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.
When cultured on a defined diet, ethanol was an efficient substrate for lipid synthesis in wild-type Drosophila melanogaster larvae. At certain dietary levels both ethanol and sucrose could displace the other as a lipid substrate. In wild-type larvae more than 90% of the flux from ethanol to lipid was metabolized via the alcohol dehydrogenase (ADH) system. The ADH and aldehyde dehydrogenase activities of ADH were modulated in tandem by dietary ethanol, suggesting that ADH provided substrate for lipogenesis by degrading ethanol to acetaldehyde and then to acetic acid. The tissue activity of catalase was suppressed by dietary ethanol, implying that catalase was not a major factor in ethanol metabolism in larvae. The activities of lipogenic enzymes, sn-glycerol-3-phosphate dehydrogenase, fatty acid synthetase (FAS), and ADH, together with the triacylglycerol (TG) content of wild-type larvae increased in proportion to the dietary ethanol concentration to 4.5% (v/v). Dietary ethanol inhibited FAS and repressed the accumulation of TG in ADH-deficient larvae, suggesting that the levels of these factors may be subject to a complex feedback control.This research was supported by National Institutes of Health Grant GM-28779 to B.W.G. and a Monash University Research Grant to S.W.M.  相似文献   

2.
The methanol metabolism in acatalasemic mice was studied by administering [14C]methanol and [14C]formic acid to acatalasemic and normal mice and determining the radioactivity of exhaled carbon dioxide. Methanol metabolism was also studied in acatalasemic and normal mice treated with 3-amino-1,2,4-triazole (AT), which is known to be an inhibitor of catalase (EC 1.11.1.6). The metabolism of methanol and formic acid was inhibited in acatalasemic mice as seen by reduced [14C]CO2 production. Similar results were obtained when AT was given prior to the methanol injection into the normal and acatalasemic mice. The results indicate the peroxidative activity of catalase plays the major role in the methanol metabolism in mice. On the other hand similar studies with [1-14C] ethanol showed that the metabolism of ethanol was not inhibited in acatalasemic mice.  相似文献   

3.
Oxidation of methanol, ethanol, propanol, and butanol by the microsomal fraction of rat liver homogenate is described. This microsomal alcohol-oxidizing system is dependent on NADPH and molecular oxygen and is partially inhibited by CO, features which are common for microsomal drug-metabolizing enzymes. The activity of the microsomal alcohol-oxidizing system could be dissociated from the alcohol peroxidation via catalase-H2O2 by differences in substrate specificity, since higher aliphatic alcohols react only with the microsomal system, but not with catalase-H2O2. Following solubilization of microsomes by ultrasonication and treatment with deoxycholate, the activity of the microsomal alcohol-oxidizing system was separated from contaminating catalase by DEAE-cellulose column chromatography, ruling out an obligatory involvement of catalase-H2O2 in the activity of the NADPH-dependent microsomal alcohol-oxidizing system. In intact hepatic microsomes, the catalase inhibitor sodium azide slightly decreased the oxidation of methanol and ethanol, but not that of propanol and butanol, indicating a facultative role of contaminating catalase in the microsomal oxidation of lower aliphatic alcohols only. It is suggested that the microsomal alcohol-oxidizing system accounts, at least in part, for that fraction of hepatic alcohol metabolism which is independent of the pathway involving alcohol dehydrogenase activity.  相似文献   

4.
Methanol is one of the most common short-chain alcohols in fermenting fruits, the natural food of the fruit fly, Drosophila melanogaster. The larvae cope continuously with methanol at various concentrations in order to survive and develop. In the present article, we found toxicities of dietary methanol and formaldehyde were enhanced by piperonyl butoxide, but not by 3-amino-1, 2, 4-triazole, 4-methylpyrazole, diethylmeleate, and triphenyl phosphate, when assessing by the combination index method. These results reveal that cytochrome P450 monooxygenases (CYPs), rather than catalases, alcohol dehydrogenases, glutathione S-transferases, and esterases, participate in methanol metabolism. Moreover, methanol exposure dramatically increased CYP activity. The ratios of the CYP activities in treated larvae to those in control reached, respectively, up to 3.0-, 3.9-, and 2.7-fold, at methanol concentrations of 22.6, 27.9, and 34.5 mg/g diet. In addition, methanol exposure greatly up-regulated the mRNA expression level of five Cyp genes, which were Cyp304a1, Cyp9f2, Cyp28a5, Cyp4d2, and Cyp4e2. Their resulting proteins were suggested as the candidate enzymes for methanol metabolism in D. melanogaster larvae.  相似文献   

