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 31 毫秒
1.
Perfusion of isolated rat livers with ethanol at a concentration of 2g/l (%o) resulted in a release of glutamate-pyruvate-transaminase (GPT) and sorbitol dehydrogenase (SDH) into the perfusate as markers of toxicity. Inhibition of alcohol dehydrogenase by 4-methylpyrazole or of aldehyde dehydrogenase by cyanamide totally abolished ethanol hepatotoxicity despite of a severalfold increase in acetaldehyde concentration in the perfusate. Addition of superoxide dismutase or catalase clearly suppressed the ethanol-induced release of GPT and SDH, suggesting that 02∼ and H20, are involved in this process. Also. chelation of iron ions by means of desferrioxamine displayed a clear inhibitory action, suggesting the involvement of an iron-catalyzed Haber-WeiB-reaction leading to the formation of OH radicals in the hepatotoxic response to ethanol. Our data suggest that during the metabolism of acetaldehyde primary reactive oxygen species ('02∼, H202) are produced which may interact to yield hydroxyl or OH-like radicals, which possibly represent the hepatotoxic principle of ethanol.  相似文献   

2.
Sequential addition of vanadyl sulfate to a phosphate-buffered solution of H2O2 released oxygen only after the second batch of vanadyl. Ethanol added to such reaction mixtures progressively decreased oxygen release and increased oxygen consumption during oxidation of vanadyl by H2O2. Inclusion of ethanol after any of the three batches of vanadyl resulted in varying amounts of oxygen consumption, a property also shared by other alcohols (methanol, propanol and octanol). On increasing the concentration of ethanol, vanadyl sulfate or H2O2, both oxygen consumption and acetaldehyde formation increased progressively. Formation of acetaldehyde decreased with increase in the ratio of vanadyl:H2O2 above 2:1 and was undetectable with ethanol at 0.1 mM. The reaction mixture which was acidic in the absence of phosphate buffer (pH 7.0), released oxygen immediately after the first addition of vanadyl and also in presence of ethanol soon after initial rapid consumption of oxygen, with no accompanying acetaldehyde formation. The results underscore the importance of some vanadium complexes formed during vanadyl oxidation in the accompanying oxygen-transfer reactions.  相似文献   

3.
Pyrococcus furiosus is a hyperthermophilic archaeon that grows optimally at 100°C by the fermentation of peptides and carbohydrates to produce acetate, CO2, and H2, together with minor amounts of ethanol. The organism also generates H2S in the presence of elemental sulfur (S0). Cell extracts contained NADP-dependent alcohol dehydrogenase activity (0.2 to 0.5 U/mg) with ethanol as the substrate, the specific activity of which was comparable in cells grown with and without S0. The enzyme was purified by multistep column chromatography. It has a subunit molecular weight of 48,000 ± 1,000, appears to be a homohexamer, and contains iron (~1.0 g-atom/subunit) and zinc (~1.0 g-atom/subunit) as determined by chemical analysis and plasma emission spectroscopy. Neither other metals nor acid-labile sulfur was detected. Analysis using electron paramagnetic resonance spectroscopy indicated that the iron was present as low-spin Fe(II). The enzyme is oxygen sensitive and has a half-life in air of about 1 h at 23°C. It is stable under anaerobic conditions even at high temperature, with half-lives at 85 and 95°C of 160 and 7 h, respectively. The optimum pH for ethanol oxidation was between 9.4 and 10.2 (at 80°C), and the apparent Kms (at 80°C) for ethanol, acetaldehyde, NADP, and NAD were 29.4, 0.17, 0.071, and 20 mM, respectively. P. furiosus alcohol dehydrogenase utilizes a range of alcohols and aldehydes, including ethanol, 2-phenylethanol, tryptophol, 1,3-propanediol, acetaldehyde, phenylacetaldehyde, and methyl glyoxal. Kinetic analyses indicated a marked preference for catalyzing aldehyde reduction with NADPH as the electron donor. Accordingly, the proposed physiological role of this unusual alcohol dehydrogenase is in the production of alcohols. This reaction simultaneously disposes of excess reducing equivalents and removes toxic aldehydes, both of which are products of fermentation.  相似文献   

