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.
Bacterial 2,3-butanediol dehydrogenases   总被引:3,自引:0,他引:3  
Enterobacter aerogenes, Aeromonas hydrophila, Serratia marcescens and Staphylococcus aureus possessing L(+)-butanediol dehydrogenase produced mainly meso-butanediol and small amounts of optically active butanediol; Acetobacter suboxydans, Bacillus polymyxa and Erwinia carotovora containing D(-)-butanediol dehydrogenase produced more optically active butanediol than meso-butanediol. Resting and growing cells of these organisms oxidized only one enantiomer of racemic butanediol. The D(-)-butanediol dehydrogenase from Bacillus polymyxa was partially purified (30-fold) with a specific activity of 24.5. Except NAD and NADH no other cofactors were required. Optimum pH-values for oxidation and reduction were pH 9 and pH 7, respectively. The optimum temperature was about 60°C. The molecular weight was 100000 to 107000. The K m-values were 3.3 mM for D(-)-butanediol, 6.25 mM for meso-butanediol, 0.53 mM for acetoin, 0.2 mM for NAD, 0.1 mM for NADH, 87 mM for diacetyl, 38 mM for 1,2-propanediol; 2,3-pentanedion was not a substrate for this enzyme. The L(+)-butanediol dehydrogenase from Serratia marcescens was purified 57-fold (specific activity 22.3). Besides NAD or NADH no cofactors were required. The optimum value for oxidation was about pH 9 and for reduction pH 4.5. The optimum temperature was 32–36°C. The molecular weight was 100000 to 107000. The K m-values were 5 mM for meso-butanediol, 10 mM for racemic butanediol, 6.45 for acetoin, 1 mM for NAD, 0.25 mM for NADH, 2.08 mM for diacetyl, 16.7 mM for 2,3-pentanedion and 11.8 mM for 1,2-propanediol.Abbreviations Bud 2,3-butanediol - DH dehydrogenase  相似文献   

2.
We utilized a photoautotrophic organism to synthesize 1,2‐propanediol from carbon dioxide and water fueled by light. A synthetic pathway comprising mgsA (methylglyoxal synthase), yqhD (aldehyde reductase), and adh (alcohol dehydrogenase) was inserted into Synechocystis sp. PCC6803 to convert dihydroxyacetone phosphate to methylglyoxal, which is subsequently reduced to acetol and then to 1,2‐propanediol. 1,2‐propanediol could be successfully produced by Synechocystis, at an approximate rate of 55 μmol h?1 gCDW?1. Surprisingly, maximal productivity was observed in the stationary phase. The production of 1,2‐propanediol was clearly coupled to the turn‐over of intracellular glycogen. Upon depletion of the glycogen pool, product formation stopped. Reducing the carbon flux to glycogen significantly decreased final product titers. Optimization of cultivation conditions allowed final product titers of almost 1 g L?1 (12 mM), which belongs to the highest values published so far for photoautotrophic production of this compound.
  相似文献   