5.
Alcohol metabolism via alcohol dehydrogenase (ADH) and catalase was studied in perfused rat livers by measuring the oxidation of methanol and butanol, selective substrates for catalase and ADH, respectively. In livers from fasted rats, basal rates of methanol uptake of 15 +/- 1 mumol/g/h were decreased significantly to 8 +/- 2 mumol/g/h by addition of butanol. Concomitantly, pyridine nucleotide fluorescence detected from the liver surface was increased by butanol but not methanol. Both effects of butanol were blocked by an inhibitor of ADH, 4-methylpyrazole, consistent with the hypothesis that elevation of the NADH redox state by butanol inhibited H2O2 production via NAD+-requiring peroxisomal beta-oxidation, leading indirectly to diminished rates of catalase-dependent methanol uptake. In support of this idea, both butanol and butyraldehyde inhibited H2O2 generation. The NADH redox state was also elevated by xylitol, causing a 75% decrease in rates of methanol uptake by livers from fasted rats. This effect was not observed in livers from fed rats unless malate-aspartate shuttle activity was reduced by infusion of the transaminase inhibitor aminooxyacetate. Taken together, these data indicate that generation of reducing equivalents from ADH in the cytosol inhibits H2O2 generation leading to significantly diminished rates of peroxidation of alcohols via catalase. This phenomenon may represent an important physiological mechanism of regulation of ethanol oxidation in intact cells.  相似文献   

6.
NADP-malic enzyme (NADP-ME) (E.C. 1.1.1.40) is situated in the cytosol of Drosophila melanogaster. Both the tissue activity and CRM level of NADP-ME parallel changes in the dosage of a gene, Men+, located in region 87C2-3 to 87D1-2 of the third chromosome. The tissue activity of NADP-ME is very high in early third instar larvae, providing about 33% of the NADPH at this life stage. The tissue activity declines during pupal development but increases as the adult ages. The concentration of NADP-ME CRM and tissue activity are coordinately increased in third instar larvae by dietary carbohydrate and decreased by dietary lipid.  相似文献   

7.
W. J. Mackay  G. C. Bewley 《Genetics》1989,122(3):643-652
Activated oxygen species have been demonstrated to be the important agents in oxygen toxicity by disrupting the structural and functional integrity of cells through lipid peroxidation events, DNA damage and protein inactivation. The biological consequences of free radical damage have long been hypothesized to be a causal agent in many aging-related diseases. Catalase (H2O2:H2O2 oxidoreductase; EC 1.15.1.1) is one of several enzymes involved in the scavenging of oxygen free radicals and free radical derivatives. The structural gene for catalase in Drosophila melanogaster has been localized to region 75D1-76A on chromosome 3L by dosage responses to segmental aneuploidy. This study reports the isolation of a stable deficiency, Df(3L)CatDH104(75C1-2;75F1), that uncovers the catalase locus and the subsequent isolation of six acatalasemic mutants. All catalase mutants are viable under standard culture conditions and recessive lethal mutations within the 75Cl-F1 interval have been shown not to affect catalase activity. Two catalase mutations are amorphic while four are hypomorphic alleles of the Cat+ locus. The lack of intergenic complementation between the six catalase mutations strongly suggests that there is only one functional gene in Drosophila. One acatalesemic mutation was mapped to position 3-47.0 which resides within the catalase dosage sensitive region. While complete loss of catalase activity confers a severe viability effect, residual levels are sufficient to restore viability to wild type levels. These results suggest a threshold effect for viability and offer an explanation for the general lack of phenotypic effects associated with the known mammalian acatalasemics.  相似文献   

8.
The localization of peroxidase activity in methanol-grown cells of the yeast Hansenula polymorphia has been studied by a method based on cytochemical staining with diaminobenzidine (DAB). The oxidation product of DAB occurred in microbodies, which characteristically develop growth on or methanol, and in the intracristate space of the mitochondria. The staining of microbodies was H2O2 dependent, appeared to be optimal at pH 10.5, diminished below pH 10 and was inhibited by 20 mM 3-amino 1,2,4 triazole (AT). In contrast to these observations, the reaction in the mitochondria was not H2O2 dependent and not notably affected by differences in pH in the range of 8.5 to 10.5. Microbodies and mitochondria were also stained when H2O2 was replaced by methanol. Appropriate control experiments indicated that in this case methanol oxidase generated the H2O2 for the peroxidative conversion of DAB by catalase. These results suggest that catalase is located in the microbodies of methanol-grown yeasts. A model for a possible physiological function of the microbodies during growth on methanol is put forward.  相似文献   