4.
After a general introduction, the main pathways of ethanol metabolism (alcohol dehydrogenase, catalase, coupling of catalase with NADPH oxidase and microsomal ethanol-oxidizing system) are shortly reviewed. The cytochrome P450 isoform (CYP2E1) specifically involved in ethanol oxidation is discussed. The acetaldehyde metabolism and the shift of the NAD/NADH ratio in the cellular environment (reductive stress) are stressed. The toxic effects of acetaldehyde are mentioned. The ethanol-induced oxidative stress: the increased MDA formation by incubated liver preparations, the absorption of conjugated dienes in mitochondrial and microsomal lipids and the decrease in the most unsaturated fatty acids in liver cell membranes are discussed. The formation of carbon-centered (1-hydroxyethyl) and oxygen-centered (hydroxyl) radicals during the metabolism of ethanol is considered: the generation of hydroxyethyl radicals, which occurs likely during the process of univalent reduction of dioxygen, is highlighted and is carried out by ferric cytochrome P450 oxy-complex (P450–Fe3+O2·−) formed during the reduction of heme-oxygen. The ethanol-induced lipid peroxidation has been evaluated, and it has been shown that plasma F2-isoprostanes are increased in ethanol toxicity.  相似文献   

5.
The effect of pargyline on the uptake of acetaldehyde (in the presence of pyrazole) by isolated rat liver cells was studied after incubating the liver cells for 0, 10, 30, 45, and 60 min with 0.40, 1.30, and 2.6 mm pargyline. Without any incubation period, pargyline had no effect on acetaldehyde uptake. With increasing time of incubation, there was a progressive increase in the extent of inhibition of acetaldehyde uptake by pargyline. This suggests the possibility that pargyline is metabolized to the effective inhibitor or the incubation period allows pargyline to reach its site(s) of action. Pargyline was also a more effective inhibitor of the uptake of lower concentrations of acetaldehyde, e.g., 0.167 mm, than of higher concentrations (1.0 mm) of acetaldehyde, especially after short incubation periods or when pyrazole was omitted from the reaction medium. After a 20- to 30-min incubation period, pargyline inhibited the control rate of ethanol oxidation by the liver cells, as well as the accelerated rate of ethanol oxidation found in the presence of pyruvate or an uncoupling agent. Pargyline had no effect on hepatic oxygen consumption. During ethanol oxidation, a time-dependent release of acetaldehyde into the medium was observed. Pyruvate, by increasing the rate of ethanol oxidation, increased the output of acetaldehyde five- to tenfold. Pargyline increased the output of acetaldehyde two- to threefold, despite decreasing the rate of ethanol metabolism by the liver cells. These data indicate that pargyline inhibits the low Km aldehyde dehydrogenase in intact rat liver cells and that this enzyme plays the major role in oxidizing the acetaldehyde which arises during the metabolism of ethanol. Although most of the acetaldehyde generated during the oxidation of ethanol is removed by the liver cells in an effective manner, changes in the activity of aldehyde dehydrogenase or the rate of acetaldehyde generation significantly alter the hepatic output of acetaldehyde.  相似文献   

6.
Previous studies provided indirect evidence that hydroxyl radicals are involved in the oxidation of primary aliphatic alcohols by rat liver microsomes. In the current study, three ·OH scavengers were used as chemical probes to evaluate ·OH production by microsomes. The scavengers and their products were 3-thiomethylpropanal (methional) and 2-keto-4-thiomethylbutyric acid, which yield ethylene gas, and dimethylsulfoxide, which yields methane gas. We observed that microsomes actively generate the appropriate hydrocarbon gas from each scavenger when electron transport is initiated with NADPH. Hydrocarbon gas production is augmented by 0.5 mm azide, an agent which inhibits catalase and, thereby, permits H2O2 to accumulate. However, no metabolism of scavengers occurs when H2O2 is added in the absence of microsomes. These results are consistent with a presumed role for H2O2 as a precursor of hydroxyl radicals. In addition, no metabolism of scavengers occurs when azide and H2O2 are added either to boiled microsomes or to intact microsomes in the absence of electron transport (NADPH-generating system omitted). Therefore, both H2O2 and simultaneous electron transport are required. Ethanol inhibits the metabolism of the scavengers. Similarly, the scavengers inhibit the oxidation of ethanol to acetaldehyde; inhibition in the presence of azide is competitive. These latter results indicate a competition between the scavengers and ethanol for metabolically generated ·OH in microsomes. The specificity of this interaction is evident from the observation that the scavengers do not affect the activities of microsomal aminopyrine demethylase or aniline hydroxylase. Two model ·OH-generating systems (Fenton's reagent and iron-EDTA-ascorbate) were also studied and they produced acetaldehyde from ethanol and hydrocarbon gases from the scavengers. These results, as a whole, tend to verify a role for ·OH in the microsomal oxidation of alcohols.  相似文献   

