the ferrocytochrome c occurred (prevented by catalase or catalase plus Superoxide dismutase). After the H2O2, in turn, was mostly consumed, the next phase commenced in which the cytochrome c became reduced for a second time. This reductive phase was 52% inhibited by superoxide dismutase. In the subsequent and final phase of the process, a progressive oxidation of cytochrome c lasting at least 24 h was observed. Of the initial reduction of ferricytochrome c, at most 37% can be attributed to direct reduction by 6-hydroxydopamine or its semiquinone. This initial net reduction of cytochrome c was inhibited 51% by superoxide dismutase and 41% by catalase. However, since either catalase or superoxide dismutase inhibited the autoxidation of 6-hydroxydopamine by at least as much as it slowed the reduction of cytochrome c, their effects in slowing the reduction of cytochrome c resulted largely from the decreased production of those free radicals which reduce ferricytochrome c, and only in part from accelerated removal. Elimination of the actions of transition metal ions (whether by passage of the buffer solutions through Chelex 100 resins or by addition of desferrioxamine to the reaction medium) slowed both the reoxidation and rereduction by up to 96%. Addition of mannitol decreased the rate of the first reoxidation by 25% and increased the rate of the rereduction by 7%. In general, the oscillations are explicable in terms of changes in the steady state levels of O2 and H2O2, with metal ions playing a major role and hydroxyl radicals a minor role in both the reoxidation and rereduction.  相似文献   

4.
Superoxide dismutase in leukocytes   总被引:1,自引:0,他引:1  
P Patriarca  P Dri  F Rossi 《FEBS letters》1974,43(3):247-251
  相似文献   

5.
Superoxide dismutase and chilling injury in Chlorella ellipsoidea   总被引:7,自引:0,他引:7  
The relationship between superoxide dismutase (SOD) and chilling injury was examined in chilling-sensitive and chilling-resistant strains of Chlorella ellipsoidea. The sensitive strain contained less SOD than the resistant strain. Moreover, all of the SOD in the sensitive strain was the H2O2-sensitive, iron-containing SOD, whereas most of the SOD in the resistant strain was the H2O2-resistant, manganese-containing SOD. Illumination further enhanced the disparity in SOD content between the sensitive and resistant strains since the SOD in the former declined during illumination, whereas the SOD in the latter strain did not. It was possible to elevate the SOD content of the sensitive strain and to increase the proportion of MnSOD by prior growth in the presence of 50 microM paraquat. The SOD content of the cultures after 5 h of illumination at 4 degrees C fell in the order sensitive strain less than paraquat-induced sensitive strain less than resistant strain. The resistance of these cultures to chilling injury was related to SOD content. This was the case whether resistance was assessed in terms of growth rate after chilling, bleaching of chlorophyll during chilling, or loss of viability during chilling. It thus appears likely that O2- is an agent of chilling injury.  相似文献   

6.
The intrinsic factor receptor from guinea pig ilea has been characterized following purification by affinity chromatography. The purified receptor complexed to intrinsic factor-cobalamin (holo-receptor) had an anhydrous molecular weight of 680,000, a Stokes radius of 10.9 nm, and a sedimentation coefficient of 15.1 S. In contrast, unsaturated receptor (apo-receptor) resolved into distinct "large" and "small" molecular species having, respectively, (i) molecular weights of 700,000 and 350,000; (ii) Stokes radii of 11.1 and 7.06 nm; (iii) sedimentation coefficients of 15.5 S and 11.9 S; (iv) association constants for binding the intrinsic factor-cobalamin complex of 7.3 and 2.5 X 10(10) liters mol-1; and (v) two isoproteins for the larger species (pI's 4.05 and 4.80), and one isoprotein for the smaller species (pI 4.90). A rabbit anti-receptor serum gave only one precipitation line in immunodiffusion against either the large or the small receptor, and each one of these two lines fused completely with the one of two lines which formed with the purified preparation containing both receptor species. Autoradiography of the precipitin lines obtained when the receptor was coupled to intrinsic factor-cyano[57Co]cobalamin demonstrated that both species of receptor were functional. The reaction of complete immunologic identity, the similar electrical properties and similar kinetics for binding to intrinsic factor, and the observed molecular weight differences indicate that the small and large apo-receptors are chemically interrelated, and suggest that the large receptor may consist of two small functional proteins.  相似文献   

