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.
Anaerobic granules degrading pentachlorophenol (PCP) with specific PCP removal activity up to 14.6 mg/g of volatile suspended solids per day were developed in a laboratory-scale anaerobic upflow sludge blanket reactor at 28 degrees C, by using a mixture of acetate, propionate, butyrate, and methanol as the carbon source. The reactor was able to treat synthetic wastewater containing 40 to 60 mg of PCP per liter at a volumetric loading rate of up to 90 mg/liter of reactor volume per day, with a hydraulic retention time of 10.8 to 15 h. PCP removal of more than 99% was achieved. Results of adsorption of PCP by granular biomass indicated that the PCP removal by the granules was due to biodegradation rather than adsorption. A radiotracer assay demonstrated that the PCP-degrading granules mineralized [14C]PCP to 14CH4 and 14CO2. Toxicity test results indicated that syntrophic propionate degraders and acetate-utilizing methanogens were more sensitive to PCP than syntrophic butyrate degraders. The PCP-degrading granules also exhibited a higher tolerance to the inhibition caused by PCP for methane production and degradation of acetate, propionate, and butyrate, compared with anaerobic granules unadapted to PCP.  相似文献   

2.
Of four chlorinated guaiacols, tetrachloroguaiacol at 62 M inhibited acetate methanogenesis, the strongest decreasing activity by 50%. 4,5,6-Trichloroguaiacol, 4,5-dichloroguaiacol, and 4-chloroguaiacol showed 50% inhibition at 0.13, 0.32, and 1.50 mM, respectively. Degradation test results of volatile fatty acids (acetic, propionic, and butyric acid) by anaerobic digester sludge (stored 5 weeks) indicated that syntrophic butyrate degraders of this sludge were more sensitive to tetrachloroguaiacol than acetoclastic methanogens and syntrophic propionate degraders.  相似文献   

3.
Granular sludge from an upflow anaerobic sludge blanket reactor treating synthetic waste water containing a mixture of volatile fatty acids and nitrate showed a removal efficiency of nearly 100% for both nitrogen and carbon. This activity was achieved by a combined process of denitrification and methanogenesis under conditions of surplus carbon. Under batch conditions the two processes proceeded clearly separated in time with first denitrification dominating and excluding methanogenesis. However, as soon as nitrate was depleted, methane production was initiated, showing that the inhibition of methanogenesis by nitrate was reversible. Of the volatile fatty acids supplied to the reactor, i.e. acetate, propionate, and butyrate, the denitrifying population clearly preferred butyrate and propionate even though acetate could also be metabolized. Consequently, growth of syntrophic volatile fatty acid degraders was suppressed by the denitrifiers in cases of low C:N ratios in the medium, leaving acetate as the major substrate for methanogenesis.Abbreviations UASB upflow anaerobic sludge blanket - COD chemical oxygen demand - VFA volatile fatty  相似文献   

4.
The pathway of propionate conversion in a syntrophic coculture of Smithella propionica and Methanospirillum hungatei JF1 was investigated by (13)C-NMR spectroscopy. Cocultures produced acetate and butyrate from propionate. [3-(13)C]propionate was converted to [2-(13)C]acetate, with no [1-(13)C]acetate formed. Butyrate from [3-(13)C]propionate was labeled at the C2 and C4 positions in a ratio of about 1:1.5. Double-labeled propionate (2,3-(13)C) yielded not only double-labeled acetate but also single-labeled acetate at the C1 or C2 position. Most butyrate formed from [2,3-(13)C]propionate was also double labeled in either the C1 and C2 atoms or the C3 and C4 atoms in a ratio of about 1:1.5. Smaller amounts of single-labeled butyrate and other combinations were also produced. 1-(13)C-labeled propionate yielded both [1-(13)C]acetate and [2-(13)C]acetate. When (13)C-labeled bicarbonate was present, label was not incorporated into acetate, propionate, or butyrate. In each of the incubations described above, (13)C was never recovered in bicarbonate or methane. These results indicate that S. propionica does not degrade propionate via the methyl-malonyl-coenzyme A (CoA) pathway or any other of the known pathways, such as the acryloyl-CoA pathway or the reductive carboxylation pathway. Our results strongly suggest that propionate is dismutated to acetate and butyrate via a six-carbon intermediate.  相似文献   