7.
Rolf A. Løvstad 《Biometals》2003,16(3):435-439
Xanthine oxidase reduces molecular oxygen to H2O2 and superoxide radicals during its catalytic action on xanthine, hypoxanthine or acetaldehyde. Ascorbate is catalytically oxidized by the superoxide radicals generated, when present in the reaction solution (Nishikimi 1975). The present study shows that iron ions markedly stimulate the enzyme dependent ascorbate oxidation, by acting as a red/ox-cycling intermediate between the oxidase and ascorbate. An apparent Km-value of 10.8 M characterized the iron stimulatory effect on the reaction at pH 6.0. Reduced transition-state metals can be oxidized by H2O2 through a Fenton-type reaction. Catalase was found to reduce the effect of iron on the enzyme dependent ascorbate oxidation, strongly suggesting that H2O2, produced during catalysis, is involved in the oxidation of ferrous ions.  相似文献   

8.
Superoxide radicals, a species known to mobilize ferritin iron, and their interaction with catalytic iron have been implicated in the pathogenesis of alcohol-induced liver injury. The mechanism(s) by which ethanol metabolism generates free radicals and mobilizes catalytic iron, however, is not fully defined. In this investigation the role of hepatic aldehyde oxidase in the mobilization of catalytic iron from ferritin was studied in vitro. Iron mobilization due to the metabolism of ethanol to acetaldehyde by alcohol dehydrogenase was increased 100% by the addition of aldehyde oxidase. Iron release was favored by low pH and low oxygen concentration. Mobilization of iron due to acetaldehyde metabolism by aldehyde oxidase was completely inhibited by superoxide dismutase but not by catalase suggesting that superoxide radicals mediate mobilization. Acetaldehyde-aldehyde oxidase mediated reduction of ferritin iron was facilitated by incubation with menadione, an electron acceptor for aldehyde oxidase. Mobilization of ferritin iron due to the metabolism of acetaldehyde by aldehyde oxidase may be a fundamental mechanism of alcohol-induced liver injury.  相似文献   

9.
《Life sciences》1992,51(21):PL195-PL200
We have recently shown that Helicobacter pylori possesses marked alcohol dehydrogenase (ADH) activity and is capable - when incubated with an ethanol containing solution in vitro - of producing large amounts of acetaldehyde. In the present study we report that some drugs commonly used for the eradication of H. pylori and for the treatment of gastroduodenal diseases are potent ADH inhibitors and, consequently, effectively prevent bacterial oxidation of ethanol to acetaldehyde. Colloidal bismuth subcitrate (CBS), already at a concentration of 0.01 mM, inhibited H. pylori ADH by 93% at 0.5 M ethanol and decreased oxidation of 22 mM ethanol to acetaldehyde to 82% of control. At concentrations above 5 mM, CBS almost totally inhibited acetaldehyde formation. Omeprazole, a drug also known to suppress growth of H. pylori, also inhibited H. pylori ADH and suppressed bacterial acetaldehyde formation significantly to 69% of control at a drug concentration of 0.1 mM. By contrast, the H2-receptor antagonists ranitidine and famotidine showed only modest effect on bacterial ADH and acetaldehyde production. We suggest that inhibition of bacterial ADH and a consequent suppression of acetaldehyde production from endogenous or exogenous ethanol may be a novel mechanism by which CBS and omeprazole exert their effect both on the growth of H. pylori as well as on H. pylori associated gastric injury.  相似文献   

