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.
Microbial diversity of a mesophilic hydrogen-producing sludge   总被引:7,自引:0,他引:7  
A hydrogen-producing sludge degraded 99% of glucose at 36 degrees C and pH 5.5, producing a methane-free biogas (comprising 64% hydrogen) and an effluent comprising mostly butyrate, acetate, and ethanol. The yield was 0.26 l H2 g(-1) glucose and the production rate per gram of volatile suspended solids was 4.6 1 H2 day(-1). A 16S rDNA library was constructed from the sludge for microbial species determination. A total of 96 clones were selected for plasmids recovery, screened by denaturing gradient gel electrophoresis, and sequenced for rDNA. Based on the phylogenetic analysis of the rDNA sequences, 64.6% of all the clones were affiliated with three Clostridium species (Clostridiaceae), 18.8% with Enterobacteriaceae, and 3.1% with Streptococcus bovis (Streptococcaceae). The remaining 13.5% belonged to eight operational taxonomic units, the affiliations of which were not identified.  相似文献   

2.
In different cell systems, the lipido-sterolic extract of Serenoa repens (LSESr, Permixon inhibits both type 1 and type 2 5alpha-reductase activity (5alphaR1 and 5alphaR2). LSESr is mainly constituted of fatty acids (90+/-5%) essentially as free fatty acids (80%). Among these free fatty acids, the main components are oleic and lauric acids which represent 65% and linoleic and myristic acids 15%.To evaluate the inhibitory effect of the different components of LSESr on 5alphaR1 or 5alphaR2 activity, the corresponding type 1 and type 2 human genes have been cloned and expressed in the baculovirus-directed insect cell expression system Sf9. The cells were incubated at pH 5.5 (5alphaR2) and pH 7.4 (5alphaR1) with 1 or 3nM testosterone in presence or absence of various concentrations of LSESr or of its different components. Dihydrotestosterone formation was measured with an automatic system combining HPLC and an on-line radiodetector.The inhibition of 5alphaR1 and 5alphaR2 activity was only observed with free fatty acids: esterified fatty acids, alcohols as well as sterols assayed were inactive. A specificity of the fatty acids in 5alphaR1 or 5alphaR2 inhibition has been found. Long unsaturated chains (oleic and linolenic) were active (IC(50)=4+/-2 and 13+/-3 microg/ml, respectively) on 5alphaR1 but to a much lesser extent (IC(50)>100 and 35+/-21 microg/ml, respectively) on 5alphaR2. Palmitic and stearic acids were inactive on the two isoforms. Lauric acid was active on 5alphaR1 (IC(50)=17+/-3 microg/ml) and 5alphaR2 (IC(50)=19+/-9 microg/ml). The inhibitory activity of myristic acid was evaluated on 5alphaR2 only and found active on this isoform (IC(50)=4+/-2 microg/ml).The dual inhibitory activity of LSESr on 5alpha-reductase type 1 and type 2 can be attributed to its high content in free fatty acids.  相似文献   

