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.
To enrich syntrophic acetate‐oxidizing bacteria (SAOB), duplicate chemostats were inoculated with sludge from syntrophic acetate oxidation (SAO)‐dominated systems and continuously supplied with acetate (0.4 or 7.5 g l?1) at high‐ammonia levels. The chemostats were operated under mesophilic (37°C) or thermophilic (52°C) temperature for about six hydraulic retention times (HRT 28 days) and were sampled over time. Irrespective of temperature, a methane content of 64–69% and effluent acetate level of 0.4–1.0 g l?1 were recorded in chemostats fed high acetate. Low methane production in the low‐acetate chemostats indicated that the substrate supply was below the threshold for methanization of acetate via SAO. Novel representatives within the family Clostridiales and genus Syntrophaceticus (class Clostridia) were identified to represent putative SAOB candidates in mesophilic and thermophilic conditions respectively. Known SAOB persisted at low relative abundance in all chemostats. The hydrogenotrophic methanogens Methanoculleus bourgensis (mesophilic) and Methanothermobacter thermautotrophicus (thermophilic) dominated archaeal communities in the high‐acetate chemostats. In line with the restricted methane production in the low‐acetate chemostats, methanogens persisted at considerably lower abundance in these chemostats. These findings strongly indicate involvement in SAO and tolerance to high ammonia levels of the species identified here, and have implications for understanding community function in stressed anaerobic processes.  相似文献   

2.
Coal mining is responsible for 11% of total anthropogenic methane emission thereby contributing considerably to climate change. Attempts to harvest coalbed methane for energy production are challenged by relatively low methane concentrations. In this study, we investigated whether nutrient and acetate amendment of a non‐producing sub‐bituminous coal well could transform the system to a methane source. We tracked cell counts, methane production, acetate concentration and geochemical parameters for 25 months in one amended and one unamended coal well in Australia. Additionally, the microbial community was analysed with 16S rRNA gene amplicon sequencing at 17 and 25 months after amendment and complemented by metagenome sequencing at 25 months. We found that cell numbers increased rapidly from 3.0 × 104 cells ml?1 to 9.9 × 107 in the first 7 months after amendment. However, acetate depletion with concomitant methane production started only after 12–19 months. The microbial community was dominated by complex organic compound degraders (Anaerolineaceae, Rhodocyclaceae and Geobacter spp.), acetoclastic methanogens (Methanothrix spp.) and fungi (Agaricomycetes). Even though the microbial community had the functional potential to convert coal to methane, we observed no indication that coal was actually converted within the time frame of the study. Our results suggest that even though nutrient and acetate amendment stimulated relevant microbial species, it is not a sustainable way to transform non‐producing coal wells into bioenergy factories.  相似文献   

3.
Hydrogen and methane were simultaneously produced in a two‐phase reactor, operated to separate the reactions of hydrogen and methanogen production. Each reactor was inoculated with a seed enriched with different microbial consortia. The first phase was operated with a hydraulic retention time of 7 days and at an organic loading rate of 7.7 g VS L?1 d?1 that produced a stable pH of 5.5. This suppressed the growth of methanogens and as a result, the off gas contained up to 27% hydrogen. The second phase was operated with a hydraulic retention time of 12 days and at an organic loading rate of 3.6 g VS L?1 d?1. This permitted the growth of hydrogenotrophs and methanogens to produce methane at a concentration of 60%. Examination of the microbial population of the two reactors both microscopically and using PCR, showed an effective separation of hydrogen‐ and methane‐producing microbial communities. The study revealed that the suppression of hydrogentrophs and methanogens can be achieved by adopting rapid method that leads the growth of hydrogen‐ and methane‐producing granules in phase‐separated anaerobic environment.  相似文献   

