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.
Squalene-2,3-epoxide-cycloartenol cyclase and cycloeucalenol-obtusifoliol isomerase activities were found in microsomal fractions of corn (Zea mays) embryos. Squalene-2,3-epoxide, 1-trans-1′-norsqualene-2,3-epoxide and 1-cis-1′-norsqualene-2,3-epoxide were incubated. Squalene-2,3-epoxide was cyclized giving only cycloartenol, whereas 1-trans-1′-norsqualene-2,3-epoxide gave 31-norcycloartenol and 31-norlanosterol with a reduced yield, 1-cis-1′-norsqualene-2,3-epoxide was not significantly cyclized.  相似文献   

2.
Abstract

A novel series of pyrimidine nucleoside analogues was synthesized. 2,3-Dideoxy-2,3-anhydro-β-D-lyxofuranose was opened by sodium azide to give the corresponding azido compound, which was reduced by lithium aluminium hydride to lead to 2,3-dideoxy-2,3-epimino-β-D-ribofuranose. Pyrimidine bases were glycosylated with this synthon to give potential antiviral molecules: 1-(2,3-dideoxy-2,3-epimino-β-D-ribofuranosyl)pyrimidines.  相似文献   

3.
Utilization of renewable feedstocks for the production of bio-based bulk chemicals, such as 2,3-butanediol (2,3-BDO), by engineered strains of the non-pathogenic yeast, Saccharomyces cerevisiae, has recently become an attractive option. In this study, to realize rapid production of 2,3-BDO, a flocculent, 2,3-BDO-producing S. cerevisiae strain YPH499/dPdAdG/BDN6-10/FLO1 was constructed from a previously developed 2,3-BDO-producing strain. Continuous 2,3-BDO fermentation was carried out by the flocculent strain in an airlift bioreactor. The strain consumed more than 90 g/L of glucose, which corresponded to 90% of the input, and stably produced more than 30 g/L of 2,3-BDO over 380 h. The maximum 2,3-BDO productivity was 7.64 g/L/h at a dilution rate of 0.200/h, which was higher than the values achieved by continuous fermentation using pathogenic bacteria in the previous reports. These results demonstrate that continuous 2,3-BDO fermentation with flocculent 2,3-BDO-producing S. cerevisiae is a promising strategy for practical 2,3-BDO production.  相似文献   

4.
Klebsiella oxytoca naturally produces a large amount of 2,3-butanediol (2,3-BD), a promising bulk chemical with wide industrial applications, along with various byproducts. In this study, the in silico gene knockout simulation of K. oxytoca was carried out for 2,3-BD overproduction by inhibiting the formation of byproducts. The knockouts of ldhA and pflB genes were targeted with the criteria of maximization of 2,3-BD production and minimization of byproducts formation. The constructed K. oxytoca ΔldhA ΔpflB strain showed higher 2,3-BD yields and higher final concentrations than those obtained from the wild-type and ΔldhA strains. However, the simultaneous deletion of both genes caused about a 50 % reduction in 2,3-BD productivity compared with K. oxytoca ΔldhA strain. Based on previous studies and in silico investigation that the agitation speed during 2,3-BD fermentation strongly affected cell growth and 2,3-BD synthesis, the effect of agitation speed on 2,3-BD production was investigated from 150 to 450 rpm in 5-L bioreactors containing 3-L culture media. The highest 2,3-BD productivity (2.7 g/L/h) was obtained at 450 rpm in batch fermentation. Considering the inhibition of acetoin for 2,3-BD production, fed-batch fermentations were performed using K. oxytoca ΔldhA ΔpflB strain to enhance 2,3-BD production. Altering the agitation speed from 450 to 350 rpm at nearly 10 g/L of acetoin during the fed-batch fermentation allowed for the production of 113 g/L 2,3-BD, with a yield of 0.45 g/g, and for the production of 2.1 g/L/h of 2,3-BD.  相似文献   