9.
Thiourea and diethylthiourea, two compounds which react with hydroxyl radicals, inhibited NADPH-dependent microsomal oxidation of ethanol and 1-butanol. Inhibition by both compounds was more effective in the presence of the catalase inhibitor, azide. Inhibition by thiourea was noncompetitive with respect to ethanol in the absence of azide but was competitive in the presence of azide. Urea, a compound which does not react with hydroxyl radicals or H2O2, was without effect. Thiourea had no effect on NADH- and NADH-cytochrome c reductase, NADPH oxidase, and NADH- and NADPH-dependent oxygen uptake. Thiourea inhibited the activities of aniline hydroxylase and aminopyrine demethylase. Thiourea, but no other hydroxyl radical scavengers, e.g., dimethyl sulfoxide, mannitol, and benzoate, reacted directly with H202 and decreased H2O2 accumulation in the presence of azide. Therefore the actions of thiourea are complex because it can react with both hydroxyl radicals and H2O2. Differences between the actions of thiourea and those previously reported for dimethyl sulfoxide, mannitol, and benzoate, e.g., effects on drug metabolism, effectiveness of inhibition in the absence of azide, or kinetics of the inhibition, probably reflect the fact that thiourea reacts directly with H2O2 whereas the other agents do not. The current results remain consistent with the concept that microsomal oxidation of alcohols involves interactions of the alcohols with hydroxyl radicals generated from microsomal electron transfer.  相似文献   

10.
11.
Olfactory responses of Drosophila melanogaster larvae to a homologous series of primary alcohols (methanol ... decanol) were tested. Alcohols at either extreme of the chain lengths studied (methanol, ethanol and decanol) evoked no significant responses. Heptanol and nonanol both produced dose-independent responses, larvae being attracted to heptanol and repulsed by nonanol. The remaining alcohols elicited dose-related attractive responses. Responses to hexanol and nonanol decline with increasing larval age. Genetic differences were found for the response to heptanol, with larvae from a Japanese strain, Katsunuma, being indifferent to this substance. Chromosome exchange revealed that a major factor involved in the response to heptanol is located on chromosome II; factors on chromosome III quantitatively modulate this response. Three mutant strains were isolated following EMS mutagenesis of chromosome III. These three strains, IndifferentA, IndifferentB and IndifferentC, show incomplete or total anosmia when stimulated with nonanol. Adult flies from these strains show similar effects. IndifferenB and C strains are dominant over the Canton-S control strain; the IndifferentA strain shows semi-dominance. Results are discussed in the light of the ecology of Drosophila larvae and the relation between olfactory stimulus and receptor conformation and number.  相似文献   

12.
Nitric oxide involvement in Drosophila immunity.   总被引:5,自引:0,他引:5  
A J Nappi  E Vass  F Frey  Y Carton 《Nitric oxide》2000,4(4):423-430
The augmented production of nitric oxide (NO) was observed during the hemocyte-mediated melanotic encapsulation responses of Drosophila melanogaster and D. teissieri. When introduced into the hemocoel of D. melanogaster larvae, NO activated the gene encoding the antimicrobial peptide Diptericin. These observations, together with previous studies documenting the production of superoxide anion (O(*-)(2)) and H(2)O(2) in immune-challenged Drosophila, provide evidence that reactive intermediates of both oxygen (ROI) and nitrogen (RNI) constitute a part of the cytotoxic arsenal employed by Drosophila in defense against both microbial pathogens and eukaryotic parasites. These ROI and RNI appear to represent an evolutionarily conserved innate immune response that is mediated by regulatory proteins that are homologous to those of mammalian species.  相似文献   

