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 46 毫秒
1.
The abc1 /coq8 gene deletion respiratory-deficient mutant NBp17 of fission yeast Schizosaccharomyces pombe displayed a phenotypic fermentation pattern with enhanced production of glycerol and acetate, and also possessed oxidative stress-sensitive phenotypes to H2O2, menadione, tBuOOH, Cd2+, and chromate in comparison with its parental respiratory-competent strain HNT. As a consequence of internal stress-inducing mutation, adaptation processes to restore the redox homeostasis of mutant NBp17 cells were detected in minimal glucose medium. Mutant NBp17 produced significantly increased amounts of O2•− and H2O2 as a result of the decreased internal glutathione concentration and the only slightly increased glutathione reductase activity. The Cr(VI) reduction capacity and hence the OH production ability were decreased. The mutant cells demonstrated increased specific activities of superoxide dismutases and glutathione reductase (but not catalase) to detoxify at least partially the overproduction of reactive oxygen species. All these features may be explained by the decreased redox capacity of the mutant cells. Most notably, mutant NBp17 hyperaccumulated yellow CdS.  相似文献   

2.
Treatment of bovine pulmonary artery smooth muscle with the O2•− generating system hypoxanthine plus xanthine oxidase stimulated MMP-2 activity and PKC activity; and inhibited Na+ dependent Ca2+ uptake in the microsomes. Pretreatment of the smooth muscle with SOD (the O2•− scavenger) and TIMP-2 (MMP-2 inhibitor) prevented the increase in MMP-2 activity and PKC activity, and reversed the inhibition of Na+ dependent Ca2+ uptake in the microsomes. Pretreatment with calphostin C (a general PKC inhibitor) and rottlerin (a PKCδ inhibitor) prevented the increase in PKC activity and reversed O2•− caused inhibition of Na+ dependent Ca2+ uptake without causing any change in MMP-2 activity in the microsomes of the smooth muscle. Treatment of the smooth muscle with the O2•− generating system revealed, respectively, 36 kDa RACK-1 and 78 kDa PKCδ immunoreactive protein profile along with an additional 38 kDa immunoreactive fragment in the microsomes. The 38 kDa band appeared to be the proteolytic fragment of the 78 kDa PKCδ since pretreatment with TIMP-2 abolished the increase in the 38 kDa immunoreactive fragment. Co-immunoprecipitation of PKCδ and RACK-1 demonstrated O2•− dependent increase in PKCδ-RACK-1 interaction in the microsomes. Immunoblot assay elicited an immunoreactive band of 41 kDa Giα in the microsomes. Treatment of the smooth muscle tissue with the O2•− generating system causes phosphorylation of Giα in the microsomes and pretreatment with TIMP-2 and rottlerin prevented the phosphorylation. Pretreatment of the smooth muscle tissue with pertussis toxin reversed O2•− caused inhibition of Na+ dependent Ca2+ uptake without affecting the protease activity and PKC activity in the microsomes. We suggest the existence of a pertussis toxin sensitive G protein mediated mechanism for inhibition of Na+ dependent Ca2+ uptake in microsomes of bovine pulmonary artery smooth muscle under O2•− triggered condition, which is regulated by PKCδ dependent phosphorylation and sensitive to TIMP-2 for its inhibition. (Mol Cell Biochem xxx: 107–117, 2005)  相似文献   

3.
Available data indicate that superoxide anion (O2•− ) is released from mitochondria, but apart from VDAC (voltage dependent anion channel), the proteins involved in its transport across the mitochondrial outer membrane still remain elusive. Using mitochondria of the yeast Saccharomyces cerevisiae mutant depleted of VDAC (Δpor1 mutant) and the isogenic wild type, we studied the role of the TOM complex (translocase of the outer membrane) in the efflux of O2•− from the mitochondria. We found that blocking the TOM complex with the fusion protein pb2-DHFR decreased O2•− release, particularly in the case of Δpor1 mitochondria. We also observed that the effect of the TOM complex blockage on O2•− release from mitochondria coincided with the levels of O2•− release as well as with levels of Tom40 expression in the mitochondria. Thus, we conclude that the TOM complex participates in O2•− release from mitochondria.  相似文献   