5.
6.
The effectiveness of some chelating agents to mobilize cadmium from Chinese hamster ovary cells after chronic exposure (20 hr), as well as from cytosolic metallothionein, was studied. In the first protocol, the most effective substance was 2,3-dimercaptopropanol, followed by 2,3-dimercaptopropane-1-sulfonate and 2,3-dimercaptosuccinic acid, whereas CaNa33-diethylenetriamine pentaacetic acid × 5H2O showed less effect. Simultaneous incubation of cells with cadmium and the chelating agent resulted in a different order of effectiveness: CaNa3 DTPA prevented cadmium uptake almost totally, 2,3-mercaptopropanol by 75% and 2,3-dimercaptopropane-1-sulfonate by 35%. Neither CaNa3-diethylenetriamine pentaacetic acid × 5H2O nor 2,3-dimercaptosuccinic acid had altered the distribution of cadmium between the cytosolic protein fractions after a 2 hr incubation of cells, whereas after this period, 2,3-dimercaptopropanol had removed all cadmium from metallothionein, and 2,3-dimercaptopropane-1-sulfonate about 50%. None of the chelating agents had reduced the amount of Cd bound to high molecular weight proteins. In the cell free system, 2,3-dimercaptopropanol and 2,3-dimercaptopropane-1-sulfonate were equally effective and removed all cadmium from metallothionein within ten minutes. CaNa3-diethylenetriamine pentaacetic acid × 5H2O, however, even after 60 min, had removed only 50% of the cadmium. The remaining cadmium was found distributed to the high molecular weight and lower molecular weight protein fractions.Abbreviations BAL 2,3-dimercaptopropanol - CHO Chinese hamster ovary cells - DMPS 2,3-dimercaptopropane-1-sulfonate - DMSA 2,3-dimercaptosuccinic acid - DTPA CaNa3-diethylenetriaminepentaacetic acid × 5 H2O - HMW proteins high molecular weight proteins - MT metallothionein  相似文献   

7.
Nonadapted freshwater sediment slurries and sediment slurries adapted to dechlorinate 2,3-dichloropyridine (2,3-Cl2Pyd), 2,3-dichloroaniline (2,3-Cl2Anl), 2,3-dichlorophenol (2,3-Cl2PhOH), 3,5-dichloropyridine (3,5-Cl2Pyd), 3,5-dichloroaniline (3,5-Cl2Anl) and 3,5-dichlorophenol (3,5-Cl2PhOH) were studied to determine the rate, range and extent of biotransformation of structurally related compounds under anaerobic conditions. 2,3-dichloroanisole (2,3-Cl2Ans) and 3,5-dichloroanisole (3,5-Cl2Ans) were initially demethylated, producing 2,3-Cl2PhOH and 3,5-Cl2PhOH as intermediate transformation products. All other dichloroaromatic compounds examined were initially dechlorinated. The rates of dechlorination of 2,3-Cl2PhOH, 2,3-Cl2Anl, and 2,3-Cl2Pyd were significantly lower (5–15 times) in nonadapted sediment slurries compared to sediment slurries adapted to 2,3-Cl2Anl or 2,3-Cl2Pyd. In 2,3-Cl2PhOH adapted sediment, the rate of dechlorination of 2,3-Cl2PhOH was 15 times greater than in nonadapted sediment; however, the rates of dechlorination of 2,3-Cl2Anl and 2,3-Cl2Pyd were similar for 2,3-Cl2PhOH-adapted and nonadapted sediment slurries. In adapted and nonadapted sediment slurries, 2,3-Cl2PhOH, 2,3-Cl2Anl, and 2,3-Cl2Pyd were preferentially dechlorinated at the ortho, meta, and meta positions, respectively. Additionally, 2,3-Cl2Pyd adapted sediment slurries dechlorinated 2,3-Cl2PhOH and 2,3-Cl2Pyd at both ortho and meta positions.Rates of dechlorination of 3,5-Cl2PhOH, 3,5-Cl2Anl, and 3,5-Cl2Pyd were lower (2–4 times) in nonadapted sediment slurries compared to sediment slurries adapted to 3,5-Cl2Anl or 3,5-Cl2Pyd. In 3,5-Cl2PhOH adapted sediment, the rate of dechlorination of 3,5-Cl2PhOH was approximately 10 times greater than in nonadapted sediment. In contrast, rates of dechlorination of 3,5-Cl2Anl and 3,5-Cl2Pyd were similar in 3,5-Cl2PhOH-adapted and nonadapted sediment slurries. A single meta chlorine was removed for all 3,5-dichloroaromatic compounds tested except 3,5-Cl2Ans, which was initially demethylated. These results illustrate differences in the specificity and cross-reactivity of microbial populations adapted to structurally related dichloroaromatic compounds.  相似文献   