13.
Purified Drosophila lebanonensis alcohol dehydrogenase (Adh) revealed one enzymically active zone in starch gel electrophoresis at pH 8.5. This zone was located on the cathode side of the origin. Incubation of D. lebanonensis Adh with NAD+ and acetone altered the electrophoretic pattern to more anodal migrating zones. D. lebanonensis Adh has an Mr of 56,000, a subunit of Mr of 28 000 and is a dimer with two active sites per enzyme molecule. This agrees with a polypeptide chain of 247 residues. Metal analysis by plasma emission spectroscopy indicated that this insect alcohol dehydrogenase is not a metalloenzyme. In studies of the substrate specificity and stereospecificity, D. lebanonensis Adh was more active with secondary than with primary alcohols. Both alkyl groups in the secondary alcohols interacted hydrophobically with the alcohol binding region of the active site. The catalytic centre activity for propan-2-ol was 7.4 s-1 and the maximum velocity of most secondary alcohols was approximately the same and indicative of rate-limiting enzyme-coenzyme dissociation. For primary alcohols the maximum velocity varied and was much lower than for secondary alcohols. The catalytic centre activity for ethanol was 2.4 s-1. With [2H6]ethanol a primary kinetic 2H isotope effect of 2.8 indicated that the interconversion of the ternary complexes was rate-limiting. Pyrazole was an ethanol-competitive inhibitor of the enzyme. The difference spectra of the enzyme-NAD+-pyrazole complex gave an absorption peak at 305 nm with epsilon 305 14.5 X 10(3) M-1 X cm-1. Concentrations and amounts of active enzyme can thus be determined. A kinetic rate assay to determine the concentration of enzyme active sites is also presented. This has been developed from active site concentrations established by titration at 305 nm of the enzyme and pyrazole with NAD+. In contrast with the amino acid composition, which indicated that D. lebanonensis Adh and the D. melanogaster alleloenzymes were not closely related, the enzymological studies showed that their active sites were similar although differing markedly from those of zinc alcohol dehydrogenases.  相似文献   

14.
Extracts prepared from heads of Drosophila melanogaster show high-affinity binding (KD = 1.9 nM) of [3H]saxitonin, a compound known to bind to and block voltage-sensitive sodium channels in other organisms. The interaction between saxitoxin and the Drosophila saxitoxin receptor is non-cooperative and reversible with a half-life of 18.3 s for binding at 4 degrees C. The saturable binding is specifically inhibited by tetrodotoxin with a K1 = 0.30 nM. The number of saturable binding sites in the extract is 97 fmol/mg protein. Since approx. 50% of the binding activity is recovered in the extract, the number of binding sites in the head is estimated to be 6.4 fmol/mg head. Nerve conduction in Drosophila larvae is completely blocked after 20 min in a bathing solution containing 200 nM tetrodotoxin. A comparison between the binding and the electrophysiological studies in Drosophila and other organisms suggests that the Drosophila saxitoxin receptor is part of the voltage-sensitive sodium channel involved in the propagation of action potentials. A mutant (ttxs), which is abnormally sensitive to dietary tetrodotoxin, is shown to be indistinguishable from wild type with respect to [3H]saxitonin-binding properties and physiological sensitivity to tetrodotoxin. These studies provide techniques which can be used to identify mutants with defects in the saxitoxin-binding component of the sodium channel.  相似文献   

15.
In order to study the mutagenic effect of exposure to tritium, Drosophila melanogaster larvae were treated with tritiated water (3H2O) or tritiated thymidine (3H-TdR) during development. Dose rates ranged from 0.0058 to 0.058 rad/h per nucleus for 3H-TdR and from 0.049 to 0.122 rad/h for 3H2O. Induction of mutations was measured by the appearance of somatic mutations in the eyes of an unstable strain of Drosophila melanogaster. Both substances caused a significant increase in mutation frequency. With the assumption that each mutation observed in this assay is caused by one DNA break, the effectiveness of tritium to create DNA breaks is estimated to be 0.20 breaks per decay for 3H-TdR and 0.27 breaks per decay for 3H2O.  相似文献   

16.
Exposure of early third instar larvae of Drosophila melanogaster to a nonlethal dose of ethanol was detrimental to larvae lacking alcohol dehydrogenase (ADH) but beneficial to wild-type larvae in terms of surviving a later ethanol tolerance test, indicating that one of the important functions of the ADH system is to supply derivatives of ethanol to larvae that in turn promote ethanol tolerance. High intracellular concentrations of ethanol in ADH-deficient (Adhn2) larvae fed ethanol were accompanied by a decrease in the cell membrane infoldings of fat body cells, suggesting that the capacities to absorb and release molecules were reduced. Marked effects of ethanol on the endoplasmic reticulum and mitochondria of ADH-deficient larvae were also evident. The absence of similar changes in wild-type larvae that were fed moderate levels of ethanol showed that the ADH system kept the intracellular level of ethanol at a concentration low enough to avoid cell damage. A cytometric analysis of electron micrographs showed that there were ethanol-induced reductions in glycogen, lipid, and protein stores in the fat body cells of ADH-deficient larvae fed 1.25% ethanol (v/v) compared with null larvae fed an ethanol-free diet. This finding implied that the capacities to synthesize or store these compounds may be limited by high intracellular concentrations of ethanol. The cytometric analysis also revealed that the consumption of diets containing 2.5% and 4.5% ethanol by Canton-S wild-type larvae for 3 days after 4 days of feeding on an ethanol-free diet resulted in decreases in glycogen and protein deposits in fat body cells, but increased the amount of lipid deposits compared to larvae fed an ethanol-free diet. This observation, coupled with the greater weight of wild-type adults that were fed a growth-limiting concentration of ethanol compared with control adults, suggested that a metabolic defense mechanism in larvae is to convert toxic ethanol to nontoxic storage products. Dietary ethanol alone and in combination with isopropanol stimulated an increase in the size of the NAD-pool in larvae, a condition that may favor the activity of ADH. A low dietary level of isopropanol (1%) completely blocked glycogen deposition in wild-type larvae, whereas ethanol did not. Thus ethanol and isopropanol exert some different toxic effects on larval fat bodies.  相似文献   