3.
Prolonged infusions of lipid and heparin that achieve high physiological free fatty acid (FFA) concentrations inhibit hepatic (and peripheral) insulin sensitivity in humans. These infusions are composed largely of polyunsaturated fatty acids (PUFA; linoleic and linolenic). It is not known whether fatty acid composition per se affects hepatic glucose metabolism in humans. To address this issue, we examined the impact of enteral infusions of either palm oil (48% palmitic, 35% oleic, and 8% linoleic acids) or safflower oil (6% palmitic, 12% oleic, 74% linoleic acids) in 14 obese nondiabetic subjects. (2)H(2)O was administered to determine the contribution of gluconeogenesis to endogenous glucose production (EGP), and a primed continuous infusion of [6,6-(2)H]glucose was administered to assess glucose appearance. As a result of the lipid infusions, plasma FFA concentrations increased significantly in both the palm oil (507.5 +/- 47.4 to 939.3 +/- 61.3 micromol/l, P < 0.01) and safflower oil (588.2.0 +/- 43.0 to 857.8 +/- 68.7 micromol/l, P < 0.01) groups after 4 h. EGP was similar at baseline (12.4 +/- 1.8 vs. 11.2 +/- 1.0 micromol x kg FFM(-1) x min(-1)). During a somatostatin-insulin clamp, the glucose infusion rate was significantly lower (AUC glucose infusion rate 195.8 +/- 50.7 vs. 377.8 +/- 38.0 micromol/kg FFM, P < 0.01), and rates of EGP were significantly higher (10.7 +/- 1.4 vs. 6.5 +/- 1.5 micromol x kg FFM(-1) x min(-1), P < 0.01) after palm oil compared with safflower oil, respectively. Baseline rates of gluconeogenesis and glycogenolysis were also similar. However, after lipid infusion, rates of glycogenolysis were suppressed by safflower oil but not by palm oil. Thus these studies demonstrate, for the first time in humans, a differential effect of saturated fatty acids and PUFA on hepatic glucose metabolism.  相似文献   

4.
pH- and fatty acid-induced conformational changes in human serum albumin were investigated by fluorescence-energy transfer, determining the distance between Trp-214 and bound bilirubin at 25 degrees C. This distance changes significantly with the pH, being 2.52 +/- 0.01 nm at pH 6, 2.31 +/- 0.04 nm at pH 9, 2.13 +/- 0.07 nm at pH 11.0 and 2.77 nm at pH 11.9. The influence of different fatty acids on the distance was also determined. At pH 7.4 medium-chain fatty acids seem to increase this distance, whereas long-chain fatty acids, at low concentrations, decrease the distance between the two chromophores. The contraction of the protein carrying long-chain saturated fatty acids is even more pronounced at pH 9.  相似文献   

5.
The performance of a mesophilic two-stage system generating hydrogen and methane continuously from sucrose (10-30 g/L) was investigated. A hydrogen-generating CSTR followed by an upflow anaerobic filter were both inoculated with anaerobically digested sewage sludge, and ORP, pH, gas output, %H(2), %CH(4) and %CO(2) monitored. pH was controlled with NaOH, KOH or Ca(OH)(2). Using NaOH as alkali with 10 g/L sucrose, yields of 1.62 +/- 0.2 mol H(2)/mol hexose added and 323 mL CH(4)/gCOD added to the hydrogen and methane reactors respectively were achieved. The overall chemical oxygen demand (COD) reduction was 92.6% with 0.90 +/- 0.1 g/L sodium and 316 +/- 40 mg/L residual acetate in the methane reactor. Operation at 20 g/L sucrose and NaOH as alkali led to impaired volatile fatty acid (VFA) degradation in the methane reactor with 2.23 +/- 0.2 g/L sodium, 1,885 mg/L residual acetate, a hydrogen yield of 1.47 +/- 0.1 mol/mol hexose added, a methane yield of 294 mL/gCOD added and an overall COD reduction of 83%. Using Ca(OH)(2) as alkali with 20 g/L sucrose gave a hydrogen yield of 1.29 +/- 0.3 mol/mol hexose added, a methane yield of 337 mL/gCOD added and improved the overall COD reduction to 91% with residual acetate concentrations of 522 +/- 87 mg/L. Operation at 30 g/L sucrose with Ca(OH)(2) gave poorer overall COD reduction (68%), a hydrogen yield of 1.47 +/- 0.2 mol/mol hexose added, a methane yield of 138 mL/gCOD added and residual acetate 7,343 +/- 715 mg/L. It was shown that sodium toxicity and overloading are important issues for successful anaerobic digestion of effluent from biohydrogen reactors in high rate systems.  相似文献   