8.
Abstract

1-(2,3-Dideoxy-2-C-hydroxymethyl-β-D-threo-pentofuranosyl)-, 1-(2,3-didehydro-2,3-dideoxy-2-C-hydroxymethyl-β-D-glycero-pentofuranosyl)- and 1-(2-C-azidomethyl-2,3-didehydro-2,3-dideoxy-β-D-glycero-pentofuranosyl)uracuracil, thymine and cytosine were synthesized and evaluated for their anti-HIV activities. A key step of the synthesis involves a novel alcohol transposition of2-methylene-nucleoside analogues.  相似文献   

9.

2,3-Butanediol (2,3-BDO) is of considerable importance in the chemical, plastic, pharmaceutical, cosmetic, and food industries. The main bacterial species producing this compound are considered pathogenic, hindering large-scale productivity. The species Paenibacillus brasilensis is generally recognized as safe (GRAS) and is phylogenetically similar to P. polymyxa, a species widely used for 2,3-BDO production. Here, we demonstrate, for the first time, that P. brasilensis strains produce 2,3-BDO. Total 2,3-BDO concentrations for 15 P. brasilensis strains varied from 5.5 to 7.6 g/l after 8 h incubation at 32 °C in modified YEPD medium containing 20 g/l glucose. Strain PB24 produced 8.2 g/l of 2,3-BDO within a 12-h growth period, representing a yield of 0.43 g/g and a productivity of 0.68 g/l/h. An increase in 2,3-BDO production by strain PB24 was observed using higher concentrations of glucose, reaching 27 g/l of total 2,3-BDO in YEPD containing about 80 g/l glucose within a 72-h growth period. We sequenced the genome of P. brasilensis PB24 and uncovered at least six genes related to the 2,3-BDO pathway at four distinct loci. We also compared gene sequences related to the 2,3-BDO pathway in P. brasilensis PB24 with those of other spore-forming bacteria, and found strong similarity to P. polymyxa, P. terrae, and P. peoriae 2,3-BDO-related genes. Regulatory regions upstream of these genes indicated that they are probably co-regulated. Finally, we propose a production pathway from glucose to 2,3-BDO in P. brasilensis PB24. Although the gene encoding S-2,3-butanediol dehydrogenase (butA) was found in the genome of P. brasilensis PB24, only R,R-2,3- and meso-2,3-butanediol were detected by gas chromatography under the growth conditions tested here. Our findings can serve as a basis for further improvements to the metabolic capabilities of this little-studied Paenibacillus species in relation to production of the high-value chemical 2,3-butanediol.