17.
The organic hydroperoxide cumene hydroperoxide is capable of oxidizing ethanol to acetaldehyde in the presence of either catalase, purified cytochrome P-450 or rat liver microsomes. Other hemoproteins like horseradish peroxidase, cytochrome c or hemoglobin were ineffective. In addition to ethanol, higher alcohols like 1-propanol, 1-butanol and 1-pentanol are also oxidized to their corresponding aldehydes to a lesser extent. Other organic hydroxyperoxides will replace cumene hydroperoxide in oxidizing ethanol but less effectively. The cumene-hydroperoxide-dependent ethanol oxidation in microsomes was inhibited partially by cytochrome P-450 inhibitors but was unaffected by catalase inhibitors. Phenobarbital pretreatment of rats increased the specific activity of the cumene-hydroperoxide-dependent ethanol oxidation per mg of microsomes about seven-fold. The evidence suggests that cytochrome P-450 rather than catalase is the enzyme responsible for hydroperoxide-dependent ethanol oxidation. However, when H2O2 is used in place of cumene hydroperoxide, the microsomal ethanol oxidation closely resembles the catalase system.  相似文献   

18.
Here we investigated H2O2 production and detoxification in the hematophagous hemiptera, Rhodnius prolixus. Superoxide dismutase (SOD) catalyzes the dismutation of superoxide radical (O2-). This reaction produces hydrogen peroxide, which is scavenged by antioxidant enzymes such as catalase (CAT). SOD and CAT activities were found in all tissues studied, being highest in the midgut. CAT was dose-dependently inhibited in vivo by injections of 3-amino-1,2,4-triazole (AT). Insects treated with AT showed a twofold increase in H2O2 levels. Injection of DL-buthionine-[S, R]-sulfoximine (BSO), an inhibitor of glutathione synthesis, also resulted in a fourfold increase in H2O2, together with stimulation of CAT activity. Simultaneous administration of both AT and BSO had a synergistic effect on midgut H2O2 content. Taken all together, our results suggest that CAT and glutathione-dependent mechanisms cooperate to control H2O2 concentration in the midgut cell and prevent hydroxyl radical generation by Fenton reaction in this tissue.  相似文献   

19.
Modifications to the alkaline Comet assay by using lesion-specific endonucleases, such as formamidopyrimidine-DNA glycosylase (FPG) and endonuclease III (ENDOIII, also known as Nth), can detect DNA bases with oxidative damage. This modified assay can be used to assess the genotoxic/carcinogenic potential of environmental chemicals. The goal of this study was to validate the ability of this modified assay to detect oxidative stress-induced genotoxicity in Drosophila melanogaster (Oregon R(+)). In this study, we used three well known chemical oxidative stress inducers: hydrogen peroxide (H(2)O(2)), cadmium chloride (CdCl(2)) and copper sulfate (CuSO(4)). Third instar larvae of D. melanogaster were fed various concentrations of the test chemicals (50-200μM) mixed with a standard Drosophila food for 24h. Alkaline Comet assays with and without the FPG and ENDOIII enzymes were performed with midgut cells that were isolated from the control and treated larvae. Our results show a concentration-dependent increase (p<0.05-0.001) in the migration of DNA from the treated larvae. ENDOIII treatment detected more oxidative DNA damage (specifically pyrimidine damage) in the H(2)O(2) exposed larvae compared to FPG or no enzyme treatment (buffer only). In contrast, FPG treatment detected more oxidative DNA damage (specifically purine damage) in CuSO(4) exposed larvae compared to ENDOIII. Although previously reported to be a potent genotoxic agent, CdCl(2) did not induce more oxidative DNA damage than the other test chemicals. Our results show that the modified alkaline Comet assay can be used to detect oxidative stress-induced DNA damage in D. melanogaster and thus may be applicable for in vivo genotoxic assessments of environmental chemicals.  相似文献   

20.
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