4.
Stored product mites can often infest stored products, but currently there is little information regarding the efficacy of pesticides that can be used for control. In this study we evaluated several common pesticides formulated from single active ingredients (a.i.) or commercially available mixtures (chlorpyrifos, deltamethrin, beta-cyfluthrin, and a combination of deltamethrin and S-bioallethrin), plus an acaricide composed of permethrin, pyriproxyfen and benzyl benzolate, for efficacy against Acarus siro, Tyrophagus putrescentiae, and Aleuroglyphus ovatus. The pesticides were incorporated into the mite diets in a dose range of 10–1000 μg a.i. g−1 diet. Concentrations for suppression of 50 and 90% population growth and eradication (rC0) of mites were fit to linear regression models. None of the tested pesticides gave complete eradication of A. siro, which was the most tolerant of the three mite species tested. The most effective pesticide Allergoff 175 CS was a combination product (a nano-capsule suspension of permethrin, pyriproxyfen and benzyl benzolate) labeled for dust mites, with rC0 range of 463–2453 μg a.i. (permethrin) g−1 diet depending on the species. Least effective were chlorpyrifos and deltamethrin.  相似文献   

5.
Upon mitogen sensitization, lymphocytes undergo proliferation by oxyradical-based mechanisms. Through continuous resting–restimulation cycles, lymphocytes accumulate auto-induced oxidative lesions which lead to cell dysfunction and limit their viability. Astaxanthin (ASTA) is a nutritional carotenoid that shows notable antioxidant properties. This study aims to evaluate whether the in vitro ASTA treatment can limit oxyradical production and auto-oxidative injury in human lymphocytes. Activated lymphocytes treated with 5 μM ASTA showed immediate lower rates of O2•−/H2O2 production whilst NO and intracellular Ca2+ levels were concomitantly enhanced (≤4 h). In long-term treatments (>24 h), the cytotoxicity test for ASTA showed a sigmoidal dose–response curve (LC50 = 11.67 ± 0.42 μM), whereas higher activities of superoxide dismutase and catalase in 5 μM ASTA-treated lymphocytes were associated to significant lower indexes of oxidative injury. On the other hand, lower proliferative scores of ASTA lymphocytes might be a result of diminished intracellular levels of pivotal redox signaling molecules, such as H2O2. Further studies are necessary to establish the ASTA-dose compensation point between minimizing oxidative damages and allowing efficient redox-mediated immune functions, such as proliferation, adhesion, and oxidative burst.  相似文献   

6.
Cadmium-induced initial changes in the production of reactive oxygen species (ROS) and antioxidant mechanism were investigated in soybean (Glycine max L. cv. Don Mario 4800 RR) leaves. Whole plants (WP) and plants without roots (PWR) were exposed to 0.0, 10.0 and 40.0 μM Cd for 0, 4, 6 and 24 h. Compared to PWR, a higher level of endogenous Cd in WP was associated with a lower oxidative stress measured in terms of lipid peroxidation. Furthermore, O2 •− content decreased in the leaves of Cd-treated WP, whereas it increased in those of Cd-treated PWR. Although O2 •− accumulation in PWR was associated with a decrease in superoxide dismutase (SOD) activity, O2 •− diminution in WP leaves was not related to any increase in SOD activity. H2O2 content increased in the leaves of both Cd-treated WP and PWR, and it was concomitant with a corresponding decline in catalase (CAT) and ascorbate peroxidase (APX) activities. When diphenyl iodonium (DPI), an inhibitor of NADPH oxidase, was added, H2O2 content remained unchanged in Cd-treated WP, suggesting that NADPH oxidase does not participate in the early hours of Cd toxicity. Taken together, our results showed that early ROS evolution and oxidative damage were different in WP and PWR. This suggests that the response in soybean leaves during the early hours of Cd toxicity is probably modulated by the root.  相似文献   