10.
The solubilization and subsequent separation of the hepatic microsomal ethanol-oxidizing system from alcohol dehydrogenase and catalase activities by DEAE-cellulose column chromatography is described. Absence of alcohol dehydrogenase in the column eluates exhibiting microsomal ethanol-oxidizing system activity was demonstrated by the failure of NAD+ to promote ethanol oxidation at pH 9.6. Differentiation of the microsomal ethanol-oxidizing system from alcohol dehydrogenase was further shown by the apparent Km for ethanol (7.2 mm, insensitivity of the microsomal ethanol-oxidizing system to the alcohol dehydrogenase inhibitor pyrazole (0.1 mm) and by the failure of added alcohol dehydrogenase to increase the ethanol oxidation. Absence of catalatic activity in these fractions was demonstrated by spectrophotometric and polarographic assay. Differentiation of the microsomal ethanol-oxidizing system from the peroxidatic activity of catalase was shown by the apparent Km for oxygen (8.3 μm), insensitivity of the microsomal ethanol-oxidizing system to the catalase inhibitors azide and cyanide, and by the lack of a H2O2-generating system (glucose-glucose oxidase) to sustain ethanol oxidation in the eluates. The oxidation of ethanol to acetaldehyde by the alcohol dehydrogenase- and catalase-free fractions required NADPH and oxygen and was inhibited by CO. The column eluates showing microsomal ethanol-oxidizing system activity contained cytochrome P-450, NADPH-cytochrome c reductase, and phospholipids and also metabolized aminopyrine, benzphetamine, and aniline.  相似文献   

11.
《Free radical research》2013,47(4-5):195-206
In situ photolysis at 20oC (argon plasma light source, $, $ 200 mm) of oxygen-free solutions containing 2mM H202 and heat-denatured, single-stranded (sS)DNA from calf-thymus resulted in the ESR spectra of the 6-hydroxy-5,6-dihydro-thymin-5-yl {1} and 5-methyleneuracil {3} radicals linked to the sugar-phosphate backbone. They were generated by reaction of OH radicals with DNA. By comparison of the decay characteristics of the ESR signals with rate constants from pulse-conductivity measurements [E. Bothe, G.A. Qureshi and D. Schulte-Frohlinde, Z. Naturforsch. 38c 1030, (1983)] the thymine-derived radicals {1} and {3} can be excluded as precursors of the fast, dominating component of strand breakage of ssDNA. In the absence of H202 from native, doubie-stranded (ds)DNA an ESR signal was obtained (singlet, g ~ 2.004, $1/2 ~ 0.8 mT) which was assigned to the deprotonated guanine radical cation, {G'(-H)} of a DNA subunit. It is assumed that by the UV irradiation the guanine radical cation, {G+}, is generated, either by monophotonic photoionisation or by electron transfer to pyrimidine bases. By rapid transfer of the bridging proton from {G+} to the hydrogen bonded cytosine {G'(-H)} is formed. When photolysis of dsDNA was carried out in the presence of H202, reaction of photolytically generated OH resulted in peroxyl radicals and purine radicals. The oxygen for formation of the peroxyl radicals is probably produced by reaction of {G' (-H)} with H202. Photolysis of N20-saturated solutions containing dsDNA or ssDNA provided another possibility of generation of OH radicals. Under those conditions the OH-induced radicals {1} and {3} were obtained not only from ssDNA but also from dsDNA.  相似文献   

12.
Induction of P450HE1 by pyridine was compared with that by ethanol, and the resulting potentiation of the pneumotoxicity and hepato-toxicity following carbon tetrachloride inhalation by pyridine was examined. Rats were treated with ethanol as either a 10% solution in the drinking water or as a daily bolus (3 ml/kg, ip) dose for 7 days or one bolus dose of pyridine (200 mg/kg, ip) and compared for P450IIE1 apoprotein content by immunoblot analysis. Ethanol in the drinking water and pyridine elevated both hepatic and pulmonary P450IIE1 apoprotein content, but bolus dose ethanol did not. The induction was greatest in the pyridine group. In the interaction study, rats were treated with pyridine (200 mg/kg, ip) and 12 hours later were exposed to CC14 (8000 ppm for 3 hours). Pulmonary injury and hepatic damage were assessed 24 hours later by bronchoalveolar lavage fluid (BALF) analysis [γ-glutamyl transpeptidase (GGT), lactate dehydrogenase (LDH), and total protein] and serum sorbitol dehydrogenase (SDH) activity, respectively. Pyridine alone had no effect on BALF or SDH but enhanced GGT and LDH release into the BALF and SDH release into the serum when compared with CC14 exposure alone. Evaluation of the liver at the light microscopic level revealed characteristic CCl4-induced centrilobular necrosis which was potentiated by pyridine. No changes were observed in the lung by light microscopic evaluation. Pyridine induced pulmonary and hepatic microsomal apoprotein levels of cytochrome P450IIE1 two- and 2- to sixfold, respectively. Exposure to CC14 decreased hepatic but not pulmonary P450IIE1 levels. Induction of cytochrome P450IIE1 by pyridine increases the bioactivation of CC14 in both the liver and lung, leading to enhanced toxicity.  相似文献   