4.
We demonstrate that the coulombic efficiency (CE) of a microbial electrolytic cell (MEC) fueled with a fermentable substrate, ethanol, depended on the interactions among anode respiring bacteria (ARB) and other groups of micro‐organisms, particularly fermenters and methanogens. When we allowed methanogenesis, we obtained a CE of 60%, and 26% of the electrons were lost as methane. The only methanogenic genus detected by quantitative real‐time PCR was the hydrogenotrophic genus, Methanobacteriales, which presumably consumed all the hydrogen produced during ethanol fermentation (~30% of total electrons). We did not detect acetoclastic methanogenic genera, indicating that acetate‐oxidizing ARB out‐competed acetoclastic methanogens. Current production and methane formation increased in parallel, suggesting a syntrophic interaction between methanogens and acetate‐consuming ARB. When we inhibited methanogenesis with 50 mM 2‐bromoethane sulfonic acid (BES), the CE increased to 84%, and methane was not produced. With no methanogenesis, the electrons from hydrogen were converted to electrical current, either directly by the ARB or channeled to acetate through homo‐acetogenesis. This illustrates the key role of competition among the various H2 scavengers and that, when the hydrogen‐consuming methanogens were present, they out‐competed the other groups. These findings also demonstrate the importance of a three‐way syntrophic relationship among fermenters, acetate‐consuming ARB, and a H2 consumer during the utilization of a fermentable substrate. To obtain high coulombic efficiencies with fermentable substrates in a mixed population, methanogens must be suppressed to promote new interactions at the anode that ultimately channel the electrons from hydrogen to current. Biotechnol. Bioeng. 2009;103: 513–523. © 2009 Wiley Periodicals, Inc.  相似文献   

5.
The conversion routes of carbon monoxide (CO) at 55°C by full-scale grown anaerobic sludges treating paper mill and distillery wastewater were elucidated. Inhibition experiments with 2-bromoethanesulfonate (BES) and vancomycin showed that CO conversion was performed by a hydrogenogenic population and that its products, i.e. hydrogen and CO2, were subsequently used by methanogens, homo-acetogens or sulfate reducers depending on the sludge source and inhibitors supplied. Direct methanogenic CO conversion occurred only at low CO concentrations [partial pressure of CO (P CO) <0.5 bar (1 bar=105 Pa)] with the paper mill sludge. The presence of hydrogen decreased the CO conversion rates, but did not prevent the depletion of CO to undetectable levels (<400 ppm). Both sludges showed interesting potential for hydrogen production from CO, especially since after 30 min exposure to 95°C, the production of CH4 at 55°C was negligible. The paper mill sludge was capable of sulfate reduction with hydrogen, tolerating and using high CO concentrations (P CO>1.6 bar), indicating that CO-rich synthesis gas can be used efficiently as an electron donor for biological sulfate reduction.  相似文献   

6.
The short-term effects of temperature on methanogenesis from acetate or CO2 in a thermophilic (58°C) anaerobic digestor were studied by incubating digestor sludge at different temperatures with 14C-labeled methane precursors (14CH3COO or 14CO2). During a period when Methanosarcina sp. was numerous in the sludge, methanogenesis from acetate was optimal at 55 to 60°C and was completely inhibited at 65°C. A Methanosarcina culture isolated from the digestor grew optimally on acetate at 55 to 58°C and did not grow or produce methane at 65°C. An accidental shift of digestor temperature from 58 to 64°C during this period caused a sharp decrease in gas production and a large increase in acetate concentration within 24 h, indicating that the aceticlastic methanogens in the digestor were the population most susceptible to this temperature increase. During a later period when Methanothrix sp. was numerous in the digestor, methanogenesis from 14CH3COO was optimal at 65°C and completely inhibited at 75°C. A partially purified Methanothrix enrichment culture derived from the digestor had a maximum growth temperature near 70°C. Methanogenesis from 14CO2 in the sludge was optimal at 65°C and still proceeded at 75°C. A CO2-reducing Methanobacterium sp. isolated from the digestor was capable of methanogenesis at 75°C. During the period when Methanothix sp. was apparently dominant, sludge incubated for 24 h at 65°C produced more methane than sludge incubated at 60°C, and no acetate accumulated at 65°C. Methanogenesis was severely inhibited in sludge incubated at 70°C, but since neither acetate nor H2 accumulated, production of these methanogenic substrates by fermentative bacteria was probably the most temperature-sensitive process. Thus, there was a correlation between digestor performance at different temperatures and responses to temperature by cultures of methanogens believed to play important roles in the digestor.  相似文献   