3.
Optically active alcohols are important building blocks as versatile chiral synthons for asymmetric syntheses of pharmaceuticals and agrochemicals. The aim of this paper is to efficiently prepare chiral 2‐pentanol by means of microorganisms. The gene of dihydroxyacetone reductase (EC 1.1.1.6) from a methylotrophic yeast, Hansenula ofunaensis, was cloned and chiral 2‐pentanol was produced by the recombinant Escherichia coli harboring the gene. The gene encoding the enzyme was cloned from an H. ofunaensis genomic library. In the deduced amino acid sequence of 364 residues, the NAD(H) binding motif and the cysteine residues that correspond to the cysteine ligands in the zinc atom were conserved, as they are in alcohol dehydrogenases from other origins. Dihydroxyacetone reductase was similar to alcohol dehydrogenases of prokaryotes. For the production of chiral compounds, an E. coli HB101 strain was transformed. The H. ofunaensis gene product, dihydroxyacetone reductase, catalyzed the NAD+‐dependent oxidation of 2‐pentanol to 2‐pentanone as well as the corresponding reverse reactions, showing specificity towards the secondary alcohol in (R)‐configuration. From 100 mM 2‐pentanone, (R)‐2‐pentanol (98 mM, > 99.9 % enantiometric excess, e.e.) was obtained in a 30‐min reaction with resting cells of the E. coli HB101 strain harboring the expression plasmid, pSG‐HOD1, which possesses the genes of both dihydroxyacetone reductase and glucose dehydrogenase as an NADH reproducing system. The stereospecificity changed during the reduction, depending on the pH. E. coli HB101 was also transformed by the expression plasmid, pSE‐HOD4, in which the gene of glucose dehydrogenase was removed from pSG‐HOD1, and designated as E. coli HB101 (pSE‐HOD4). E. coli HB101 (pSE‐HOD4) oxidized only (R)‐2‐pentanol in 100 mM of the racemate (R:S = 52:48), and the reaction medium was enriched with (S)‐2‐pentanol (48 mM, 98 % e.e.) after 30 min of incubation. The reaction was sufficiently promoted without the other additives. E. coli transformants expressing the gene of this enzyme could be particularly advantageous to the production of optically active 2‐pentanol.  相似文献   

4.
The ability of bacterial strains to assimilate glycerol derived from biodiesel facilities to produce metabolic compounds of importance for the food, textile and chemical industry, such as 1,3‐propanediol (PD), 2,3‐butanediol (BD) and ethanol (EtOH), was assessed. The screening of 84 bacterial strains was performed using glycerol as carbon source. After initial trials, 12 strains were identified capable of consuming raw glycerol under anaerobic conditions, whereas 5 strains consumed glycerol under aerobiosis. A plethora of metabolic compounds was synthesized; in anaerobic batch‐bioreactor cultures PD in quantities up to 11.3 g/L was produced by Clostridium butyricum NRRL B‐23495, while the respective value was 10.1 g/L for a newly isolated Citrobacter freundii. Adaptation of Cl. butyricum at higher initial glycerol concentration resulted in a PDmax concentration of ~32 g/L. BD was produced by a new Enterobacter aerogenes isolate in shake‐flask experiments, under fully aerobic conditions, with a maximum concentration of ~22 g/L which was achieved at an initial glycerol quantity of 55 g/L. A new Klebsiella oxytoca isolate converted waste glycerol into mixtures of PD, BD and EtOH at various ratios. Finally, another new C. freundii isolate converted waste glycerol into EtOH in anaerobic batch‐bioreactor cultures with constant pH, achieving a final EtOH concentration of 14.5 g/L, a conversion yield of 0.45 g/g and a volumetric productivity of ~0.7 g/L/h. As a conclusion, the current study confirmed the utilization of biodiesel‐derived raw glycerol as an appropriate substrate for the production of PD, BD and EtOH by several newly isolated bacterial strains under different experimental conditions.  相似文献   

5.
Klebsiella pneumoniae is known to produce meso-2,3-butanediol and 2S,3S-butanediol, whereas 2R,3R-butanediol was detected in the culture broth of K. pneumoniae CGMCC 1.6366. The ratio of 2R,3R-butanediol to all isomers obtained using glycerol as the carbon source was higher than that obtained using glucose as the carbon source. Therefore, enzymes involved in glycerol metabolism are likely related to 2R,3R-butanediol formation. In vitro reactions show that glycerol dehydrogenase catalyzes the stereospecific conversion of R-acetoin to 2R,3R-butanediol and S-acetoin to meso-2,3-butanediol. Butanediol dehydrogenase exhibits high (S)-enantioselectivity in ketone reduction. Genes encoding glycerol dehydrogenase, α-acetolactate decarboxylase, and butanediol dehydrogenase were individually disrupted in K. pneumoniae CGMCC 1.6366, and the 2,3-butanediol synthesis characteristics of these mutants were investigated. K. pneumoniae ΔdhaD lost the ability to synthesize 2R,3R-butanediol. K. pneumoniae ΔbudA showed reduced 2R,3R-butanediol synthesis. However, K. pneumoniae ΔbudC produced a high level of 2R,3R-butanediol, and R-acetoin was accumulated in the broth. The metabolic characteristics of these mutants and in vitro experiment results demonstrated the mechanism of the 2,3-butanediol stereoisomer synthesis pathway. Glycerol dehydrogenase, encoded by dhaD, exhibited 2R,3R-butanediol dehydrogenase activity and was responsible for 2R,3R-butanediol synthesis from R-acetoin. This enzyme also contributed to meso-2,3-butanediol synthesis from S-acetoin. Butanediol dehydrogenase, encoded by budC, was the only enzyme that catalyzed the conversion of diacetyl to S-acetoin and further to 2S,3S-butanediol.  相似文献   