  相似文献   

10.
2,3-Dihydroxybiphenyl dioxygenase from Pseudomonas cepacia Et 4 was found to catalyze the ring fission of 2,3-dihydroxydiphenylether in the course of diphenylether degradation. The enzyme was purified and characterized. It had a molecular mass of 240 kDa and is dissociated by SDS into eight subunits of equal mass (31 kDa). The purified enzyme was found to be most active with 2,3-dihydroxybiphenyl as substrate and showed moderate activity with 2,3-dihydroxydiphenylether, catechol and some 3-substituted catechols. The K m-value of 1 M for 2,3-dihydroxydiphenylether indicated a high affinity of the enzyme towards this substrate. The cleavage of 2,3-dihydroxydiphenylether by 2,3-dihydroxybiphenyl dioxygenase lead to the formation of phenol and 2-pyrone-6-carboxylate as products of ring fission and ether cleavage without participation of free intermediates. Isotope labeling experiments carried out with 18O2 and H2 18O indicated the incorporation of 18O from the atmosphere into the carboxyl residue as well as into the carbonyl oxygen of the lactone moiety of 2-pyrone-6-carboxylate. Based on these experimental findings the reaction mechanism for the formation of phenol and 2-pyrone-6-carboxylate is proposed in accordance with the mechanism suggested by Kersten et al. (1982).Non-standard abbreviations DPE diphenylether - 2,3-dihydroxy-DPE 2,3-dihydroxydiphenylether - PCA 2-pyrone-6-carboxylic acid - 2,3-dihydroxy-BP dioxygenase 2,3-dihydroxybiphenyl dioxygenase - GC gas chromatography  相似文献   

11.
The present work aims to block 2,3-butanediol synthesis in acetoin fermentation of Bacillus subtilis. First, we constructed a recombinant strain BS168D by deleting the 2,3-butanediol dehydrogenase gene bdhA of the B. subtilis168, and there was almost no 2,3-butanediol production in 20?g/L of glucose media. The acetoin yield of BS168D reached 6.61?g/L, which was about 1.5 times higher than that of the control B. subtilis168 (4.47?g/L). Then, when the glucose concentration was increased to 100?g/L, the acetoin yield reached 24.6?g/L, but 2.4?g/L of 2,3-butanediol was detected at the end of fermentation. The analysis of 2,3-butanediol chiral structure indicated that the main 2,3-butanediol production of BS168D was meso-2,3-butanediol, and the bdhA gene was only responsible for (2R,3R)-2,3-butanediol synthesis. Therefore, we speculated that there may exit another pathway relating to the meso-2,3-butanediol synthesis in the B. subtilis. In addition, the results of low oxygen condition fermentation showed that deletion of bdhA gene successfully blocked the reversible transformation between acetoin and 2,3-butanediol and eliminated the effect of dissolved oxygen on the transformation.  相似文献   

12.
2,3-Diphosphoglycerate (2,3-DPG), an intracellular metabolite of glycolytic pathway is known to affect the oxygen binding capacity of haemoglobin and mechanical properties of the red blood cells. 2,3-DPG levels have been reported to be elevated during anaemic conditions including visceral leishmaniasis. 2,3-DPG activity in P. falciparum infected red blood cells, particularly in cells infected with different stages of the parasite and its relationship with structural integrity of the cells is not known. Chloroquine sensitive and resistant strains of P. falciparum were cultured in vitro and synchronized cultures of ring, trophozoite and schizont stage rich cells along with the uninfected control erythrocytes were assayed for 2,3-DPG activity and osmotic fragility. It was observed that in both the strains, in infected erythrocytes the 2,3-DPG activity gradually decreased and osmotic fragility gradually increased as the parasite matured from ring to schizont stage. The decrease in 2,3-DPG may probably be due to increased pyruvate kinase activity of parasite origin, which has been shown in erythrocytes infected with several species of Plasmodium. The absence of compensatory increase in 2,3-DPG in P. falciparum infected erythrocytes may aggravate hypoxia due to anaemia in malaria and probably may contribute to hypoxia in cerebral malaria. As 2,3-DPG was not found to be increased in erythrocytes parasitized with P. falciparum, the increased osmotic fragility observed in these cells is not due to increased 2,3-DPG as has been suggested in visceral leishmaniasis.  相似文献   