6.
The effects of L-carnitine on myocardial glycolysis, glucose oxidation, and palmitate oxidation were determined in isolated working rat hearts. Hearts were perfused under aerobic conditions with perfusate containing either 11 mM [2-3H/U-14C]glucose in the presence or absence of 1.2 mM palmitate or 11 mM glucose and 1.2 mM [1-14C]palmitate. Myocardial carnitine levels were elevated by perfusing hearts with 10 mM L-carnitine. A 60-min perfusion period resulted in significant increases in total myocardial carnitine from 4376 +/- 211 to 9496 +/- 473 nmol/g dry weight. Glycolysis (measured as 3H2O production) was unchanged in carnitine-treated hearts perfused in the absence of fatty acids (4418 +/- 300 versus 4547 +/- 600 nmol glucose/g dry weight.min). If 1.2 mM palmitate was present in the perfusate, glycolysis decreased almost 2-fold compared with hearts perfused in the absence of fatty acids. In carnitine-treated hearts this drop in glycolysis did not occur (glycolytic rates were 2911 +/- 231 to 4629 +/- 460 nmol glucose/g dry weight.min, in control and carnitine-treated hearts, respectively. Compared with control hearts, glucose oxidation rates (measured as 14CO2 production from [U-14C]glucose) were unaltered in carnitine-treated hearts perfused in the absence of fatty acids (1819 +/- 169 versus 2026 +/- 171 nmol glucose/g dry weight.min, respectively). In the presence of 1.2 mM palmitate, glucose oxidation decreased dramatically in control hearts (11-fold). In carnitine-treated hearts, however, glucose oxidation was significantly greater than control hearts under these conditions (158 +/- 21 to 454 +/- 85 nmol glucose/g dry weight.min, in control and carnitine-treated hearts, respectively). Palmitate oxidation rates (measured as 14CO2 production from [1-14C]palmitate) decreased in the carnitine-treated hearts from 728 +/- 61 to 572 +/- 111 nmol palmitate/g dry weight.min. This probably occurred secondary to an increase in overall ATP production from glucose oxidation (from 5.4 to 14.5% of steady state myocardial ATP production). The results reported in this study provide direct evidence that carnitine can stimulate glucose oxidation in the intact fatty acid perfused heart. This probably occurs secondary to facilitating the intramitochondrial transfer of acetyl groups from acetyl-CoA to acetylcarnitine, thereby relieving inhibition of the pyruvate dehydrogenase complex.  相似文献   

7.
The effect of propionate concentrations on biodegradation of human waste (night soil) was studied at 10 degrees C. Propionate was toxic for the biomethanation at all the pH tested (6.0, 7.0 and 8.0). The maximum reduction in biogas production in presence of 200 mM propionate was observed at pH 7.0 followed by 8.0. The methane content in biogas also followed a similar trend and at pH 7.0 an 11.5% decrease was observed. Propionate caused the reduction of methanogenic count by an approximately 2log value. Total volatile fatty acids increased with the increase in propionate concentration and particularly accumulation of propionate was observed. The results were also compared with the 30 degrees C fermentation.  相似文献   

8.
To determine whether the effects of fatty acids on the diabetic heart during ischemia involve altered glycolytic ATP and proton production, we measured energetics and intracellular pH (pH(i)) by using (31)P NMR spectroscopy plus [2-(3)H]glucose uptake in isolated rat hearts. Hearts from 7-wk streptozotocin diabetic and control rats, perfused with buffer containing 11 mM glucose, with or without 1.2 mM palmitate or the ketone bodies, 4 mM beta-hydroxybutyrate plus 1 mM acetoacetate, were subjected to 32 min of low-flow (0.3 ml x g wet wt(-1) x min(-1)) ischemia, followed by 32 min of reperfusion. In control rat hearts, neither palmitate nor ketone bodies altered the recovery of contractile function. Diabetic rat hearts perfused with glucose alone or with ketone bodies, had functional recoveries 50% lower than those of the control hearts, but palmitate restored recovery to control levels. In a parallel group with the functional recoveries, palmitate prevented the 54% faster loss of ATP in the diabetic, glucose-perfused rat hearts during ischemia, but had no effect on the rate of ATP depletion in control hearts. Palmitate decreased total glucose uptake in control rat hearts during low-flow ischemia, from 106 +/- 17 to 52 +/- 12 micromol/g wet wt, but did not alter the total glucose uptake in the diabetic rat hearts, which was 42 +/- 5 micromol/g wet wt. Recovery of contractile function was unrelated to pH(i) during ischemia; the glucose-perfused control and palmitate-perfused diabetic hearts had end-ischemic pH(i) values that were significantly different at 6.36 +/- 0.04 and 6.60 +/- 0.02, respectively, but had similar functional recoveries, whereas the glucose-perfused diabetic hearts had significantly lower functional recoveries, but their pH(i) was 6.49 +/- 0.04. We conclude that fatty acids, but not ketone bodies, protect the diabetic heart by decreasing ATP depletion, with neither having detrimental effects on the normal rat heart during low-flow ischemia.  相似文献   