6.
Due to its availability, low‐price, and high degree of reduction, glycerol has become an attractive carbon source for the production of fuels and reduced chemicals. Using the platform we have established from the identification of key pathways mediating fermentative metabolism of glycerol, this work reports the engineering of Escherichia coli for the conversion of glycerol into 1,2‐propanediol (1,2‐PDO). A functional 1,2‐PDO pathway was engineered through a combination of overexpression of genes involved in its synthesis from the key intermediate dihydroxyacetone phosphate (DHAP) and the manipulation of the fermentative glycerol utilization pathway. The former included the overexpression of methylglyoxal synthase (mgsA), glycerol dehydrogenase (gldA), and aldehyde oxidoreductase (yqhD). Manipulation of the glycerol utilization pathway through the replacement of the native E. coli PEP‐dependent dihydroxyacetone kinase (DHAK) with an ATP‐dependent DHAK from C. freundii increased the availability of DHAP allowing for higher 1,2‐PDO production. Analysis of the major fermentative pathways indentified ethanol as a required co‐product while increases in 1,2‐PDO titer and yield were achieved through the disruption of the pathways for acetate and lactate production. Combination of these key metabolic manipulations resulted in an engineered E. coli strain capable of producing 5.6 g/L 1,2‐PDO, at a yield of 21.3% (w/w). This strain also performed well when crude glycerol, a by‐product of biodiesel production, was used as the substrate. The titer and yield achieved in this study were favorable to those obtained with the use of E. coli for the production of 1,2‐PDO from common sugars. Biotechnol. Bioeng. 2011; 108:867–879. © 2010 Wiley Periodicals, Inc.  相似文献   

7.
Escherichia coli cells co‐expressing genes coding for Candida tenuis xylose reductase and Candida boidinii formate dehydrogenase were used for the bioreduction of o‐chloroacetophenone with in situ coenzyme recycling. The product, (S)‐1‐(2‐chlorophenyl)ethanol, is a key chiral intermediate in the synthesis of polo‐like kinase 1 inhibitors, a new class of chemotherapeutic drugs. Production of the alcohol in multi‐gram scale requires intensification and scale‐up of the biocatalyst production, biotransformation, and downstream processing. Cell cultivation in a 6.9‐L bioreactor led to a more than tenfold increase in cell concentration compared to shaken flask cultivation. The resultant cells were used in conversions of 300 mM substrate to (S)‐1‐(2‐chlorophenyl)ethanol (e.e. >99.9%) in high yield (96%). Results obtained in a reaction volume of 500 mL were identical to biotransformations carried out in 1 mL (analytical) and 15 mL (preparative) scale. Optimization of product isolation based on hexane extraction yielded 86% isolated product. Biotransformation and extraction were accomplished in a stirred tank reactor equipped with pH and temperature control. The developed process lowered production costs by 80% and enabled (S)‐1‐(2‐chlorophenyl)ethanol production within previously defined economic boundaries. A simple and efficient way to synthesize (S)‐1‐(2‐chlorophenyl)ethanol in an isolated amount of 20 g product per reaction batch was demonstrated. Biotechnol. Bioeng. 2013; 110: 2311–2315. © 2013 Wiley Periodicals, Inc.  相似文献   