7.
The flow of carbon under sulfate‐reducing conditions within a benzene‐mineralizing enrichment culture was analysed using fully labelled [13C6]‐benzene. Over 180 days of incubation, 95% of added 13C‐benzene was released as 13C‐carbon dioxide. DNA extracted from cultures that had degraded different amounts of unlabelled or 13C‐labelled benzene was centrifuged in CsCl density gradients to identify 13C‐benzene‐assimilating organisms by density‐resolved terminal restriction fragment length polymorphism analysis and cloning of 16S rRNA gene fragments. Two phylotypes showed significantly increased relative abundance of their terminal restriction fragments in ‘heavy’ fractions of 13C‐benzene‐incubated microcosms compared with a 12C‐benzene‐incubated control: a member of the Cryptanaerobacter/Pelotomaculum group within the Peptococcaceae, and a phylotype belonging to the Epsilonproteobacteria. The Cryptanaerobacter/Pelotomaculum phylotype was the most frequent sequence type. A small amount of 13C‐methane was aceticlastically produced, as concluded from the linear relationship between methane production and benzene degradation and the detection of Methanosaetaceae as the only methanogens present. Other phylotypes detected but not 13C‐labelled belong to several genera of sulfate‐reducing bacteria, that may act as hydrogen scavengers for benzene oxidation. Our results strongly support the hypothesis that benzene is mineralized by a consortium consisting of syntrophs, hydrogenotrophic sulfate reducers and to a minor extent of aceticlastic methanogens.  相似文献   

8.
Butyrate is an important intermediate in the anaerobic degradation of organic matter. In sulfate-depleted environments butyrate is oxidized to acetate and hydrogen by obligate proton reducers, in syntrophic association with hydrogen-consuming methanogens. This paper describes two enrichments of endospore-forming bacteria degrading butyrate in consortia with methanogens. The isolates are readily established in coculture with H2-consuming, sulfate-reducing bacteria by pasteurizing the culture. The two original enrichments differed in that one grew to an optically dense culture while the second grew in clumps. Examination by scanning electron microscopy showed that clumping resulted from the production of large amounts of extracellular polymer. Several H2-consuming methanogens were identified in the enrichments. Some of them grew closely associated to the butyrate degraders. This attachment to the hydrogen producer may permit some methanogens to compete for the growth substrate against other bacteria having higher substrate affinity.  相似文献   

9.
10.
Anaerobic digestion (AD) processes are known to effectively convert organic waste to CO2 and CH4, but much of the microbial ecology remains unclear. Specifically, we have limited insights into symbiotic syntroph and methanogen (‘syntrophy’) acid degradation, although they are essential for preventing process deterioration. Also, we often observed many uncharacterized or uncultivated organisms, but poorly understood their role(s) in relation to syntrophy. To define syntrophy‐associated populations, this study enriched methanogenic communities with propionate, butyrate, benzoate, acetate, formate and H2 from two different inocula over 3 years. 16S pyrotag analysis revealed core populations of known syntrophs (six clades) and methanogens (nine clades) associated with acid degradation, and evidence for substrate‐ and/or inoculum‐dependent specificity in syntrophic partnerships. Based on comprehensive re‐evaluation of publically available microbial community data for AD, the known syntrophs and methanogens identified were clearly representatives of the AD‐associated syntrophs and methanogens. In addition, uncultivated clades related to Bacteroidetes, Firmicutes, Actinobacteria and Chloroflexi were ubiquitously found in AD and enrichments. These organisms may be universally involved in AD syntrophic degradation, but only represented <23% of the yet‐to‐be‐cultivated organisms (89 of 390 clades). Thus, the contribution of these uncultured organisms in AD remains unclear and warrants further investigation.  相似文献   