9.
The metabolic impact of the murine adipocyte fatty acid-binding protein (AFABP/aP2) on lipid metabolism was investigated in the AFABP/aP2(-/-) mouse and compared with wild-type C57BL/6J littermates. Mice were weaned on a high-fat diet (59% of energy from fat) and acclimated to meal feeding. Stable isotopes were administered, and indirect calorimetry was performed to quantitate fatty acid flux, dietary fatty acid utilization, and substrate oxidation. Consistent with previous in situ and in vitro studies, fasting serum nonesterified fatty acid (NEFA) release was significantly reduced in AFABP/aP2(-/-) (17.1 +/- 9.0 vs. 51.9 +/- 22.9 mg.kg(-1).min(-1)). AFABP/aP2(-/-) exhibited higher serum NEFA (1.4 +/- 0.6 vs. 0.8 +/- 0.4 mmol/l, AFABP/aP2(-/-) vs. C57BL/6J, respectively) and triacylglycerol (TAG; 0.23 +/- 0.09 vs. 0.13 +/- 0.10 mmol/l) and accumulated more TAG in liver tissue (2.9 +/- 2.3 vs. 1.1 +/- 0.8% wet wt) in the fasted state. For the liver-TAG pool, 16.4 +/- 7.3% of TAG-fatty acids were derived from serum NEFA in AFABP/aP2(-/-). In contrast, a significantly greater portion of C57BL/6J liver-TAG was derived from serum NEFA (42.3 +/- 25.5%) during tracer infusion. For adipose-TAG stores, only 0.29 +/- 0.04% was derived from serum NEFA in AFABP/aP2(-/-), and, in C57BL/6J, 1.85 +/- 0.97% of adipose-TAG was derived from NEFA. In addition, AFABP/aP2(-/-) preferentially oxidized glucose relative to fatty acids in the fed state. These data demonstrate that in vivo disruption of AFABP/aP2(-/-) leads to changes in the following two major metabolic processes: 1) decreased adipose NEFA efflux and 2) preferential utilization of glucose relative to fatty acids.  相似文献   

10.
The ability of the neonatal rat to oxidize the branched-chain amino acids leucine and valine and their corresponding keto acids was evaluated. In vivo, about 20% of orally administered labeled amino or keto acids were oxidized in 6 h, after which time little further oxidation occurred. In perfused neonatal liver the amino acids were oxidized at only 5-10% the rate of the keto acids. The oxidation of the keto acids showed a saturable dependence on concentration. The decarboxylation of ketoisocaproate (KIC) had a maximal rate of 40.1 +/- 1.6 mumol/h/g liver with an apparent Km of 0.27 +/- 0.03 mM, and decarboxylation of ketoisovalerate (KIV) had a maximal rate of 37.9 +/- 1.9 mumol/h/g liver and an apparent Km of 0.28 +/- 0.04 mM. KIC was ketogenic, producing mainly acetoacetate at a maximal rate of 44.5 +/- 1.6 mumol/h/g liver with an apparent Km of 0.27 +/- 0.03 mM. On the other hand, KIV was not gluconeogenic, although the perfused neonatal liver was able to produce glucose from lactate. During liver perfusion, KIV did not produce measurable quantities of either propionic or beta-aminoisobutyric acids, which are possible end products of KIV metabolism. Decanoic acid inhibited the decarboxylation of both keto acids to the same extent with a maximal effect at 0.4 mM fatty acid. At saturating levels, KIC was less ketogenic than decanoate. Inhibition of endogenous fatty acid oxidation by 2-tetradecylglycidic acid had no effect on keto acid oxidation. These data suggest that branched-chain amino acids derived from milk proteins are probably not quantitatively significant sources of either ketone bodies or glucose in the neonatal rat.  相似文献   