13.
A hitherto unknown amino sugar (Compound A), detected in acid hydrolyzates of lipopolysaccharides of Rhodopsuedomonas viridis and Rhodopseudomonas palustris, is present in the Lipid A component but not in the O-specific part of the lipopolysaccharides. 2-Amino-2-deoxy-D-glucose is lacking in the purified Lipid A of both strains. Compound A, characterized by a very high migration in paper electrophoresis was obtained in a pure state by ion-exchange chromatography and shown by m.s of the alditol acetate to be a 2,3-diamino-2,3-dideoxyhexose. G.I.c. and periodate oxidation excluded all possible stereoisomers with the exception of 2,3-diamino-2,3-dideoxyglucose and 2,3-diamino-2,3-dideoxyidose. G.I.c. of the alditol acetates of Compound A and of the glucose derivative suggests that Compound A is 2,3-diamino-2,3-dideoxyglucose. The significance of the occurrence of this new aminodeoxy sugar in the lipid A component of Rhodopsuedomonas viridis and Rhodopseudomonas palustris O-antigens for the biological properties of the respective lipopolysaccharides and for the taxonomy of the Rhodospirillaceae family is discussed.  相似文献   

14.
John G. Bruno 《Luminescence》1998,13(3):139-145
Electrochemiluminescence (ECL) of 200 ppm 2,3-diaminonaphthalene (2,3-DAN) was studied alone and in conjunction with 100 ppm of 34 different metal and non-metal ions and revealed three relatively intense ECL responses from interactions of 2,3-DAN with Au+, Fe+3 and V+5. ECL responses from Cr+6 or Ru+3 with 2,3-DAN were less intense, but noteworthy, as was the coloured fluorescent product of the non-metal ion Se+4 interaction with 2,3-DAN. Several intense 2,3-DAN–metal ion ECL reactions were studied in greater detail and revealed various titration curves with ionic detection limits in the low ppm range, using a fixed level (200 ppm) of 2,3-DAN. © 1998 John Wiley & Sons, Ltd.  相似文献   

15.
Enzymatic reduction of acetoin into 2,3-butanediol (2,3-BD) typically requires the reduced nicotinamide adenine dinucleotide (NADH) or its phosphate form (NADPH) as electron donor. Efficiency of 2,3-BD biosynthesis, therefore, is heavily influenced by the enzyme specificity and the cofactor availability which varies dynamically. This work describes the engineering of cofactor flexibility for 2,3-BD production by simultaneous overexpression of an NADH-dependent 2,3-BD dehydrogenase from Klebsiella pneumoniae (KpBudC) and an NADPH-specific 2,3-BD dehydrogenase from Clostridium beijerinckii (CbAdh). Co-expression of KpBudC and CbAdh not only enabled condition versatility for 2,3-BD synthesis via flexible utilization of cofactors, but also improved production stereo-specificity of 2,3-BD without accumulation of acetoin. With optimization of medium and fermentation condition, the co-expression strain produced 92 g/L of 2,3-BD in 56 h with 90% stereo-purity for (R,R)-isoform and 85% of maximum theoretical yield. Incorporating cofactor flexibility into the design principle should benefit production of bio-based chemical involving redox reactions.  相似文献   

16.
2,3-Butanediol containing less than 3% of themesoform has been obtained from samples containing up to 50% of themesoform. The diacetate was obtained by esterification with acetic anhydride in the presence of traces of sulfuric acid as a catalyst and was then purified. When the diacetate was held at 4°C, crystals of racemic 2,3-butanediol diacetate formed, and these were separated by filtration. The diacetate was then transformed back to 2,3-butanediol by transesterification with methanol in the presence of sodium methylate as a catalyst. The resulting 2,3-butanediol contained less than 3% of themesoform. For an original batch of 2,3-butanediol containing 50%dland 50%meso,this method can isolate up to 70% of the racemate content. If the original 2,3-butanediol contains too muchmesoform, racemic 2,3-butanediol diacetate does not crystallize, but 2,3-butanediol containing up to 60% of themesoform can be enriched up to 70% racemate by distillation.  相似文献   