7.
Summary.  Methyl-jasmonate (MeJA) has been proposed to be involved in the evocation of defense reactions, as the oxidative burst in plants, substituting the elicitors or enhancing their effect. 48 h dark- and sterilely cultured (axenic) aeroponic sunflower seedling roots excised and treated with different concentrations of MeJA showed a strong and quick depression of the H+ efflux rate, 1.80 μM MeJA totally stopping it for approximately 90 min and then reinitiating it again at a lower rate than controls. These results were wholly similar to those obtained with nonsterilely cultured roots and have been interpreted as mainly based on H+ consumption for O2 •− dismutation to H2O2. Also K+ influx was strongly depressed by MeJA, even transitorily reverting to K+ efflux. These results were consistent with those associated to the oxidative burst in plants. MeJA induced massive H2O2 accumulation in the middle lamella and intercellular spaces of both the root cap cells and the inside tissues of the roots. The native acidic extracellular peroxidase activity of the intact (nonexcised) seedling roots showed a sudden enhancement (by about 52%) after 5 min of MeJA addition, maintained for approximately 15 min and then decaying again to control rates. O2 uptake by roots gave similar results. These and other results for additions of H2O2 or horseradish peroxidase, diphenylene iodonium, and sodium diethyldithiocarbamate trihydrate to the reaction mixture with roots were all consistent with the hypothesis that MeJA induced an oxidative burst, with the generation of H2O2 being necessary for peroxidase activity. Results with peroxidase activity of the apoplastic fluid were in accordance with those of the whole root. Finally, MeJA enhanced NADH oxidation and inhibited hexacyanoferrate(III) reduction by axenic roots, and diphenylene iodonium cancelled out these effects. Redox activities by CN- preincubated roots were also studied. All these results are consistent with the hypothesis that MeJA enhanced the NAD(P)H oxidase of a redox chain linked to the oxidative burst, so enhancing the generation of O2 •− and H2O2, O2 uptake, and peroxidase activity by roots. Received July 12, 2002; accepted October 2, 2002; published online May 21, 2003 RID="*"  相似文献   

8.
To characterise the NADH oxidase activity of both xanthine dehydrogenase (XD) and xanthine oxidase (XO) forms of rat liver xanthine oxidoreductase (XOR) and to evaluate the potential role of this mammalian enzyme as an O2 •− source, kinetics and electron paramagnetic resonance (EPR) spectroscopic studies were performed. A steady-state kinetics study of XD showed that it catalyses NADH oxidation, leading to the formation of one O2 •− molecule and half a H2O2 molecule per NADH molecule, at rates 3 times those observed for XO (29.2 ± 1.6 and 9.38 ± 0.31 min−1, respectively). EPR spectra of NADH-reduced XD and XO were qualitatively similar, but they were quantitatively quite different. While NADH efficiently reduced XD, only a great excess of NADH reduced XO. In agreement with reductive titration data, the XD specificity constant for NADH (8.73 ± 1.36 μM−1 min−1) was found to be higher than that of the XO specificity constant (1.07 ± 0.09 μM−1 min−1). It was confirmed that, for the reducing substrate xanthine, rat liver XD is also a better O2 •− source than XO. These data show that the dehydrogenase form of liver XOR is, thus, intrinsically more efficient at generating O2 •− than the oxidase form, independently of the reducing substrate. Most importantly, for comparative purposes, human liver XO activity towards NADH oxidation was also studied, and the kinetics parameters obtained were found to be very similar to those of the XO form of rat liver XOR, foreseeing potential applications of rat liver XOR as a model of the human liver enzyme.  相似文献   