11.
Methanogenesis in thermophilic biogas reactors   总被引:2,自引:0,他引:2  
Methanogenesis in thermophilic biogas reactors fed with different wastes is examined. The specific methanogenic activity with acetate or hydrogen as substrate reflected the organic loading of the specific reactor examined. Increasing the loading of thermophilic reactors stabilized the process as indicated by a lower concentration of volatile fatty acids in the effluent from the reactors. The specific methanogenic activity in a thermophilic pilot-plant biogas reactor fed with a mixture of cow and pig manure reflected the stability of the reactor. The numbers of methanogens counted by the most probable number (MPN) technique with acetate or hydrogen as substrate were further found to vary depending on the loading rate and the stability of the reactor. The numbers of methanogens counted with antibody probes in one of the reactor samples was 10 times lower for the hydrogen-utilizing methanogens compared to the counts using the MPN technique, indicating that other non-reacting methanogens were present. Methanogens that reacted with the probe againstMethanobacterium thermoautotrophicum were the most numerous in this reactor. For the acetate-utilizing methanogens, the numbers counted with the antibody probes were more than a factor of 10 higher than the numbers found by MPN. The majority of acetate utilizing methanogens in the reactor wereMethanosarcina spp. single cells, which is a difficult form of the organism to cultivatein vitro. No reactions were observed with antibody probes raised againstMethanothrix soehngenii orMethanothrix CALS-1 in any of the thermophilic biogas reactors examined. Studies using 2-14C-labeled acetate showed that at high concentrations (more than approx. 1 mM) acetate was metabolized via the aceticlastic pathway, transforming the methyl-group of acetate into methane. When the concentration of acetate was less than approx. 1 mM, most of the acetate was oxidized via a two-step mechanism (syntrophic acetate oxidation) involving one organism oxidizing acetate into hydrogen and carbon dioxide and a hydrogen-utilizing methanogen forming the products of the first microorganism into methane. In thermophilic biogas reactors, acetate oxidizing cultures occupied the niche ofMethanothrix species, aceticlastic methanogens which dominate at low acetate concentrations in mesophilic systems. Normally, thermophilic biogas reactors are operated at temperatures from 52 to 56° C. Experiments using biogas reactors fed with cow manure showed that the same biogas yield found at 55° C could be obtained at 61° C after a long adaptation period. However, propionate degradation was inhibited by increasing the temperature.  相似文献   

12.
The cotton whitefly, Bemisia tabaci (Gennadius) B‐biotype, is fed on by a wide variety of generalist predators, but there is little information on these predator–prey interactions, especially under field conditions. In this study, a real‐time polymerase chain reaction (PCR) assay was developed to quantify B. tabaci B‐biotype remains in predator gut. The B. tabaci B‐biotype genomic DNA copy number was referred to the actual amount of BT1 isolate, the B. tabaci B‐biotype specific DNA fragment. The numbers of BT1 isolate in one B. tabaci B‐biotype egg, individual adult and a single red‐eyed nymph were 2.56 × 103, 2.56 × 104, and 1.29 × 104 copies, respectively. When Propylaea japonica adults fed on one, two, four, eight or 16 red‐eyed nymphs, the detected numbers of BT1 isolate ranged from 2.77 × 104 to 4.05 × 105 copies, forming a strong linear relationship (R2 = 0.9899). Following the consumption of two red‐eyed nymphs, prey DNA was detectable in 100% of P. japonica at t = 0, decreasing to 80.0% and 60.0% after 1–4 h and 8 h of digestion, respectively, with 3.36 × 104–1.25 × 103 BT1 isolate copies. The predation by field‐collected predators, 26 larvae of P. japonica, and of Harmonia axyridis each, Chrysopa spp. larvae (Chrysopa pallens and C. formosa, 18 individuals in total), and a single adult of Scymnus hoffmanni, 19 adults of Orius sauteri and nine adult spiders (Erigonnidium graminicolum and Neoscona doenitzi), on B. tabaci B‐biotype were quantified. Of the 99 analysed predator individuals, 3.65 × 102–4.60 × 105 copies of BT1 isolate, equivalent to 0.8–18.8 red‐eyed nymphs were detected. These results suggest that TaqMan real‐time PCR technology may provide a rapid and sensitive method for quantifying B. tabaci B‐biotype remains in predator guts and will be invaluable in assessing the food web relationship between prey and arthropod predators.  相似文献   