11.
The antimicrobial activity of C2-C18 fatty acids was determined in vitro in cultures of two strains of Escherichia coli grown on glucose. Antimicrobial activity was expressed as IC50 (a concentration at which only 50% of the initial glucose in the cultures was utilized). Utilization of glucose was inhibited by caprylic acid (IC50 0.30-0.85 g/L) and capric acid (IC50 1.25-2.03 g/L). Neither short-chain fatty acids (C2-C6) nor fatty acids with longer chain (C12-C18) influenced substrate utilization. Caproic acid, however, decreased cell yield in cultures of E. coli in a dose-dependent manner. No inhibition of glucose utilization was produced with unsaturated fatty acids, oleic and linoleic. Calcium ions added in excess reversed the antimicrobial effect of capric acid, but not that of caprylic acid. Antimicrobial activity of caprylic and capric acid decreased when the bacteria were grown in the presence of straw particles, or repeatedly subcultured in a medium containing these compounds at low concentrations. Counts of viable bacteria determined by plating decreased after incubation with caprylic and capric acid (30 min; 1 g/L) at pH 5.2 from > 10(9) to approximately 10(2)/mL. A reduction of a mere 0.94-1.96 log10 CFU was observed at pH 6.5-6.6. It can be concluded that caprylic acid, and to a lesser extent also capric acid, has a significant antimicrobial activity toward E. coli. Effects of other fatty acids were not significant or absent.  相似文献   

12.
This study characterizes the response of the microbial communities of a laboratory-scale mesophilic biogas process, fed with a synthetic substrate based on cellulose and egg albumin, to single pulses of glucose overloading (15 or 25 times the daily feed based on VS). The microbial biomass and community structure were determined from analyses of membrane phospholipids. The ratio between phospholipid fatty acids (PLFAs; eubacteria and eucaryotes) and di-ethers (PLEL; archaea) suggested that methanogens constituted 4-8% of the microbial biomass. The glucose addition resulted in transient increases in the total biomass of eubacteria while there were only small changes in community structure. The total gas production rate increased, while the relative methane content of the biogas and the alkalinity decreased. However, the biomass of methanogens was not affected by the glucose addition. The results show that the microbial communities of biogas processes can respond quickly to changes in the feeding rate. The glucose overload resulted in a transient general stimulation of degradation rates and almost a doubling of eubacterial biomass, although the biomass increase corresponded to only 7% of the glucose C added.  相似文献   

13.
During microbial methanogensis of diluted distillery waste or spent wash (initial COD 25,000 mg/l) in the presence of sodium acetate, sodium propionate, or sodium butyrate at the concentration of 2000 mg/l, biogas containing 22.0% to 39.4% methane was produced in 20 days in a semicontinuous fermentation system. The analysis of the volatile fatty acid spectrum of the effluent showed accumulation of 46.3% branched chain fatty acids. When the fermentation medium was supplemented with modified Smith and Mah (SM) medium containing electrolytes and 1% sodium acetate, production of methane went up to 59.0% in 18 days. With the addition of a strain of Candida sp. the coculture produced about the same volume of methane (61%) but the time required was reduced (14 days). COD (16600.8 mg/l) was reduced and the effluent contained only 3% branched chain fatty acids adn 96% straight chain fatty acids. Fortification with the sodium salts of three branched-chain fatty acids as sources of carbon in SM medium (alone) reduced methane production. Only 0.0% to 14.2% methane was recorded. By combining these acids with straight chain fatty acids in SM medium, methane production increased significantly (about 4-fold) in proportion to the concentration of straight chain fatty acids added.  相似文献   