9.
Summary The effect of low concentrations of hydrogen peroxide (H2O2) (5 × 10−7−9.5 × 10−7 M) on cell growth and antibody production was investigated with murine hybridoma cells (Mark 3 and anti-hPL) in culture. Cell growth, measured by flow cytometry with morphological parameters, was significantly stimulated by H2O2 (8 × 10−7 M) but H2O2 concentration of 7 × 10−6 M and above increased cell death. H2O2 stimulation of antibody production was nonsignificant. The metabolism of cells treated with 8 × 10−7 or 1 × 10−5 M H2O2 was similar to that of the control in terms of glucose and glutamine consumption, lactate and ammonia production, and amino acid concentrations in the medium. The concentrations of lactate dehydrogenase, a marker of cell death, in test and control cells were similar. However, concentrations of intracellular free radicals measured by flow cytometry with dihydrorhodamine 123 (DHR 123) and dichlorofluorescein diacetate (DCFH-DA) as fluorochromes were different. The reactive oxygen species content of cells in 8 × 10−7 M H2O2 was similar to that of the controls, but there was a sudden, marked production of superoxide anions (detected with DHR 123) and H2O2 or peroxides (detected with DCFH-DA) by cells incubated with 1 × 10−5 M H2O2 which increased with increasing H2O2 until cell death.  相似文献   

10.
In the absence of added Fe2+, the ATPase activity of isolatedSchizosaccharomyces pombe plasma membranes (5–7 μmolP i per mg protein per min) is moderately inhibited by H2O2 in a concentration-dependent manner. Sizable inactivation occurs only at 50–80 mmol/L H2O2. The process, probably a direct oxidative action of H2O2 on the enzyme, is not induced by the indigenous membrane-bound iron (19.3 nmol/mg membrane protein), is not affected by the radical scavengers mannitol and Tris, and involves a decrease of both theK m of the enzyme for ATP and theV of ATP splitting. On exposing the membranes to the Fenton reagent (50 μmol/L Fe2+ +20 mmol/L H2O2), which causes a fast production of HO radicals, the ATPase is 50–60% inactivated and 90% of added Fe2+ is oxidized to Fe3+ within 1 min. The inactivation occurs only when Fe2+ is added before H2O2 and can thus bind to the membranes. The lack of effect of radical scavengers (mannitol, Tris) indicates that HO radicals produced in the bulk phase play no role in inactivation. Blockage of the inactivation by the iron chelator deferrioxamine implies that the process requires the presence of Fe2+ ions bound to binding sites on the enzyme molecules. Added catalase, which competes with Fe2+ for H2O2, slows down the inactivation but in some cases increases its total extent, probably due to the formation of the superoxide radical that gives rise to delayed HO production.  相似文献   

11.
[Cu2+•Cys-Gly-His-Lys] stimulates thermolysin (TLN) activity at low concentration (below 10 μM) and inhibits the enzyme at higher concentration, with binding affinities of 2.0 and 4.9 μM, respectively. The metal-free Cys-Gly-His-Lys peptide also stimulates TLN activity, with an apparent binding affinity of 2.2 μM. Coordination of copper through deprotonated imine nitrogens, the histidyl nitrogen, and the free N-terminal amino group is consistent with the characteristic absorption spectrum of a Cu2+–amino-terminal copper and nickel binding motif (λ max ∼ 525 nm). The lack of thiol coordination is suggested by both the absence of a thiol to Cu2+ charge transfer band and electrochemical studies, since the electrode potential (vs. Ag/AgCl) 0.84 V (ΔE = 92 mV) for the Cu3+/2+ redox couple obtained for [Cu2+•Cys-Gly-His-Lys] was found to be in close agreement with that of a related complex [Cu2+•Lys-Gly-His-Lys]+ (0.84 V, ΔE = 114 mV). The N-terminal cysteine appears to be available as a zinc-anchoring residue and plays a critical functional role since the [Cu2+•Lys-Gly-His-Lys]+ homologue exhibits neither stimulation nor inhibition of TLN. Under oxidizing conditions (ascorbate/O2) the catalyst is shown to mediate the complete irreversible inactivation of TLN at concentrations where enzyme activity would otherwise be stimulated. The observed rate constant for inactivation of TLN activity was determined as k obs = 7.7 × 10−2 min−1, yielding a second-order rate constant of (7.7 ± 0.9) × 104 M−1 min−1. Copper peptide mediated generation of reactive oxygen species that subsequently modify active-site residues is the most likely pathway for inactivation of TLN rather than cleavage of the peptide backbone.  相似文献   