13.
CpG‐related single nucleotide polymorphisms (CGS) have the potential to perturb DNA methylation; however, their effects on Alzheimer disease (AD) risk have not been evaluated systematically. We conducted a genome‐wide association study using a sliding‐window approach to measure the combined effects of CGSes on AD risk in a discovery sample of 24 European ancestry cohorts (12,181 cases, 12,601 controls) from the Alzheimer's Disease Genetics Consortium (ADGC) and replication sample of seven European ancestry cohorts (7,554 cases, 27,382 controls) from the International Genomics of Alzheimer's Project (IGAP). The potential functional relevance of significant associations was evaluated by analysis of methylation and expression levels in brain tissue of the Religious Orders Study and the Rush Memory and Aging Project (ROSMAP), and in whole blood of Framingham Heart Study participants (FHS). Genome‐wide significant (p < 5 × 10?8) associations were identified with 171 1.0 kb‐length windows spanning 932 kb in the APOE region (top p < 2.2 × 10?308), five windows at BIN1 (top p = 1.3 × 10?13), two windows at MS4A6A (top p = 2.7 × 10?10), two windows near MS4A4A (top p = 6.4 × 10?10), and one window at PICALM (p = 6.3 × 10‐9). The total number of CGS‐derived CpG dinucleotides in the window near MS4A4A was associated with AD risk (p = 2.67 × 10?10), brain DNA methylation (p = 2.15 × 10?10), and gene expression in brain (p = 0.03) and blood (p = 2.53 × 10?4). Pathway analysis of the genes responsive to changes in the methylation quantitative trait locus signal at MS4A4A (cg14750746) showed an enrichment of methyltransferase functions. We confirm the importance of CGS in AD and the potential for creating a functional CpG dosage‐derived genetic score to predict AD risk.  相似文献   

14.
We established Fe(III)‐reducing co‐cultures of two species of metal‐reducing bacteria, the Gram‐positive Desulfotomaculum reducens MI‐1 and the Gram‐negative Geobacter sulfurreducens PCA. Co‐cultures were given pyruvate, a substrate that D. reducens can ferment and use as electron donor for Fe(III) reduction. G. sulfurreducens relied upon products of pyruvate oxidation by D. reducens (acetate, hydrogen) for use as electron donor in the co‐culture. Co‐cultures reduced Fe(III) to Fe(II) robustly, and Fe(II) was consistently detected earlier in co‐cultures than pure cultures. Notably, faster cell growth, and correspondingly faster pyruvate oxidation, was observed by D. reducens in co‐cultures. Global comparative proteomic analysis was performed to observe differential protein abundance during co‐culture vs. pure culture growth. Proteins previously associated with Fe(III) reduction in G. sulfurreducens, namely c‐type cytochromes and type IV pili proteins, were significantly increased in abundance in co‐cultures relative to pure cultures. D. reducens ribosomal proteins were significantly increased in co‐cultures, likely a reflection of faster growth rates observed for D. reducens cells while in co‐culture. Furthermore, we developed multiple reaction monitoring (MRM) assays to quantitate specific biomarker peptides. The assays were validated in pure and co‐cultures, and protein abundance ratios from targeted MRM and global proteomic analysis correlate significantly.  相似文献   