14.
Thermophilic biohydrogen production from glucose with trickling biofilter   总被引:3,自引:0,他引:3  
Thermophilic H2 production from glucose was studied at 55-64 degrees C for 234 days using a continuous trickling biofilter reactor (TBR) packed with a fibrous support matrix. Important parameters investigated included pH, temperature, hydraulic retention time (HRT), and glucose concentration in the feed. The optimal pH and temperature were 5.5 and 60 degrees C, respectively. With decreasing HRT or increasing inlet glucose concentration, volumetric H2 production rate increased but the H2 production yield to glucose decreased gradually. The biogas composition was almost constant at 53 +/- 4% (v/v) of H2 and 47 +/- 4% (v/v) of CO2. No appreciable CH4 was detected when the reactor was under a normal operation. The carbon mass balance showed that, in addition to cell mass, lactate, n-butyrate, CO2, and acetate were major products that comprised more than 85% of the carbon consumed. The maximal volumetric H2 production rate and H2 yield to glucose were 1,050 +/- 63 mmol H2/l.d and 1.11 +/- 0.12 mol H2/mol glucose, respectively. These results indicate that the thermophilic TBR is superior to most suspended or immobilized reactor systems reported thus far. This is the first report on continuous H2 production by a thermophilic TBR system.  相似文献   

15.
Galanin, a 29 amino acid neuropeptide, was recently isolated from pig intestine. We studied the localization, nature and effect of galanin in pig pancreas. Galanin immunoreactive nerve fibers were regularly found in the pancreas. A peptide chromatographically similar to synthetic galanin was identified in pancreas extracts. The effect of galanin on the endocrine and exocrine secretion was studied in isolated pancreases, perfused with a synthetic medium containing 3.5, 5 or 8 mmol/l glucose and synthetic galanin (10(-10)-10(-8) mol/l). There was no effect on the basal exocrine secretion. The output of insulin, glucagon, somatostatin and pancreatic polypeptide (PP) was measured in the effluent. There was no effect on PP secretion. At a perfusate glucose concentration of 5 mmol/l, galanin at 10(-9) mol/l increased insulin secretion by 55 +/- 14% (mean +/- S.E.M., n = 5) of basal secretion, and at 10(-8) mol/l by 58 +/- 27% (n = 6). At 8 mmol/l glucose, insulin secretion increased by 25 +/- 10% (n = 6) and 62 +/- 17% (n = 8). At 5 mmol/l glucose glucagon secretion was increased by 15 +/- 3% (n = 5) by galanin at 10(-9) mol/l and by 29 +/- 11% (n = 5) by galanin at 10(-8) mol/l, and at 8 mmol/l glucose by 66 +/- 27% and 41 +/- 25%. Somatostatin secretion was inhibited to 72 +/- 2% (n = 5) of basal secretion by galanin at 10(-9) mol/l and to 65 +/- 7% (n = 7) at galanin at 10(-8) mol/l, both at 5 mmol/l glucose. At 8 mmol/l the figures were 83 +/- 6% and 70 +/- 10%. Insulin secretion in response to square wave increases in glucose concentration from 3.5 to 11 mmol/l (n = 5) increased 2-fold during simultaneous perfusion with galanin (10(-8) mol/l).  相似文献   