13.
Significant quantitative differences in end-product yields by two strains of Clostridium thermocellum and one strain of Thermoanaerobium brockii were observed during cellobiose fermentation. Most notably, the ethanol/H2 and lactate/acetate ratios were drastically higher for T. brockii as compared with C. thermocellum strains LQRI and AS39. Exogenous H2 addition (0.4 to 1.0 atm) during culture growth increased the ethanol/acetate ratio of both T. brockii and AS39 but had no effect on LQRI. All strains had an operative Embden-Meyerhof glycolytic pathway and displayed catabolic activities of fructose-1,6-diphosphate–activated lactate dehydrogenase, coenzyme A acetylating pyruvate and acetaldehyde dehydrogenase, hydrogenase, ethanol dehydrogenase, and acetate kinase. Enzyme kinetic properties (apparent Km, Vmax, and Q10 values) and the specificity of electron donors/acceptors for different oxidoreductases involved in pyruvate conversion to fermentation products were compared in the three strains. Both species contained ferredoxin-linked pyruvate dehydrogenase and pyridine nucleotide oxidoreductases. Ferredoxin-nicotinamide adenine dinucleotide (NAD) reductase activity was significantly higher in T. brockii than in AS39 and was not detectable in LQRI. H2 production and hydrogenase activity were inversely related to ferredoxin-NAD reductase activity in the three strains. Ferredoxin-NAD phosphate reductase activity was present in cell extracts of both species. Alcohol dehydrogenase activity in C. thermocellum was NAD dependent, unidirectional, and inhibited by low concentrations of NAD and ethanol. Ethanol dehydrogenase activity of T. brockii was both NAD and NADP linked, reversible, and not inhibited by low levels of reaction products. The high lactate yield of T. brockii correlated with increased fructose-1,6-diphosphate. The relation of catabolic enzyme activity and quantitative differences in intracellular electron flow and fermentation product yields of these thermophilic bacteria is discussed.  相似文献   

14.
Methanogenium organophilum, a non-autotrophic methanogen able to use primary and secondary alcohols as hydrogen donors, was grown on ethanol. Per mol of methane formed, 2 mol of ethanol were oxidized to acetate. In crude extract, an NADP+-dependent alcohol dehydrogenase (ADH) with a pH optimum of about 10.0 catalyzed a rapid (5 mol/min·mg protein; 22°C) oxidation of ethanol to acetaldehyde; after prolonged incubation also acetate was detectable. With NAD+ only 2% of the activity was observed. F420 was not reduced. The crude extract also contained F420: NADP+ oxidoreductase (0.45 mol/min·mg protein) that was not active at the pH optimum of ADH. With added acetaldehyde no net reduction of various electron acceptors was measured. However, the acetaldehyde was dismutated to ethanol and acetate by the crude extract. The dismutation was stimulated by NADP+. These findings suggested that not only the dehydrogenation of alcohol but also of aldehyde to acid was coupled to NADP+ reduction. If the reaction was started with acetaldehyde, formed NADPH probably reduced excess aldehyde immediately to ethanol and in this way gave rise to the observed dismutation. Acetate thiokinase activity (0.11 mol/min·mg) but no acetate kinase or phosphotransacetylase activity was observed. It is concluded that during growth on ethanol further oxidation of acetaldehyde does not occur via acetylCoA and acetyl phosphate and hence is not associated with substrate level phosphorylation. The possibility exists that oxidation of both ethanol and acetaldehyde is catalyzed by ADH. Isolation of a Methanobacterium-like strain with ethanol showed that the ability to use primary alcohols also occurs in genera other than Methanogenium.Non-standard abbreviations ADH alcohol dehydrogenase - Ap5ALi3 P1,P5-Di(adenosine-5-)pentaphosphate - DTE dithioerythritol (2,3-dihydroxy-1,4-dithiolbutane) - F420 N-(N-l-lactyl--l-glutamyl)-l-glutamic acid phosphodiester of 7,8-dimethyl-8-hydroxy-5-deazariboflavin-5-phosphate - Mg. Methanogenium - OD578 optical density at 578 nm - PIPES 1,4-piperazine-diethanesulfonic acid - TRICINE N-(2-hydroxy-1,1-bis[hydroxymethyl]methyl)-glycine - Tris 2-amino-2-hydroxy-methylpropane-1,3-diol - U unit (mol substrate/min)  相似文献   