12.
The parameters of a dense (1013–1014 cm−3) plasma produced by ionization of a H2 + Ti mixture in a moderate-power (W ≤ 10 MW) pulsed reflective discharge are investigated. The dynamics of the plasma density, the elemental composition of the generated plasma, the radial distribution of the electron density, the rotation velocity and frequency of the plasma layer with n p n cr, the radial electric field, the coefficient of plasma particle separation, and the coefficient of plasma recombination in the stage of plasma decay are determined.  相似文献   

13.
An iron-hexacyanide-covered microelectrode sensor has been used to continuously monitor the kinetics of hydrogen peroxide decomposition catalyzed by oxidized cytochrome oxidase. At cytochrome oxidase concentration ≈1 μM, the catalase activity behaves as a first order process with respect to peroxide at concentrations up to ≈300–400 μM and is fully blocked by heat inactivation of the enzyme. The catalase (or, rather, pseudocatalase) activity of bovine cytochrome oxi- dase is characterized by a second order rate constant of ≈2•102 M-1•sec-1 at pH 7.0 and room temperature, which, when divided by the number of H2O2 molecules disappearing in one catalytic turnover (between 2 and 3), agrees reasonably well with the second order rate constant for H2O2-dependent conversion of the oxidase intermediate FI-607 to FII-580. Accordingly, the catalase activity of bovine oxidase may be explained by H2O2 procession in the oxygen-reducing center of the enzyme yielding superoxide radicals. Much higher specific rates of H2O2 decomposition are observed with preparations of the bacterial cytochrome c oxidase from Rhodobacter sphaeroides. The observed second order rate constants (up to ≈3000 M-1•sec-1) exceed the rate constant of peroxide binding with the oxygen-reducing center of the oxidized enzyme (≈500 M-1•sec-1) several-fold and therefore cannot be explained by catalytic reaction in the a 3/CuB site of the enzyme. It is proposed that in the bacterial oxidase, H2O2 can be decomposed by reacting with the adventitious transition metal ions bound by the polyhistidine-tag present in the enzyme, or by virtue of reaction with the tightly-bound Mn2+, which in the bacterial enzyme substitutes for Mg2+ present in the mitochondrial oxidase.  相似文献   

14.
DNA-binding proteins from nutrient-starved cells (DPS) protect cells from oxidative stress by removing H2O2 and iron. A new class of DPS-like proteins has recently been identified, with DPS-like protein from Sulfolobus solfataricus (SsDPS) being the best characterized to date. SsDPS protects cells from oxidative stress and is upregulated in response to H2O2 but also in response to iron depletion. The ferroxidase active site of SsDPS is structurally similar to the active sites of manganese catalase and rat liver arginase. The present work shows that the ferroxidase center in SsDPS binds two Mn2+ ions with K D = (1/K 1 K 2)1/2 = 48(3) μM. The binding constant of the second Mn2+ is significantly higher than that of the first, inducing dinuclear Mn(II) cluster formation for all but the lowest concentrations of added Mn2+. In competition experiments, equimolar amounts of Fe2+ were unable to displace the bound manganese. EPR spectroscopy of the Mn2 2+ cluster showed signals comparable to those of other characterized dimanganese clusters. The exchange coupling for the cluster was determined, J = −1.4(3) cm−1 (H = −2JS 1 S 2), and is within the range expected for a μ1,1-carboxylato bridge between the manganese ions. Manganese-bound SsDPS showed catalase activity at a rate 10–100 times slower than for manganese catalases. EPR spectra of SsDPS after addition of H2O2 showed the appearance of an intermediate in the reaction with H2O2.  相似文献   