16.
The composition of volatile fatty acids in the biogas digester based on cattle manure as substrate and stabilised at 25°C showed that it contained 87–88% branched chain fatty acids, comprising of isobutyric and isovaleric acids, in comparison to 38 % observed in the digester operating at 35°C. Mixed cellulolytic cultures equilibrated at 25°C (C-25) and 35‡C (C-35) showed similar properties, but rates of hydrolysis were three times higher than that observed in a standard biogas digester. The proportion of isobutyric and isovaleric were drastically reduced when C-25 was grown with glucose or filter paper as substrates. The volatile fatty acids recovered from C-25 (at 25°C) inhibited growth of methanogens on acetate, whereas that from C-35 was not inhibitory. The inhibitory effects were due to the branched chain fatty acids and were observed with isobutyric acid at concentrations as low as 50 ppm. Addition of another micro-organismRhodotorula selected for growth on isobutyric completely reversed this inhibition. Results indicate that the aceticlastic methanogens are very sensitive to inhibition by branched chain fatty acids and reduction in methane formation in biogas digester at lower temperature may be due to this effect.  相似文献   

17.
Plasma glucose 2H enrichment was quantified by 2H NMR in patients with cirrhosis (n=6) and healthy subjects (n=5) fasted for 16 h and given 2H(2)O to approximately 0.5% body water. The percent contribution of glycogenolysis and gluconeogenesis to glucose production (GP) was estimated from the relative enrichments of hydrogen 5 and hydrogen 2 of plasma glucose. Fasting plasma glucose levels were normal in both groups (87+/-7 and 87+/-24 mg/dl for healthy and cirrhotic subjects, respectively). The percent contribution of glycogen to GP was smaller in cirrhotics than controls (22+/-7% versus 46+/-4%, P<0.001), while the contribution from gluconeogenesis was larger (78+/-7% versus 54+/-4%, P<0.001). In all subjects, glucose 6R and 6S hydrogens had similar enrichments, consistent with extensive exchange of 2H between body water and the hydrogens of gluconeogenic oxaloacetate (OAA). The difference in 2H-enrichment between hydrogen 5 and hydrogen 6S was significantly larger in cirrhotics, suggesting that the fractional contribution of glycerol to the glyceraldehyde-3-phosphate (G3P)-moiety of plasma glucose was higher compared to controls (19+/-6% versus 7+/-6%, P<0.01). In all subjects, hydrogens 4 and 5 of glucose had identical enrichments while hydrogen 3 enrichments were systematically lower. This reflects incomplete exchange between the hydrogen of water and that of 1-R-dihydroxyacetone phosphate (DHAP) or incomplete exchange of DHAP and G3P pools via triose phosphate isomerase.  相似文献   

18.
To test the hypothesis that intrahepatic availability of fatty acid could modify the rate of suppression of endogenous glucose production (EGP), acipimox or placebo was administered before and during a test meal. We used a modified isotopic methodology to measure EGP in 11 healthy subjects, and (1)H magnetic resonance spectroscopic measurement of hepatic triglyceride stores was also undertaken. Acipimox suppressed plasma free fatty acids markedly before the meal (0.05 +/- 0.01 mmol/l at -10 min, P = 0) and throughout the postprandial period (0.03 +/- 0.01 mmol/l at 150 min). Mean peak plasma glucose was significantly lower after the meal on acipimox days (8.9 +/- 0.4 vs. 10.1 +/- 0.5 mmol/l, P < 0.01), as was mean peak serum insulin (653.1 +/- 99.9 vs. 909 +/- 118 pmol/l, P < 0.01). Fasting EGP was similar (11.15 +/- 0.58 micromol.kg(-1).min(-1) placebo vs. 11.17 +/- 0.89 mg.kg(-1).min(-1) acipimox). The rate of suppression of EGP after the meal was almost identical on the 2 test days (4.36 +/- 1.52 vs. 3.69 +/- 1.21 micromol.kg(-1).min(-1) at 40 min). There was a significant negative correlation between the acipimox-induced decrease in peak plasma glucose and liver triglyceride content (r = -0.827, P = 0.002), suggesting that, when levels of liver fat were low, inhibition of lipolysis was able to affect glucose homeostasis. Acute pharmacological sequestration of fatty acids in triglyceride stores improves postprandial glucose homeostasis without effect on the immediate postprandial suppression of EGP.  相似文献   