15.
The subcellular distribution and relative amounts of the two isozymes, F1 and F2, of aldehyde dehydrogenase (EC 1.2.1.3) which were recently purified to homogeneity from horse liver (Eckfeldt, J., et al. (1976) J. Biol. Chem.251, 236–240) have been investigated. A fresh horse liver homogenate was fractionated on DEAE-cellulose. The results indicate that approximately 60% of the total pH 7.0 acetaldehyde dehydrogenase activity is due to the F1 isozyme and 40% is due to the F2 isozyme. Several horse livers were then fractionated into subcellular components using a differential centrifugation method. Based on the disulfiram (Antabuse) inhibition and the aldehyde concentration dependence of the enzymatic activity, it appears that the disulfiram-sensitive F1 isozyme (Km acetaldehyde ? 70 μm) is primarily cytosolic and the disulfiram-insensitive F2 isozyme (Km acetaldehyde ? 0.2 μm) is primarily mitochondrial. Fluorescence studies showed that the acetaldehyde dehydrogenase of the intact mitochondria could utilize only the endogenous pyridine nucleotide pool and not externally added NAD. Also, the ethanol dehydrogenase activity was found to be nearly 10 times the total acetaldehyde dehydrogenase activity when assaying a horse liver homogenate at pH 7.0 and with saturating substrates. The significant differences between this work and the results reported in rat liver are discussed with respect to the physiological importance of the cytosolic and mitochondrial aldehyde dehydrogenase during the ethanol oxidation in vivo.  相似文献   

16.
Abstract

The objectives of this study were to develop a robust protocol to measure the rate of hydrogen peroxide (H2O2) production in isolated perfused rat lungs, as an index of oxidative stress, and to determine the cellular sources of the measured H2O2 using the extracellular probe Amplex red (AR). AR was added to the recirculating perfusate in an isolated perfused rat lung. AR’s highly fluorescent oxidation product resorufin was measured in the perfusate. Experiments were carried out without and with rotenone (complex I inhibitor), thenoyltrifluoroacetone (complex II inhibitor), antimycin A (complex III inhibitor), potassium cyanide (complex IV inhibitor), or diohenylene iodonium (inhibitor of flavin-containing enzymes, e.g. NAD(P)H oxidase or NOX) added to the perfusate. We also evaluated the effect of acute changes in oxygen (O2) concentration of ventilation gas on lung rate of H2O2 release into the perfusate. Baseline lung rate of H2O2 release was 8.45?±?0.31 (SEM) nmol/min/g dry wt. Inhibiting mitochondrial complex II reduced this rate by 76%, and inhibiting flavin-containing enzymes reduced it by another 23%. Inhibiting complex I had a small (13%) effect on the rate, whereas inhibiting complex III had no effect. Inhibiting complex IV increased this rate by 310%. Increasing %O2 in the ventilation gas mixture from 15 to 95% had a small (27%) effect on this rate, and this O2-dependent increase was mostly nonmitochondrial. Results suggest complex II as a potentially important source and/or regulator of mitochondrial H2O2, and that most of acute hyperoxia-enhanced lung rate of H2O2 release is from nonmitochondrial rather than mitochondrial sources.  相似文献   