5.
1. Interactions in the rates of consumption of acetate, propionate and butyrate in sheep liver mitochondria were examined in the presence and absence of l-malate and alpha-oxoglutarate. 2. Acetate was not consumed in absence of ancillary substrate but utilization of acetate (7.2nmol/min per mg of protein) occurred in the presence of alpha-oxoglutarate. This consumption was abolished by propionate or butyrate but the presence of acetate did not affect consumption of propionate or butyrate. 3. Propionate consumption (10.1nmol/min per mg of protein) was unaffected by malate but was stimulated by 63% by butyrate or by 180% by alpha-oxoglutarate. 4. Butyrate consumption (3.3nmol/min per mg of protein) was stimulated by 117% by malate, by 151% by propionate and by 310% by alpha-oxoglutarate. 5. In the absence of ancillary substrates the maximum rate of total volatile fatty acid utilization (24.7nmol/min per mg of protein) occurred with a mixture of propionate and butyrate. When both propionate and butyrate were present total consumption was not affected by malate but was stimulated by 24% by alpha-oxoglutarate. With alpha-oxoglutarate present, propionate and butyrate each decreased the other's consumption by about 26%, but the total utilization was the greatest observed. 6. The inhibition of acetate consumption by propionate or butyrate is unexplained, but the remaining effects are consistent with an interaction of propionate and butyrate through oxaloacetate together with a general limitation imposed by a need for GTP to rephosphorylate AMP formed during activation of the volatile fatty acids.  相似文献   

6.
The effects of acetate, propionate, and butyrate on the anaerobic thermophilic conversion of propionate by methanogenic sludge and by enriched propionate-oxidizing bacteria in syntrophy with Methanobacterium thermoautotrophicum delta H were studied. The methanogenic sludge was cultivated in an upflow anaerobic sludge bed (UASB) reactor fed with propionate (35 mM) as the sole substrate for a period of 80 days. Propionate degradation was shown to be severely inhibited by the addition of 50 mM acetate to the influent of the UASB reactor. The inhibitory effect remained even when the acetate concentration in the effluent was below the level of detection. Recovery of propionate oxidation occurred only when acetate was omitted from the influent medium. Propionate degradation by the methanogenic sludge in the UASB reactor was not affected by the addition of an equimolar concentration (35 mM) of butyrate to the influent. However, butyrate had a strong inhibitory effect on the growth of the propionate-oxidizing enrichment culture. In that case, the conversion of propionate was almost completely inhibited at a butyrate concentration of 10 mM. However, addition of a butyrate-oxidizing enrichment culture abolished the inhibitory effect, and propionate oxidation was even stimulated. All experiments were conducted at pH 7.0 to 7.7. The thermophilic syntrophic culture showed a sensitivity to acetate and propionate similar to that of mesophilic cultures described in the literature. Additions of butyrate or acetate to the propionate medium had no effect on the hydrogen partial pressure in the biogas of an UASB reactor, nor was the hydrogen partial pressure in propionate-degrading cultures affected by the two acids.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