19.
Biological hydrogen production using a membrane bioreactor   总被引:6,自引:0,他引:6  
A cross-flow membrane was coupled to a chemostat to create an anaerobic membrane bioreactor (MBR) for biological hydrogen production. The reactor was fed glucose (10,000 mg/L) and inoculated with a soil inoculum heat-treated to kill non-spore-forming methanogens. Hydrogen gas was consistently produced at a concentration of 57-60% in the headspace under all conditions. When operated in chemostat mode (no flow through the membrane) at a hydraulic retention time (HRT) of 3.3 h, 90% of the glucose was removed, producing 2200 mg/L of cells and 500 mL/h of biogas. When operated in MBR mode, the solids retention time (SRT) was increased to SRT = 12 h producing a solids concentration in the reactor of 5800 mg/L. This SRT increased the overall glucose utilization (98%), the biogas production rate (640 mL/h), and the conversion efficiency of glucose-to-hydrogen from 22% (no MBR) to 25% (based on a maximum of 4 mol-H(2)/mol-glucose). When the SRT was increased from 5 h to 48 h, glucose utilization (99%) and biomass concentrations (8,800 +/- 600 mg/L) both increased. However, the biogas production decreased (310 +/- 40 mL/h) and the glucose-to-hydrogen conversion efficiency decreased from 37 +/- 4% to 18 +/- 3%. Sustained permeate flows through the membrane were in the range of 57 to 60 L/m(2) h for three different membrane pore sizes (0.3, 0.5, and 0.8 microm). Most (93.7% to 99.3%) of the membrane resistance was due to internal fouling and the reversible cake resistance, and not the membrane itself. Regular backpulsing was essential for maintaining permeate flux through the membrane. Analysis of DNA sequences using ribosomal intergenic spacer analysis indicated bacteria were most closely related to members of Clostridiaceae and Flexibacteraceae, including Clostridium acidisoli CAC237756 (97%), Linmingia china AF481148 (97%), and Cytophaga sp. MDA2507 AF238333 (99%). No PCR amplification of 16s rRNA genes was obtained when archaea-specific primers were used.  相似文献   

20.
Placental fatty acid transfer in humans in vivo was studied using stable isotopes. Four pregnant women undergoing cesarean section received 4 h before delivery an oral dose of [(13)C]palmitic acid (PA), [(13)C]oleic acid (OA), [(13)C]linoleic acid (LA), and [(13)C]docosahexaenoic acid (DHA). Maternal blood samples were collected at -4 h (basal), -3 h, -2 h, -1 h, 0 h, and +1 h relative to time of cesarean section. At the time of birth, venous cord blood and placental tissue were collected. Fatty acid composition was determined by gas-liquid chromatography and isotopic enrichment by gas chromatography-combustion-isotope ratio mass spectrometry. (13)C-enrichment of fatty acids in the nonesterified fatty acids (NEFA) of cord plasma tended to be higher than in NEFA of placenta, with statistically significant differences for the nonesterified OA and DHA ([(13)C]PA, 0.024 +/- 0.011 vs. 0.001 +/- 0.001; [(13)C]OA, 0.042 +/- 0.008 vs. 0.005 +/- 0.003; [(13)C]LA, 0.038 +/- 0.010 vs. 0.008 +/- 0.002; [(13)C]DHA, 0.059 +/- 0.009 vs. 0.010 +/- 0.003). The ratio of tracer fatty acid concentrations of placenta to maternal plasma was significantly higher for [(13)C]DHA than for the other fatty acids ([(13)C]PA, 7.1 +/- 1%; [(13)C]OA, 3.8 +/- 0.4%; [(13)C]LA, 9.2 +/- 1.3%; [(13)C]DHA, 25.9 +/- 3.4%). These results suggest that only a part of the placental NEFA participated in fatty acid transfer, and that the placenta showed a preferential accretion of DHA relative to the other fatty acids.  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