17.
A 2,3-dihydroxybiphenyl (2,3-DHBP) dioxygenase gene from a Rhodococcus sp. strain, named RrbphCI and involved in the degradation of polychlorinated biphenyls (PCBs), was synthesized. RrbphCI was expressed in Escherichia coli and its encoded enzyme was purified. SDS–PAGE analysis indicated that the size of the protein encoded by RrbphCI was about 32 kDa. The activity of the 2,3-DHBP dioxygenase was 82.8 U/mg when the substrate was 2,3-DHBP, with optimum pH 8.0 at 30°C, and optimum temperature was 40°C at pH 8.0. The RrbphCI gene was transformed into Pseudomonas putida strain EG11, to determine the ability of the enzyme to degrade 2,3-DHBP. The wild type EG11 degraded 61.86% of supplied 2,3-DHBP and the transformed EG11 (hosting the RrbphCI gene) utilized 52.68% after 2 min of treatment at 30°C. The overexpressed and purified enzyme was able to degrade 2,3-DHBP. The 2,3-DHBP dioxygenase is a key enzyme in the PCB degradation pathway. RrbphCI and its encoded 2,3-DHBP dioxygenase may have transgenic applications in bioremediation of PCBs.  相似文献   

18.
The microbial production of high amounts of 2,3-butanediol (2,3-BD) from glycerol as a sole carbon source by the Bulgarian isolate Klebsiella pneumoniae G31 was studied in a series of fed-batch processes. The following conditions were evaluated as optimal: micro-aerobic cultivation in modified media, without pH control. Beginning at pH 8, 49.2 g/l of 2,3-BD was produced as negligible concentrations of by-products were received. The pH is the most important factor ruling the 2,3-BD production. Spontaneous pH changes and products formation in time were investigated, performing fermentations with non-controlled pH starting at different initial pH. In lack of external maintenance, the microorganism attempted to control the pH using acetate/2,3-BD alternations of the oxidative pathway of glycerol catabolism, which resulted in pH fluctuations. Thus, the culture secreted 2,3-BD at unequal portions, either allowing or detaining the acetate synthesis. More alkaline initial pH led to enhanced 2,3-BD accumulation as a response to the increased amplitudes of the pH variations. When the pH was maintained constant, the yield of 2,3-BD was very poor. These cultures remained viable only 72 h; whereas, the pH self-controlling cells lived and produced 2,3-BD up to 280 h. In conclusion, the formation of 2,3-BD is a result of an adaptive mechanism of pH self-control, responding to spontaneous pH drops during glycerol fermentation.  相似文献   

19.
Cultivation in glycerol instead of sugars inhibits 2,3-butanediol (2,3-BD) production by Bacillus amyloliquefaciens. In this study, we report that B. amyloliquefaciens readily produces 2,3-BD from biodiesel-derived glycerol in the presence of beet molasses as a co-substrate. Unexpectedly, the molasses stimulated 2,3-BD production and simultaneously reduced the duration of fermentation. Productivity of 2,3-BD was enhanced at the start of fermentation, and yields increased under continuous molasses supply. Subsequently, 2,3-BD production in molasses-supplemented fed-batch culture was observed. Prior to inoculation of fed-batch fermentation culture, 15 g/l of molasses was added to the bioreactor. After 6 h of incubation, the bioreactor was fed with a solution containing 80 % glycerol and 15 % molasses. The 2,3-BD concentration, yield, and productivity significantly improved, reaching 83.3 g/l, 0.42 g/g, and 0.87 g/l·h, respectively. To our knowledge, these results are the highest report for 2,3-BD fermentation from biodiesel-derived glycerol.  相似文献   

20.
Ethanol was a major byproduct of 2,3-butanediol (2,3-BD) fermentation by Klebsiella oxytoca ME-UD-3. In order to achieve a high efficiency of 2,3-BD production, K. oxytoca mutants deficient in ethanol formation were successfully constructed by replace the aldA gene coding for aldehyde dehydrogenase with a tetracycline resistance cassette. The results suggested that inactivation of aldA led to a significantly improved 2,3-BD production. The carbon flux to 2,3-BD was enhanced by eliminating the byproducing ethanol and at the same time reducing the accumulation of another byproduct acetoin. At last, by fed-batch culturing of the mutant, the final 2,3-BD titer up to 130 g/l with the productivity of 1.63 g/l.h and the 2,3-BD yield relative to glucose of 0.48 g/g was obtained.  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