7.
The production of polyhydroxyalkanoates from acetate and propionate by two mixed cultures well adapted to each of these substrates was evaluated. Sludge fed with acetate (A), produced a homopolymer of hydroxybutyrate (HB), whereas sludge fed with propionate (P) produced a copolymer of HB and HV (hydoxyvalerate). Switching the substrate feeds, propionate to sludge A and acetate to culture P, a terpolymer of HB, HV and hydroxymethylvalerate (HMV) was obtained with culture A and a copolymer of P(HB/HV) by sludge P. Regardless of the population used, the polymer yield and productivity were much higher for acetate than for propionate. Feeding a mixture of acetate and propionate, in equal parts, to both cultures resulted in an increase of HV units produced per C mol of propionate consumed, relative to the situation where only propionate was used. The individual use of butyrate and valerate by culture A was also studied. Butyrate produced a homopolymer whereas valerate was stored as a terpolymer of P(HB/HV/HMV). The polymer yields on acetate and butyrate were higher than those on propionate and valerate. The polymer productivity was higher for acetate and propionate than for butyrate and valerate. Results showed that the polymer composition, and consequently the polymer properties, could be manipulated by varying the volatile fatty acid feed composition and/or the population.  相似文献   

8.
Short-chain fatty acids (SCFAs), primarily acetate, propionate, and butyrate, are metabolites formed by gut microbiota from complex dietary carbohydrates. Butyrate and acetate were reported to protect against diet-induced obesity without causing hypophagia, while propionate was shown to reduce food intake. However, the underlying mechanisms for these effects are unclear. It was suggested that SCFAs may regulate gut hormones via their endogenous receptors Free fatty acid receptors 2 (FFAR2) and 3 (FFAR3), but direct evidence is lacking. We examined the effects of SCFA administration in mice, and show that butyrate, propionate, and acetate all protected against diet-induced obesity and insulin resistance. Butyrate and propionate, but not acetate, induce gut hormones and reduce food intake. As FFAR3 is the common receptor activated by butyrate and propionate, we examined these effects in FFAR3-deficient mice. The effects of butyrate and propionate on body weight and food intake are independent of FFAR3. In addition, FFAR3 plays a minor role in butyrate stimulation of Glucagon-like peptide-1, and is not required for butyrate- and propionate-dependent induction of Glucose-dependent insulinotropic peptide. Finally, FFAR3-deficient mice show normal body weight and glucose homeostasis. Stimulation of gut hormones and food intake inhibition by butyrate and propionate may represent a novel mechanism by which gut microbiota regulates host metabolism. These effects are largely intact in FFAR3-deficient mice, indicating additional mediators are required for these beneficial effects.  相似文献   

9.
Washed excised roots of rice (Oryza sativa) produced H(2), CH(4) and fatty acids (millimolar concentrations of acetate, propionate, butyrate; micromolar concentrations of isovalerate, valerate) when incubated under anoxic conditions. Surface sterilization of the root material resulted in the inactivation of the production of CH(4), a strong reduction of the production of fatty acids and a transient (75 h) but complete inhibition of the production of H(2). Radioactive bicarbonate was incorporated into CH(4), acetate, propionate and butyrate. About 20-40% of the fatty acid carbon originated from CO(2) reduction. In the presence of phosphate, CH(4) was exclusively produced from H(2)/CO(2), since phosphate selectively inhibited acetoclastic methanogenesis. Acetoclastic methanogenesis was also selectively inhibited by methyl fluoride, while chloroform or 2-bromoethane sulfonate inhibited CH(4) production completely. Production of CH(4), acetate, propionate and butyrate from H(2)/CO(2) was always exergonic with Gibbs free energies <-20 kJ mol(-1) product. Chloroform inhibited the production of acetate and the incorporation of radioactive CO(2) into acetate. Simultaneously, H(2) was no longer consumed and accumulated, indicating that acetate was produced from H(2)/CO(2). Chloroform also resulted in increased production of propionate and butyrate whose formation from CO(2) became more exergonic upon addition of chloroform. Nevertheless, the incorporation of radioactive CO(2) into propionate and butyrate was inhibited by chloroform. The accumulation of propionate and butyrate in the presence of chloroform probably occurred by fermentation of organic matter, rather than by reduction of acetate and CO(2). [U-(14)C]Glucose was indeed converted to acetate, propionate, butyrate, CO(2) and CH(4). Radioactive acetate, CO(2) and CH(4) were also products of the degradation of [U-(14)C]cellulose and [U-(14)C]xylose. Addition of chloroform and methyl fluoride did not affect the product spectrum of [U-(14)C]glucose degradation. The application of combinations of selective inhibitors may be useful to elucidate anaerobic metabolic pathways in mixed microbial cultures and natural microbial communities.  相似文献   