7.
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   

8.
The sodium- and potassium-transporting ionophore monensin induces the maturation of Caenorhabditis elegans spermatids to spermatozoa in vitro. Rearrangement of cytoplasm, fusion of membranous organelles with the plasma membrane and growth of pseudopodia, all characteristic of in vivo spermiogenesis, occur within five minutes after exposure to monensin at concentrations of 0.1–1.0 μM. This activation is dependent upon external Na+ and K+ ions but not Ca2+ ions. Monensin-activated spermatozoa have normal morphology and normal amoeboid motility. During activation spermatids twitch and rotate prior to pseudopod extension. Analysis of intermediates by transmission and scanning electron microscopy reveals that the sequence of morphogenetic events leading from the spherical spermatid to the polarized spermatozoan involves microvilli rearrangement and membranous organelle fusion, cytoplasmic polarization, then pseudopod extension.  相似文献   

9.
The permeability of phospholipid membranes to the superoxide anion (O2?) was determined using soybean phospholipid vesicles containing FMN in the internal space. The efflux of O2? generated by the illumination of FMN was so slow that more than 90% of the radicals were spontaneously disproportionated within the vesicles before they could react with cytochrome c at the membrane exterior. The amount of diffused O2? was proportional to the intravesicular concentration of O2? over a range from 1 to 10 μm which was deduced from its disproportionation rate. The permeability coefficient of the phospholipid bilayer for O2? was estimated to be 2.1 × 10?6 cm s?1 at pH 7.3 and 25 ° C. Superoxide dismutase trapped inside vesicles was not reactive with extravesicular O2? unless Triton X-100 was added. O2? generated outside spinach chloroplast thylakoids did not interact with superoxide dismutase or cytochrome c which had been enclosed in the thylakoids. Thus, chloroplast thylakoids also showed little permeability to O2?.  相似文献   

10.
The rates of synthesis and degradation of enolase and total soluble proteins slow with age in the free-living nematode, Turbatrix aceti. The half-lives are 73 and 58 h for soluble protein and enolase, respectively, in young organisms (5 days old). The respective figures are 163 and 161 h for old organisms (22–30 days old). Similar slowing of protein turnover occurs when the organisms are aged by a repeated screening procedure which avoids the use of fluorodeoxyuridine, an inhibitor of DNA synthesis normally added to aging cultures to obtain synchrony. The results support the idea that slowed protein turnover may be responsible for the formation of altered enzymes in old organisms.  相似文献   

11.
William Stillwell  H.Ti Tien 《BBA》1977,461(2):239-252
When chlorophyll-lipid microvesicles under aerobic conditions were illuminated with intense white light; (a) the media became more acidic, (b) the lipid microvesicles bleached from green to yellow and (c) oxygen was taken up from the bathing solution. The bleaching, which was followed spectrophotometrically, resulted in a decrease in the total chlorophyll content as well as the chlorophyll a : b ratio. Some bleaching, which was slowed by the presence of electron donors, occurred in the dark. Water-soluble electron donors were shown to increase the rate of oxygen consumption with the order of effectiveness being; control = KI = ferrocyanide < hydroquinone < thiourea < cysteine < NADH < Fe+2 < ascorbic acid < phenylhydrazine. Chlorophyll mediated electron transfer from donors to oxygen is similar to that of the well known Krasnovsky reactions (in organic solvents and aqueous detergent suspensions) and Mehler reaction (in chloroplast suspension). Electron acceptors and β-carotene had no effect on the oxygen consumption. Lipid-soluble quinones and α-tocopherol affected the oxygen reaction to different extents. The reactions reported here are closely related to those previously described for chlorophyll in organic solvents, “quantasomes” and chloroplasts. The demonstration of these reactions in chlorophyll-lipid microvesicles is an advance in making the chlorophyll-lipid microvesicles a better model of the thylakoid membrane.  相似文献   