8.
Saccharomyces cerevisiae NAD(H)‐dependent 2,3‐butanediol dehydrogenase (Bdh1), a medium chain dehydrogenase/reductase is the main enzyme catalyzing the reduction of acetoin to 2,3‐butanediol. In this work we focused on altering the coenzyme specificity of Bdh1 from NAD(H) to NADP(H). Based on homology studies and the crystal structure of the NADP(H)‐dependent yeast alcohol dehydrogenase Adh6, three adjacent residues (Glu221, Ile222, and Ala223) were predicted to be involved in the coenzyme specificity of Bdh1 and were altered by site‐directed mutagenesis. Coenzyme reversal of Bdh1 was obtained with double Glu221Ser/Ile222Arg and triple Glu221Ser/Ile222Arg/Ala223Ser mutants. The performance of the triple mutant for NADPH was close to that of native Bdh1 for NADH. The three engineered mutants were able to restore the growth of a phosphoglucose isomerase deficient strain (pgi), which cannot grow on glucose unless an alternative NADPH oxidizing system is provided, thus demonstrating their in vivo functionality. These mutants are interesting tools to reduce the excess of acetoin produced by engineered brewing or wine yeasts overproducing glycerol. In addition, they represent promising tools for the manipulation of the NADP(H) metabolism and for the development of a powerful catalyst in biotransformations requiring NADPH regeneration. Biotechnol. Bioeng. 2009; 104: 381–389 © 2009 Wiley Periodicals, Inc.  相似文献   

9.
The formaldehyde dehydrogenase (EC 1.2.1.1) from the yeast Pichia pastoris IFP 206 was purified to homogeneity. The enzyme had a molecular weight of 84,000 daltons and was composed of two identical subunits of a molecular weight of 39,000 daltons. The N-terminal end of the subunits is blocked. The protein showed 6,3 free -SH groups per mole and 12,5 in the presence of NAD+. Enzyme stability was increased by addition of glycerol during the purification.

The enzyme activity is NAD+ and glutathione dependent. The reaction product is S formylglutathione. The presence of an S-formylglutathione hydrolase (EC 3.1.2.12) in the cell free extract was detected. The formaldehyde dehydrogenase showed an optimum pH of 7.9 and an optimum temperature of 47°C. The activation energy was 3.2 kcal/mol. The Michaelis constants for NAD+ and S-hydroxymethyl glutathione were respectively 0.24 mM and 0.26 mM.  相似文献   

10.
Growth of biodiesel industries resulted in increased coproduction of crude glycerol which is therefore becoming a waste product instead of a valuable ‘coproduct’. Glycerol can be used for the production of valuable chemicals, e.g. biofuels, to reduce glycerol waste disposal. In this study, a novel bacterial strain is described which converts glycerol mainly to ethanol and hydrogen with very little amounts of acetate, formate and 1,2‐propanediol as coproducts. The bacterium offers certain advantages over previously studied glycerol‐fermenting microorganisms. Anaerobium acetethylicum during growth with glycerol produces very little side products and grows in the presence of maximum glycerol concentrations up to 1500 mM and in the complete absence of complex organic supplements such as yeast extract or tryptone. The highest observed growth rate of 0.116 h?1 is similar to that of other glycerol degraders, and the maximum concentration of ethanol that can be tolerated was found to be about 60 mM (2.8 g l?1) and further growth was likely inhibited due to ethanol toxicity. Proteome analysis as well as enzyme assays performed in cell‐free extracts demonstrated that glycerol is degraded via glyceraldehyde‐3‐phosphate, which is further metabolized through the lower part of glycolysis leading to formation of mainly ethanol and hydrogen. In conclusion, fermentation of glycerol to ethanol and hydrogen by this bacterium represents a remarkable option to add value to the biodiesel industries by utilization of surplus glycerol.  相似文献   