10.
Short-chain fatty acids (SCFA) are produced by fermentation of water-soluble fiber by anaerobic bacteria in the large bowel. Fiber-rich diets decrease the risk of developing inflammatory bowel disease (IBD) and butyrate enemas are effective as a therapy in some patients. Crohn's disease, one form of IBD, appears to involve an exagerated T helper-1 (Th1) lymphocyte phenotype, characterised by production of interleukin (IL)-2 and interferon (IFN)-gamma, that drives the inflammation. To examine whether SCFA influence pro- and anti-inflammatory cytokine production, rat mesenteric lymph node lymphocytes were cultured in the presence of acetate (10 mM), butyrate (1.5 mM) or propionate (2 mM) and the production of cytokines in response to concanavalin A determined. Butyrate, but not acetate or propionate, inhibited lymphocyte proliferation and IL-2 production. Acetate and propionate were able to partly prevent the inhibitory effect of butyrate on IL-2 production. Acetate and propionate increased IFN-gamma production, whereas butyrate inhibited it. Acetate and propionate in combination were able to prevent the inhibitory effect of butyrate on IFN-gamma production. IL-4 was not detected in any cultures. Acetate and propionate increased IL-10 production, which was not affected by butyrate. It is concluded that butyrate significantly inhibits Th1-type responses and that this might explain the therapeutic effect of butyrate in IBD patients. Acetate and propionate have less marked modulatory actions, and in some cases have effects that oppose those of butyrate. A combination of the three SCFA causes a shift in the T helper lymphocyte phenotype towards a more anti-inflammatory phenotype and this might explain the protective effects of fiber.  相似文献   

11.
1. Two mM propionate or butyrate inhibited the mitochondrial uptake of acetate by rat hepatocytes. 2. With propionate the inhibition was so strong that the net formation of acetate in the cytoplasm, usually masked by the mitochondrial uptake, appeared directly as a net output of acetate into the medium; showing that this net formation of acetate, reported by [Crabtree B., Gordon M.-J. and Christie S. L. (1990) Biochem. J. 270, 219-225] is not an artefact arising from a misinterpretation of isotopic data. 3. The results also suggest that propionate and butyrate inhibit peroxisomal metabolism.  相似文献   

12.
Summary The degradation of acetate, propionate and butyrate was monitored during start-up of five lab-scale methanogenic fluidized bed reactors on an artificially prepared waste water. The acetate concentration in the reactor content was found to influence the degradation of propionate but not of butyrate. In general, at acetate levels over 200 mg/l the degradation of propionate was below 60%, whereas the degradation was complete at acetate levels under 100 mg/l. The rationale of the inhibition of propionate degradation by acetate is discussed.  相似文献   

13.
A mixed culture from an anaerobic biowaste digester was enriched on propionate and used to investigate interspecies hydrogen transfer in dependence of spatial distances between propionate degraders and methanogens. From 20.3 mM propionate, 20.8 mM acetate and 15.5 mM methane were formed. Maximum specific propionate oxidation and methane formation rates were 49 and 23 mmol?mg?1?day?1, respectively. Propionate oxidation was inhibited by only 20 mM acetate by about 50 %. Intermediate formate formation during inhibited methanogensis was observed. The spatial distribution and the biovolume fraction of propionate degraders and of methanogens in relation to the total population during aggregate formation were determined. Measurements of interbacterial distances were conducted with fluorescence in situ hybridization by application of group-specific 16S rRNA-targeted probes and 3D image analyses. With increasing incubation time, floc formation and growth up to 54 μm were observed. Propionate degraders and methanogens were distributed randomly in the flocs. The methanogenic biovolume fraction was high at the beginning and remained constant over 42 days, whereas the fraction of propionate degraders increased with time during propionate feeding. Interbacterial distances between propionate degraders and methanogens decreased with time from 5.30 to 0.29 μm, causing an increase of the maximum possible hydrogen flux from 1.1 to 10.3 nmol?ml?1?min?1. The maximum possible hydrogen flux was always higher than the hydrogen formation and consumption rate, indicating that reducing the interspecies distance by aggregation is advantageous in complex ecosystems.  相似文献   