12.
Superoxide dismutase: a photochemical augmentation assay.   总被引:21,自引:0,他引:21  
Cell envelope vesicles containing bacteriorhodopsin, prepared from Halobacterium halobium, have previously been shown to accumulate glutamate to high concentration gradients when illuminated. This active transport is energized by a sodium gradient (Naout+ ? Nain+), which arises from Na+-efflux coupled to the light-induced H+-gradient. The oxidation of dimethyl phenylenediamine (DPD) by the vesicles also can drive uphill glutamate transport, and such transport is inhibited by KCN, azide, ionophores, or uncouplers. KT for glutamate is 1.4 × 10?7m under these conditions, as compared to 1.3 × 10?7m for light-induced transport. The respiration-induced transport of glutamate is dependent on high Na+ concentrations on the vesicle exterior and requires low Na+ concentrations in the interior. When Na+ of increasing concentrations is included in the vesicles, transport proceeds with increasing lags, similarly to the case of light-driven transport. In vesicles to which DPD is added first, and then KCN at increasing time intervals (5 to 15 min), glutamate transport occurs after the addition of KCN, with increasing rates, even though respiration is inhibited. This indicates that the energy generated by DPD-oxidation is conserved over several minutes. These results suggest that in the case of respiration-dependent glutamate transport the translocation is also driven by a Na+-gradient; thus, there is a single glutamate transport system independent of the source of energy. The generation of such an Na+-gradient during DPD-oxidation implies that the respiration component involved, cytochrome oxidase, is functionally equivalent to bacteriorhodopsin, which acts as a proton pump.  相似文献   

13.
The embryonic cell lineage of Caenorhabditis elegans has been traced from zygote to newly hatched larva, with the result that the entire cell lineage of this organism is now known. During embryogenesis 671 cells are generated; in the hermaphrodite 113 of these (in the male 111) undergo programmed death and the remainder either differentiate terminally or become postembryonic blast cells. The embryonic lineage is highly invariant, as are the fates of the cells to which it gives rise. In spite of the fixed relationship between cell ancestry and cell fate, the correlation between them lacks much obvious pattern. Thus, although most neurons arise from the embryonic ectoderm, some are produced by the mesoderm and a few are sisters to muscles; again, lineal boundaries do not necessarily coincide with functional boundaries. Nevertheless, cell ablation experiments (as well as previous cell isolation experiments) demonstrate substantial cell autonomy in at least some sections of embryogenesis. We conclude that the cell lineage itself, complex as it is, plays an important role in determining cell fate. We discuss the origin of the repeat units (partial segments) in the body wall, the generation of the various orders of symmetry, the analysis of the lineage in terms of sublineages, and evolutionary implications.  相似文献   

14.
Murine migration inhibitory factor (MIF) produced by concanavalin A-stimulated lymph node cells from C57BL/6 mice was fractionated by Sephadex G-100 gel filtration, density gradient electrophoresis, and isoelectrofocusing in a sucrose density gradient and assayed on in vitro-cultivated bone marrow macrophages from C57BL/6 mice. Two major MIF species, pH3-MIF with an isoelectric point of 3.0–4.3 and pH5-MIF with an isoelectric point of 4.6 to 5.2, were obtained. The similarity of murine MIF to guinea pig and human MIF is discussed.  相似文献   

15.
Peroxidase activity and hydrogen peroxide concentrations were measured by a chemiluminescent method based upon the reduction of peroxidase Compounds I and II by both eosin and EDTA. Eosin excess present in the reaction mixture was excited during the reaction to its fluorescent state. This bioenergized method allows calculation of peroxidase concentration in the range of 10?12 to 10?7m, and hydrogen peroxide concentration in the range of 10?9 to 10?5m. This approach has been applied to the estimation of peroxidase activity in human red cell membranes and hydrogen peroxide formation in the peroxidase-catalyzed oxidation of glutathione.  相似文献   