15.
An archaeal ether-linked lipid, archaeol, was determined to be a biomass indicator for methanogens both in the laboratory enriched culture and in marine sediments. The archaeol measurement method described by Ohtsubo et al. in 1993 was modified and applied to marine sediments. We compared the amount of archaeol with the cell number of methanogens or methane concentration in laboratory enriched culture of methanogens from marine sediment. Good correlations were obtained as follows: (Methane, mmol) = 11.2 × (Archaeol, mg): r= .996 or (Cell number) = 1.13 × 1011× (Archaeol, mg): r= .995, respectively. In the sediments of Tokyo Bay, archaeol was measured from approximately 46 to 561 ng/dry g sediment at the entrance to 267 to 4160 ng/dry g sediment at the innermost area. Using the coefficient from the laboratory experiment, these data corresponded to cell numbers of 5.2 × 106 to 4.7 × 108/dry g sediment. These values were 1 or 2 orders of magnitude higher than those obtained by culture methods in previous studies. Although dead or decomposed cells might be detected, archaeol measurement is useful for estimating the biomass of methanogens because of the good correlation between methane concentration and archaeol content in marine environments. In this study, we found a correlation of (Methane, mmol) = 0.012 × (Archaeol, mg): r= .932, n= 17 in marine sediments. Received December 21, 1998; accepted June 16, 1999  相似文献   

16.
Drug‐protein interaction analysis is pregnant in designing new leads during drug discovery. We prepared the stationary phase containing immobilized β2‐adrenoceptor (β 2AR) by linkage of the receptor on macroporous silica gel surface through N ,N ′‐carbonyldiimidazole method. The stationary phase was applied in identifying antiasthmatic target of protopine guided by the prediction of site‐directed molecular docking. Subsequent application of immobilized β 2AR in exploring the binding of protopine to the receptor was realized by frontal analysis and injection amount–dependent method. The association constants of protopine to β 2AR by the 2 methods were (1.00 ± 0.06) × 105M−1 and (1.52 ± 0.14) × 104M−1. The numbers of binding sites were (1.23 ± 0.07) × 10−7M and (9.09 ± 0.06) × 10−7M, respectively. These results indicated that β 2AR is the specific target for therapeutic action of protopine in vivo. The target‐drug binding occurred on Ser169 in crystal structure of the receptor. Compared with frontal analysis, injection amount–dependent method is advantageous to drug saving, improvement of sampling efficiency, and performing speed. It has grave potential in high‐throughput drug‐receptor interaction analysis.  相似文献   

17.
Previous studies suggested that methanol and acetate were the likely methanogenic precursors in the cold Zoige wetland. In this study, the contribution of the two substances to methanogenesis and the conversion in Zoige wetland were analyzed. It was determined that methanol supported the highest CH4 formation rate in the enrichments of the soil grown with Eleocharis valleculosa, and even higher at 15°C than at 30°C; while hydrogenotrophic methanogenesis was higher at 30°C. Both methanol- and acetate-using methanogens were counted at the highest (107 g−1) in the soil, whereas methanol-using acetogens (108 g−1) were ten times more abundant than either methanol- or acetate-using methanogens. Both methanol and acetate were detected in the methanogenesis-inhibited soil samples, so that both could be the primary methanogenic precursors in E. valleculosa soil. However, the levels of methanol and acetate accumulated in 2-bromoethane-sulfonate (BES)- and CHCl3-treated soils were in reverse, i.e., higher methanol in CHCl3- and higher acetate in BES-treated soil, so that methanol-derived methanogenesis could be underestimated due to the consumption by acetogens. Analysis of the soil 16S rRNA genes revealed Acetobacterum bakii and Trichococcus pasteurii to be the dominant methanol-using acetogens in the soil, and a strain of T. pasteurii was isolated, which showed the high conversion of methanol to acetate at 15°C.  相似文献   