14.
The pathway of propionate conversion in a syntrophic coculture of Smithella propionica and Methanospirillum hungatei JF1 was investigated by 13C-NMR spectroscopy. Cocultures produced acetate and butyrate from propionate. [3-13C]propionate was converted to [2-13C]acetate, with no [1-13C]acetate formed. Butyrate from [3-13C]propionate was labeled at the C2 and C4 positions in a ratio of about 1:1.5. Double-labeled propionate (2,3-13C) yielded not only double-labeled acetate but also single-labeled acetate at the C1 or C2 position. Most butyrate formed from [2,3-13C]propionate was also double labeled in either the C1 and C2 atoms or the C3 and C4 atoms in a ratio of about 1:1.5. Smaller amounts of single-labeled butyrate and other combinations were also produced. 1-13C-labeled propionate yielded both [1-13C]acetate and [2-13C]acetate. When 13C-labeled bicarbonate was present, label was not incorporated into acetate, propionate, or butyrate. In each of the incubations described above, 13C was never recovered in bicarbonate or methane. These results indicate that S. propionica does not degrade propionate via the methyl-malonyl-coenzyme A (CoA) pathway or any other of the known pathways, such as the acryloyl-CoA pathway or the reductive carboxylation pathway. Our results strongly suggest that propionate is dismutated to acetate and butyrate via a six-carbon intermediate.  相似文献   

15.
In this study, the widely used anaerobic digestion model (ADM1) was used in order to simulate the inhibition of three pharmaceuticals, propranolol hydrochloride, ofloxacin and diclofenac sodium, on two groups of microorganisms, acetogens and acetoclastic methanogens, the most sensitive microorganisms groups involved in the anaerobic digestion process. The specific maximum consumption rate and saturation constant of acetate and propionate degraders were estimated through fitting the model to experimental data taken from continuous and batch experiments. A modified non-competitive inhibition function was used, and the inhibition constants were estimated using data from Batch experiments conducted at various concentrations of pharmaceuticals using enriched cultures with propionate and acetate degraders. It was found that propranolol hydrochloride was the most inhibitory pharmaceutical to both microorganisms groups.  相似文献   

16.
Smith EA  Macfarlane GT 《Anaerobe》1997,3(5):327-337
The abilities of slurries of human faecal bacteria to ferment 20 different amino acids were investigated in batch culture incubations. Ammonia, short chain fatty acids, and in some cases, amines, were the principal products of dissimilatory metabolism. The types of SCFA produced were dependent on the chemical compositions of the test substrates. Thus, acetate and butyrate were formed from the acidic amino acid glutamate, while acetate and propionate predominated in aspartate fermentations. Breakdown of the basic amino acids lysine and arginine was rapid, and yielded butyrate and acetate, and ornithine and citrulline, respectively. The major products of histidine deamination were also acetate and butyrate. However, fermentation of sulphur-containing amino acids was slow and incomplete. Acetate, propionate and butyrate were formed from cysteine, whereas the main products of methionine metabolism were propionate and butyrate. The simple aliphatic amino acids alanine and glycine were fermented to acetate, propionate and butyrate, and acetate and methylamine, respectively. Branched-chain amino acids were slowly fermented by colonic bacteria, with the main acidic products being branched-chain fatty acids one carbon atom shorter than the parent amino acid. Low concentrations of amines were also detected in these fermentations. Aliphatic-hydroxy amino acids were rapidly deaminated by large intestinal microorganisms. Serine was primarily fermented to acetate and butyrate, while threonine was mainly metabolised to propionate. Proline was poorly utilized by intestinal bacteria, but hydroxyproline was efficiently fermented to acetate and propionate. The aromatic amino acids tyrosine, phenylalanine and tryptophan were broken down to a range of phenolic and indolic compounds.  相似文献   