16.
The ability of a number of stimulants to activate an oxidative burst (OB) in oil-elicited guinea pig peritoneal exudate macrophages (MPs) was examined. The parameters of the OB were the generation and extracellular release of Superoxide anions (O2?) and hydrogen peroxide (H2O2). We found that: (1) The cocarcinogen and skin irritant phorbol myristate acetate (PMA) was the most potent OB activator—The weak cocarcinogen 4-O-methyl PMA was a proportionally less effective OB activator; (2) The lectins concanavalin A (Con A) and wheat germ agglutinin (WGA), but not soybean, Lotus, and pokeweed lectins, were also quite effective OB activators—The ability of Con A to stimulate O2? production was abolished by succinylation and could be prevented by the presence of α-methyl-D-mannoside; (3) Other stimulators of an OB in MPs were: N-formyl-methionyl peptides, opsonized zymosan, the Ca2+ ionophore A23187, phospholipase C, NaF, antimacrophage antibody, microtubule-disrupting drugs, and sodium nitroprusside—O2? generation induced by A23187 (but not that stimulated by PMA) was dependent on extracellular Ca2+; (4) The amount of O2? produced per cell was higher at low cell densities; (5) The addition of Superoxide dismutase (SOD) to the medium totally prevented the detection of O2? and augmented twice the amount of H2O2 recovered; (6) Pretreatment of MPs with the SOD inhibitor sodium diethyldithiocarbamate had no effect on the release of O2? but blocked H2O2 release in a dose-dependent manner. These data were interpreted as indicating that the bulk of H2O2 was derived by enzymatic dismutation of O2?; (7) The common mechanism by which such a variety of stimuli provoke an OB in MPs was not elucidated. No evidence was found to suggest a role for a cyclic nucleotide messenger.  相似文献   

17.
18.
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   

19.
The autoxidation of 6-hydroxydopamine (6HODA) was virtually blocked (k2 less than 10(-15) M-1 S-1 at pH 8.0, ionic strength 0.04) by the simultaneous presence of diethylenetriaminepentaacetic acid (DTPA), catalase, and superoxide dismutase (SOD). No quinone product or oxygen consumption was detectable after 12 min under these conditions. Thus, if 6HODA is to react with molecular oxygen at a measurable rate, some other redox species is required as a coreductant. The subsequent addition of formate or mannitol proved capable of overcoming the total inhibition induced by the mixture of catalase, SOD, and DTPA. The simplest interpretation of the data is that most of the autoxidation of 6HODA, as commonly observed, involves successive reduction of a series of metal-bound species of oxygen; the actual transfer of electrons occurring within a ternary reductant-metal-oxygen transition state.  相似文献   

20.
Panicum milioides represents the first well-documented example of a higher plant species with reduced photorespiration and O2 inhibition of photosynthesis. We have investigated the biochemical mechanism(s) involved in reducing O2 sensitivity of photosynthesis in this species by parallel enzyme inhibitor experiments with thin leaf slices of P. milioides and C3 and C4Panicum species. The reduced O2 sensitivity of net photosynthesis in P. milioides gradually increased with increasing concentrations of the phosphoenolpyruvate carboxylase (EC 4.1.1.31) inhibitors, maleate and malonate. At saturating levels of inhibitor, photosynthesis in 2% O2 was decreased by about 18%, and the inhibitory effects of both 21% O2 and 49% O2 were identical to those observed with a C3Panicum species in the absence or presence of inhibitor. A significant potential for C4 photosynthesis in P. milioides, compared to its complete absence in a C3Panicum species, was demonstrated on the basis of: (a) a coupling of leaf slice CO2 fixation by phosphoenolpyruvate carboxylase with the C3 cycle; (b) NAD-malic enzyme (EC 1.1.1.39)-dependent aspartate and malate decarboxylation in leaf slices; (c) a full complement of C4 cycle enzymes in leaf extracts, including pyruvate, Pi dikinase (EC 2.7.9.1) and NAD-malic enzyme; and (d) Kranz-like leaf anatomy with numerous plasmodesmata traversing the mesophyll-bundle sheath interfacial cell wall. These data indicate that the reduced photorespiration and O2 inhibition of photosynthesis in P. milioides is due to phosphoenolpyruvate carboxylase participation, possibly by creating a limited C4-like CO2 pump, rather than an altered ribulose 1,5-bisphosphate carboxylase-oxygenase (EC 4.1.1.39).  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