15.
The present in vitro study was designed to examine the antioxidative activity of red cabbage anthocyanins (ATH) in the protection of blood plasma proteins and lipids against damage induced by oxidative stress. Fresh leaves of red cabbage were extracted with a mixture of methanol/distilled water/0.01% HCl (MeOH/H2O/HCl, 50/50/1, v/v/w). Total ATH concentration [μM] was determined with cyanidin 3-glucoside as a standard. Phenolic profiles in the crude red cabbage extract were determined using the HPLC method. Plasma samples were exposed to 100 μM peroxynitrite (ONOO) or 2 mM hydrogen peroxide (H2O2) in the presence/absence of ATH extract (5–15 μM); oxidative alterations were then assessed. Pre-incubation of plasma with ATH extract partly reduced oxidative stress in plasma proteins and lipids. Dose-dependent reduction of both ONOO and H2O2-mediated plasma protein carbonylation was observed. ATH extract partly inhibited the nitrative action of ONOO, and significantly decreased plasma lipid peroxidation caused by ONOO or H2O2. Our results demonstrate that anthocyanins present in red cabbage have inhibitory effects on ONOO and H2O2-induced oxidative stress in blood plasma components. We suggest that red cabbage ATH, as dietary antioxidants, should be considered as potentially usable nutraceuticals in the prevention of oxidative stress-related diseases.  相似文献   

16.
Cyanobacterial contamination of water has been a serious problem in recent years. Thus, the effective control of undesired cyanobacteria has become an urgent issue. We studied therefore the effects of ρ-coumaric acid and vanillic acid on toxic Microcystis aeruginosa and the allelopathic mechanisms. The results showed that the growth of toxic M. aeruginosa was significantly inhibited by ρ-coumaric acid and vanillic acid, with an EC50 of 0.26 ± 0.07 and 0.34 ± 0.05 mmol L−1, respectively. Our data also demonstrated that both ρ-coumaric acid and vanillic acid triggered the generation of superoxide anion radicals (O2 •−). The O2 •− might induce a lipid peroxidation which may change cell membrane penetrability, thereby leading to the eventual death of M. aeruginosa. Our current studies further provide evidence that some phenolic acids such as ρ-coumaric acid and vanillic acid may be a potential effective solution for aquatic management.  相似文献   

17.
Summary Protoplasts were isolated from oat (Avena sativa L.) leaves by the combination of highly purified preparations of pectin lyase, xylanase, and cellulase C1. During the enzymic isolation, superoxide radical (O 2 ) was generated from the tissues. Both the protoplasts themselves and the cell walls, exposed to enzyme treatment, produced O 2 . Hydrogen peroxide (H2O2) apparently accumulated in the reaction mixture due to the spontaneous dismutation reaction of O 2 , while a part of H2O2 may have been produced directly from cell walls by the action of enzymes. Singlet molecular oxygen (1O2) generated in the reaction mixture was detected by cholesterol oxidation in small unilamellar liposomes. It seems likely that1O2 may be generated by the peroxidase-H2O2-halide system during enzymic treatment of the leaves. The work was partially supported by the Research Project “Research and development of the improvement of bacterial and plant cells by cell fusion” of the Food and Agriculture Research and Development Association (Japan).  相似文献   

18.
The activity of 1-aminocyclopropane-1-carboxylic acid synthase (ACC synthase, ACS) and the concentrations of superoxide radical (O2−.) and hydrogen peroxide (H2O2) were measured in etiolated mungbean seedlings following their transfer to a growth chamber at 25°C after a 5-h-chilling treatment at 5°C. All of these variables increased dramatically after the transfer, and strong correlations were found between ACS activity and the concentrations of superoxide and H2O2. Exogenous applications of two generators of superoxide radicals, methylviologen (MV) and xanthine–xanthine oxidase (X–XOD), enhanced ACS activity in seedlings, but their effects were inhibited by exogenous applications of specific scavengers of O2−.. However, applications of H2O2 or specific H2O2-scavengers had no significant effects on seedlings ACS activity. The results indicate that O2−. was involved in the chilling-induced increases in ACS activity, but not H2O2. ACS activity peaked ca. 8 h after the transfer, and then declined, but the decline could be counteracted by exogenous applications of specific O2−. scavengers, this suggests that damage was caused by superoxide radicals influencing ACS activity in etiolated mungbean seedlings. Further analysis of changes in two key kinetic parameters of ACS activity—V max (maximum velocity) and K m (the Michaelis constant)—in the seedlings indicated that the presence of O2−. may reduce K m, i.e. increase substrate (S-adenosyl methionine, SAM) affinity. That would be the main mechanism responsible for the observed chilling-induced increases in ACS activity in etiolated mungbean seedlings.  相似文献   