11.
The sulfate-reducing bacterimDesulfovibrio strain HDv (DSM 6830) grew faster on (S)- and on (R, S)-1,2-propanediol (μmax 0.053 h) than on (R)-propanediol (0.017 h−1) and ethanol (0.027 h−1). From (R, S)-1,2-propanediol-grown cells, an alcohol dehydrogenase was purified. The enzyme was oxygen-labile, NAD-dependent, and decameric; the subunit mol. mass was 48 kDa. The N-terminal amino acid sequence indicated similarity to alcohol dehydrogenases belonging to family III of NAD-dependent alcohol dehydrogenases, the first 21 N-terminal amino acids being identical to those of theDesulfovibrio gigas alcohol dehydrogenase. Best substrates were ethanol and propanol (K m of 0.48 and 0.33 mM, respectively). (R, S)-1,2-Propanediol was a relatively poor substrate for the enzyme, but activities in cell extracts were high enough to account for the growth rate. The enzyme showed a preference for (S)-1,2-propanediol over (R)-1,2-propanediol. Antibodies raised against the alcohol dehydrogenase ofD. gigas showed cross-reactivity with the alcohol dehydrogenase ofDesulfovibrio strain HDv and with cell extracts of six other ethanol-grown sulfate-reducing bacteria.  相似文献   

12.
A new process for (6S)-tetrahydrofolate production from dihydrofolate was designed that used dihydrofolate reductase and an NADPH regeneration system. Glucose dehydrogenase from Gluconobacter scleroides KY3613 was used for recycling of the cofactor. The reaction mixture contained 200 mM dihydrofolate, 220mM glucose, 2mM NADP, 14.4U/ml dihydrofolate reductase, and 14.4U/ml Glucose dehydrogenase, and the reaction was complete after incubation at pH 8.0, and 40°C for 2.5 hr. With (6S)-tetrahydrofolate as the starting material, l-leucovorin was synthesized via a methenyl derivative. The purity of the l-leucovorin was 100%, and its diastereomeric purity was >99.5% d.e. as the (6S)-form.  相似文献   

13.
Phosphoserine aminotransferase (SerC) from Escherichia coli (E. coli) MG1655 is engineered to catalyze the deamination of homoserine to 4‐hydroxy‐2‐ketobutyrate, a key reaction in producing 1,3‐propanediol (1,3‐PDO) from glucose in a novel glycerol‐independent metabolic pathway. To this end, a computation‐based rational approach is used to change the substrate specificity of SerC from l ‐phosphoserine to l ‐homoserine. In this approach, molecular dynamics simulations and virtual screening are combined to predict mutation sites. The enzyme activity of the best mutant, SerCR42W/R77W, is successfully improved by 4.2‐fold in comparison to the wild type when l ‐homoserine is used as the substrate, while its activity toward the natural substrate l ‐phosphoserine is completely deactivated. To validate the effects of the mutant on 1,3‐PDO production, the “homoserine to 1,3‐PDO” pathway is constructed in E. coli by coexpression of SerCR42W/R77W with pyruvate decarboxylase and alcohol dehydrogenase. The resulting mutant strain achieves the production of 3.03 g L?1 1,3‐PDO in fed‐batch fermentation, which is 13‐fold higher than the wild‐type strain and represents an important step forward to realize the promise of the glycerol‐independent synthetic pathway for 1,3‐PDO production from glucose.  相似文献   