17.
Microbial production of different alipathic esters with flavour characteristic has been studied. Lyophilized whole cells of Rhizopus oryzae CBS 112-07 were found to be particularly suitable to catalyse the synthesis of different flavour esters (hexyl acetate, propionate, butyrate, caprylate; geranyl acetate, propionate, butyrate and 2- and 3-methylbutyl acetate, butyrate) in n-heptane. This strain was therefore utilized for the semipreparative production of geranyl butyrate by semicontinous and continous addition of the substrates with satisfactory yields (144 g l–1 in 264 h and 142 g l–1 in 48 h respectively).  相似文献   

18.
Effect of precursors on biosynthesis of monensins A and B   总被引:1,自引:0,他引:1  
Precursors of monensins (acetate, propionate, butyrate, isobutyrate) affect the total production and the relative proportion of monensins A and B. Addition of propionate into the fermentation medium causes a prevalence of monensin B whereas butyrate and isobutyrate stimulate the production of monensin A and suppress the production of monensin B.  相似文献   

19.
Acarbose inhibits starch digestion in the human small intestine. This increases the amount of starch available for microbial fermentation to acetate, propionate, and butyrate in the colon. Relatively large amounts of butyrate are produced from starch by colonic microbes. Colonic epithelial cells use butyrate as an energy source, and butyrate causes the differentiation of colon cancer cells. In this study we investigated whether colonic fermentation pathways changed during treatment with acarbose. We examined fermentations by fecal suspensions obtained from subjects who participated in an acarbose-placebo crossover trial. After incubation with [1-13C]glucose and 12CO2 or with unlabeled glucose and 13CO2, the distribution of 13C in product C atoms was determined by nuclear magnetic resonance spectrometry and gas chromatography-mass spectrometry. Regardless of the treatment, acetate, propionate, and butyrate were produced from pyruvate formed by the Embden-Meyerhof-Parnas pathway. Considerable amounts of acetate were also formed by the reduction of CO2. Butyrate formation from glucose increased and propionate formation decreased with acarbose treatment. Concomitantly, the amounts of CO2 reduced to acetate were 30% of the total acetate in untreated subjects and 17% of the total acetate in the treated subjects. The acetate, propionate, and butyrate concentrations were 57, 20, and 23% of the total final concentrations, respectively, for the untreated subjects and 57, 13, and 30% of the total final concentrations, respectively, for the treated subjects.  相似文献   

20.
1. Benzylmercapturic acid and hippuric acid were isolated from the urine of rats that had been injected subcutaneously with benzyl acetate. 2. 1-Menaphthylmercapturic acid and 1-naphthoic acid were isolated from the urine of rats after the subcutaneous injection of each of the following compounds: 1-menaphthyl alcohol and its acetate, propionate, butyrate and benzoate esters. 3. A quantitative method for determining 1-menaphthylmercapturic acid in urine was developed and used to measure the excretion of this compound in the urine of rats in the 4-day period after the subcutaneous injection of 1-menaphthyl alcohol and its acetate, propionate, butyrate and benzoate esters. 4. Chromatographic evidence was obtained for the presence of S-(1-menaphthyl)glutathione and S-(1-menaphthyl)-l-cysteine in bile collected from rats with cannulated bile ducts after the animals had been injected subcutaneously with each of the following compounds: S-(1-menaphthyl)glutathione, 1-menaphthyl acetate, propionate and butyrate. 5. Benzylmercapturic acid and 1-menaphthylmercapturic acid were isolated from the urine of rats that had been injected with sodium benzyl sulphate and sodium 1-menaphthyl sulphate respectively.  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