17.
End-product synthesis and enzyme activities involved in pyruvate catabolism, H2 synthesis, and ethanol production in mid-log (OD600  0.25), early stationary (OD600  0.5), and stationary phase (OD600  0.7) cell extracts were determined in Clostridium thermocellum ATCC 27405 grown in batch cultures on cellobiose. Carbon dioxide, hydrogen, ethanol, acetate and formate were major end-products and their production paralleled growth and cellobiose consumption. Lactate dehydrogenase, pyruvate:formate lyase, pyruvate:ferredoxin oxidoreductase, methyl viologen-dependant hydrogenase, ferredoxin-dependant hydrogenase, NADH-dependant hydrogenase, NADPH-dependant hydrogenase, NADH-dependant acetaldehyde dehydrogenase, NADH-dependant alcohol dehydogenase, and NADPH-dependant alcohol dehydrogenase activities were detected in all extracts, while pyruate dehydrogenase and formate dehydrogenase activities were not detected. All hydrogenase activities decreased (2–12-fold) as growth progressed from early exponential to stationary phase. Alcohol dehydrogenase activities fluctuated only marginally (<45%), while lactate dehydrogenase, pyruvate:formate lyase, and pyruvate:ferredoxin oxidoreductase remained constant in all cell extracts. We have proposed a pathway involved in pyruvate catabolism and end-product formation based on enzyme activity profiles in conjunction with bioinformatics analysis.  相似文献   

18.
Pelobacter carbinolicus and P. acetylenicus oxidize ethanol in syntrophic cooperation with methanogens. Cocultures with Methanospirillum hungatei served as model systems for the elucidation of syntrophic ethanol oxidation previously done with the lost “Methanobacillus omelianskii” coculture. During growth on ethanol, both Pelobacter species exhibited NAD+-dependent alcohol dehydrogenase activity. Two different acetaldehyde-oxidizing activities were found: a benzyl viologen-reducing enzyme forming acetate, and a NAD+-reducing enzyme forming acetyl-CoA. Both species synthesized ATP from acetyl-CoA via acetyl phosphate. Comparative 2D-PAGE of ethanol-grown P. carbinolicus revealed enhanced expression of tungsten-dependent acetaldehyde: ferredoxin oxidoreductases and formate dehydrogenase. Tungsten limitation resulted in slower growth and the expression of a molybdenum-dependent isoenzyme. Putative comproportionating hydrogenases and formate dehydrogenase were expressed constitutively and are probably involved in interspecies electron transfer. In ethanol-grown cocultures, the maximum hydrogen partial pressure was about 1,000 Pa (1 mM) while 2 mM formate was produced. The redox potentials of hydrogen and formate released during ethanol oxidation were calculated to be EH2 = -358±12 mV and EHCOOH = -366±19 mV, respectively. Hydrogen and formate formation and degradation further proved that both carriers contributed to interspecies electron transfer. The maximum Gibbs free energy that the Pelobacter species could exploit during growth on ethanol was −35 to −28 kJ per mol ethanol. Both species could be cultivated axenically on acetaldehyde, yielding energy from its disproportionation to ethanol and acetate. Syntrophic cocultures grown on acetoin revealed a two-phase degradation: first acetoin degradation to acetate and ethanol without involvement of the methanogenic partner, and subsequent syntrophic ethanol oxidation. Protein expression and activity patterns of both Pelobacter spp. grown with the named substrates were highly similar suggesting that both share the same steps in ethanol and acetalydehyde metabolism. The early assumption that acetaldehyde is a central intermediate in Pelobacter metabolism was now proven biochemically.  相似文献   

19.
20.
The physiology and biochemistry of Sarcina ventriculi was studied in order to determine adaptations made by the organism to changes in environmental pH. The organism altered carbon and electron flow from acetate, formate and ethanol production at neutral pH, to predominantly ethanol production at pH 3.0. Increased levels of pyruvate dehydrogenase (relative to pyruvate decarboxylase) and acetaldehyde dehydrogenase occurred when the organism was grown at neutral pH, indicating the predominance of carbon flux through the oxidative branch of the pathway for pyruvate metabolism. When the organism was grown at acid pH, there was a significant increase in pyruvate decarboxylase levels and a decrease in acetaldehyde dehydrogenase, causing flux through the non-oxidative branch of the pathway. CO2 reductase and formate dehydrogenase were not regulated as a function of growth pH. Pyruvate dehydrogenase possessed Michaelis-Menten kinetics for pyruvate with an apparent K m of 2.5 mM, whereas pyruvate decarboxylase exhibited sigmoidal kinetics, with a S0.5 of 12.0 mM. Differences in total protein banding patterns from cells grown at pH extremes suggested that synthesis of pyruvate decarboxylase and other enzymes was in part responsible for metabolic regulation of the fermentation products formed.  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