14.
Glycerol, the principal byproduct of biodiesel production, can be a valuable carbon source for bioconversion into diverse class of compounds. This article attempts to investigate the mechanistic aspects of ultrasound mediated bioconversion of glycerol to ethanol and 1,3‐propanediol (1,3‐PDO) by immobilized Clostridium pasteurianum cells on silica support. Our approach is of coupling experimental results with simulations of cavitation bubble dynamics and enzyme kinetics. In addition, the statistical analysis (ANOVA) of experimental results was also done. The glycerol uptake by cells was not affected by either immobilization or with ultrasonication. Nonetheless, both immobilization and ultrasonication were found to enhance glycerol consumption. The enhancement effect of ultrasound on glycerol consumption was most marked (175%) at the highest glycerol concentration of 25 g/L (271.7 mM). The highest glycerol consumption (32.4 mM) was seen for 10 g/L (108.7 mM) initial glycerol concentration. The immobilization of cells shifted the metabolic pathway almost completely towards 1,3‐PDO. No formation of ethanol was seen with mechanical shaking, while traces of ethanol were detected with ultrasonication. On the basis of analysis of enzyme kinetics parameters, we attribute these results to increased substrate‐enzyme affinity and decreased substrate inhibition for 1,3‐PDO dehydrogenase in presence of ultrasound that resulted in preferential conversion of glycerol into 1,3‐PDO. Biotechnol. Bioeng. 2013; 110: 1637–1645. © 2013 Wiley Periodicals, Inc.  相似文献   

15.
Klebsiella pneumoniae HR526, a new isolated 1,3‐propanediol (1,3‐PD) producer, exhibited great productivity. However, the accumulation of lactate in the late‐exponential phase remained an obstacle of 1,3‐PD industrial scale production. Hereby, mutants lacking D ‐lactate pathway were constructed by knocking out the ldhA gene encoding fermentative D ‐lactate dehydrogenase (LDH) of HR526. The mutant K. pneumoniae LDH526 with the lowest LDH activity was studied in aerobic fed‐batch fermentation. In experiments using pure glycerol as feedstock, the 1,3‐PD concentrations, conversion, and productivity increased from 95.39 g L?1, 0.48 and 1.98 g L?1 h?1 to 102. 06 g L?1, 0.52 mol mol?1 and 2.13 g L?1 h?1, respectively. The diol (1,3‐PD and 2,3‐butanediol) conversion increased from 0.55 mol mol?1 to a maximum of 0.65 mol mol?1. Lactate would not accumulate until 1,3‐PD exceeded 84 g L?1, and the final lactate concentration decreased dramatically from more than 40 g L?1 to <3 g L?1. Enzymic measurements showed LDH activity decreased by 89–98% during fed‐batch fermentation, and other related enzyme activities were not affected. NADH/NAD+ enhanced more than 50% in the late‐exponential phase as the D ‐lactate pathway was cut off, which might be the main reason for the change of final metabolites concentrations. The ability to utilize crude glycerol from biodiesel process and great genetic stability demonstrated that K. pnemoniae LDH526 was valuable for 1,3‐PD industrial production. Biotechnol. Bioeng. 2009; 104: 965–972. © 2009 Wiley Periodicals, Inc.  相似文献   

16.
Small-angle X-ray scattering was used to probe the structure of actin in the presence of cryosolvents: 1,2-propanediol, glycerol, or a mixture of both solvents. In media devoid of polymerizing salts, a radius of gyration of 23 Å is measured, as expected from the literature. In the presence of 1,2-propanediol alone, the scattering pattern begins to exhibit the characteristic slope of elongated objects with a non-negligible thickness, such as actin filaments polymerized in 40 mM KCl and 1 mM MgCl2. However, only short fragments (radius of gyration 40 Å) are generated. We infer that in a medium of low ionic strength containing 15% 1,2-propanediol, actin assumes a structure closer to that of filamentous actin. 1,2-propanediol apparently induces nucleation of oligomers, as with polymerizing salts, but no propagation occurs. Glycerol and/or propanediol induce no alteration in the structure of individual salt-polymerized actin filaments. Aggregation occurs with propanediol, even in the presence of glycerol. Glycerol alone has no such effect. No shortening is detected within the scale covered, with either solvent, although 1,2-propanediol is known to shorten actin filaments. We suggest that in the absence of salts, 1,2-propanediol induces a conformational change in monomeric actin that is necessary for nucleation. This could correlate with a conformational change of actin protomers within microfilaments observed in the presence of 1,2-propanediol by other authors using different techniques.Abbreviations SAXS small-angle X-ray scattering - G-actin globular monomeric actin - F-actin filamentous polymerized actin Correspondence to: E. Pajot-Augy  相似文献   