18.
The suitability of miscanthus biomass for anaerobic digestion has already been confirmed by several studies. However, it is rarely used as feedstock in biogas plants, mainly due to uncertainty about the optimal harvest regime with regard to the long‐term methane hectare yield and resilience of the crop to green cutting. The recommended green‐cut date for the only commercially available genotype Miscanthus × giganteus (M×g) ranges from September to November. This timeframe is too broad for agricultural practice and needs to be both narrowed down and further specified for different genotypes. The aim of this study was to identify the most suitable harvest window for an autumn green cut of miscanthus, which delivers both a high dry matter and methane yield while securing the long‐term productivity of the crop. A further objective was to quantify the effect of genotypic differences, such as leaf to stem ratio, on the substrate‐specific biogas and methane yield. For these purposes, a field trial with four genotypes (M×g, GNT1, GNT3, Sin55) was conducted over 2 years (2016/2017) and harvested at 2‐week intervals on three dates between mid‐September to mid‐October. Methane hectare yield ranged from 3,183 m³ CH4 ha?1 a?1 (Sin55) to 5,265 m³ CH4 ha?1 a?1 (M×g), which is mainly influenced by dry matter yield. The substrate‐specific methane yield was higher for the leaf (311.0 ml CH4 (g oDM)‐1) than the stem fraction (285.1 ml CH4 (g oDM)‐1) in all genotypes due to lower lignin content of leaves. Of all genotypes, M×g showed the highest and Sin55 the lowest nutrient use efficiency. We conclude that miscanthus in Germany should be harvested in October to maximize methane yields and nutrient recycling and minimize yield reduction. Additionally, to increase methane hectare yields even further, future miscanthus breeding should focus on a higher leaf proportion.  相似文献   

19.
A simple, rapid chemiluminescence (CL) method was described for the determination of piroxicam, a commonly used analgesic agent drug. A strong CL signal was detected when cerium(IV) sulphate was injected into tris‐(4,7‐diphenyl‐1,10‐phenanthrolinedisulphonic acid) ruthenium(II) (RuBPS)–piroxicam solution. The CL signal was proportional to the concentration of piroxicam in the range 2.8 × 10–8–1.2 × 10–5 mol/L. The detection limit was 2 × 10–8 mol/L and the relative standard deviation (RSD) was 3.7% (c = 7.0 × 10–7 mol/L piroxicam; n = 11). The proposed method was applied to the determination of piroxicam in pharmaceutical preparations in capsules, spiked serum and urine samples with satisfactory results. Copyright © 2008 John Wiley & Sons, Ltd.  相似文献   

20.
The objective of this study was to improve the biological water–gas shift reaction for producing hydrogen (H2) by conversion of carbon monoxide (CO) using an anaerobic thermophilic pure strain, Carboxydothermus hydrogenoformans. Specific hydrogen production rates and yields were investigated at initial biomass densities varying from 5 to 20 mg volatile suspended solid (VSS) L−1. Results showed that the gas–liquid mass transfer limits the CO conversion rate at high biomass concentrations. At 100-rpm agitation and at CO partial pressure of 1 atm, the optimal substrate/biomass ratio must exceed 5 mol CO g−1 biomass VSS in order to avoid gas–liquid substrate transfer limitation. An average H2 yield of 94 ± 3% and a specific hydrogen production rate of ca. 3 mol g−1 VSS day−1 were obtained at initial biomass densities between 5 and 8 mg VSS−1. In addition, CO bioconversion kinetics was assessed at CO partial pressure from 0.16 to 2 atm, corresponding to a dissolved CO concentration at 70°C from 0.09 to 1.1 mM. Specific bioactivity was maximal at 3.5 mol CO g−1 VSS day−1 for a dissolved CO concentration of 0.55 mM in the culture. This optimal concentration is higher than with most other hydrogenogenic carboxydotrophic species.  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