19.
Chitosan, CN, or H2O2 caused the death of epidermal cells (EC) in the epidermis of pea leaves that was detected by monitoring the destruction of cell nuclei; chitosan induced chromatin condensation and marginalization followed by the destruction of EC nuclei and subsequent internucleosomal DNA fragmentation. Chitosan did not affect stoma guard cells (GC). Anaerobic conditions prevented the chitosan-induced destruction of EC nuclei. The antioxidants nitroblue tetrazolium or mannitol suppressed the effects of chitosan, H2O2, or chitosan + H2O2 on EC. H2O2 formation in EC and GC mitochondria that was determined from 2′,7′-dichlorofluorescein fluorescence was inhibited by CN and the protonophoric uncoupler carbonyl cyanide m-chlorophenylhydrazone but was stimulated by these agents in GC chloroplasts. The alternative oxidase inhibitors propyl gallate and salicylhydroxamate prevented chitosan- but not CN-induced destruction of EC nuclei; the plasma membrane NADPH oxidase inhibitors diphenylene iodonium and quinacrine abolished chitosan- but not CN-induced destruction of EC nuclei. The mitochondrial protein synthesis inhibitor lincomycin removed the destructive effect of chitosan or H2O2 on EC nuclei. The effect of cycloheximide, an inhibitor of protein synthesis in the cytoplasm, was insignificant; however, it was enhanced if cycloheximide was added in combination with lincomycin. The autophagy inhibitor 3-methyladenine removed the chitosan effect but exerted no influence on the effect of H2O2 as an inducer of EC death. The internucleosome DNA fragmentation in conjunction with the data on the 3-methyladenine effect provides evidence that chitosan induces programmed cell death that follows a combined scenario including apoptosis and autophagy. Based on the results of an inhibitor assay, chitosan-induced EC death involves reactive oxygen species generated by the NADPH oxidase of the plasma membrane.  相似文献   

20.
In this study, we have characterized the cellular source and mechanism for the enhanced generation of reactive oxygen species (ROS) in the myocardium during Trypanosoma cruzi infection. Cardiac mitochondria of infected mice, as compared to normal controls, exhibited 63.3% and 30.8% increase in ROS-specific fluorescence of dihydroethidium (detects O2 •−) and amplex red (detects H2O2), respectively. This increase in ROS level in cardiac mitochondria of infected mice was associated with a 59% and 114% increase in the rate of glutamate/malate- (complex I substrates) and succinate- (complex II substrate) supported ROS release, respectively, and up to a 74.9% increase in the rate of electron leakage from the respiratory chain when compared to normal controls. Inhibition studies with normal cardiac mitochondria showed that rotenone induced ROS generation at the QNf-ubisemiquinone site in complex I. In complex III, myxothiazol induced ROS generation from a site located at the Qo center that was different from the Qi center of O2 •− generation by antimycin. In cardiac mitochondria of infected mice, the rate of electron leakage at complex I during forward (complex I-to-complex III) and reverse (complex II-to-complex I) electron flow was not enhanced, and complex I was not the main site of increased ROS production in infected myocardium. Instead, defects of complex III proximal to the Qo site resulted in enhanced electron leakage and ROS formation in cardiac mitochondria of infected mice. Treatment of infected mice with phenyl-α-tert-butyl-nitrone (PBN) improved the respiratory chain function, and, subsequently, decreased the extent of electron leakage and ROS release. In conclusion, we show that impairment of the Qo site of complex III resulted in increased electron leakage and O2 •− formation in infected myocardium, and was controlled by PBN.  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