17.
NAD+-dependent glycerol dehydrogenase from Cellulomonas sp. NT3060 was purified by a procedure of 10 steps involving crystallization. Dihydroxyacetone was identified as the oxidation product of glycerol with the enzyme. The purified enzyme did not lose activity on heating below 60°C. The enzyme oxidized other alcohols such as 1,2-propanediol, 2,3-butanediol and glycerol-α-monochlorohydrin, beside glycerol. The enzyme activity was inhibited by p-chloromercuribenzoate, Zn2+, Cu2+ and Cd2+. Oxidation of glyberol was activated by Na+ and reduction of dihydroxyacetone was activated by K+ at pH 7.5.  相似文献   

18.
Klebsiella pneumoniae was shown to convert glycerol to 1,3-propanediol, 2,3-butanediol and ethanol under conditions of uncontrolled pH. Formation of 2,3-butanediol starts with some hours' delay and is accompanied by a reuse of the acetate that was formed in the first period. The fermentation was demonstrated in the type strain of K. pneumoniae, but growth was better with the more acid-tolerant strain GT1, which was isolated from nature. In continuous cultures in which the pH was lowered stepwise from 7.3 to 5.4, 2,3-butanediol formation started at pH 6.6 and reached a maximum yield at pH 5.5, whereas formation of acetate and ethanol declined in this pH range. 2,3-Butanediol and acetoin were also found among the products in chemostat cultures grown at pH 7 under conditions of glycerol excess but only with low yields. At any of the pH values tested, excess glycerol in the culture enhanced the butanediol yield. Both effects are seen as a consequence of product inhibition, the undissociated acid being a stronger trigger than the less toxic diols and acid anions. The possibilities for using the fermentation type described to produce 1,3-propanediol and 2,3-butanediol almost without by-products are discussed. Received: 4 February 1998 / Received revision: 30 March 1998 / Accepted: 13 April 1998  相似文献   

19.
In this study, a newly isolated strain screened from the indoxacarb‐rich agricultural soils, Bacillus cereus WZZ006, has a high stereoselectivity to racemic substrate 5‐chloro‐1‐oxo‐2,3‐dihydro‐2‐hydroxy‐1H‐indene‐2‐carboxylic acid methyl ester. (S)‐5‐chloro‐1‐oxo‐2,3‐dihydro‐2‐hydroxy‐1H‐indene‐2‐carboxylic acid methyl ester was obtained by bio‐enzymatic resolution. After the 36‐hour hydrolysis in 50‐mM racemic substrate under the optimized reaction conditions, the e.e.s was up to 93.0% and the conversion was nearly 53.0% with the E being 35.0. Therefore, B cereus WZZ006 performed high‐level ability to produce (S)‐5‐chloro‐1‐oxo‐2,3‐dihydro‐2‐hydroxy‐1H‐indene‐2‐carboxylic acid methyl ester. This study demonstrates a new biocatalytic process route for preparing the indoxacarb chiral intermediates and provides a theoretical basis for the application of new insecticides in agricultural production.  相似文献   

20.
Chiral alcohols are important building blocks for specialty chemicals and pharmaceuticals. The production of chiral alcohols from ketones can be carried out stereo selectively with alcohol dehydrogenases (ADHs). To establish a process for cost‐effective enzyme immobilization on solid phase for application in ketone reduction, we used an established enzyme pair consisting of ADH from Rhodococcus erythropolis and formate dehydrogenase (FDH) from Candida boidinii for NADH cofactor regeneration and co‐immobilized them on modified poly‐p‐hydroxybutyrate synthase (PhaC)‐inclusion bodies that were recombinantly produced in Escherichia coli cells. After separate production of genetically engineered and recombinantly produced enzymes and particles, cell lysates were combined and enzymes endowed with a Kcoil were captured on the surface of the Ecoil presenting particles due to coiled‐coil interaction. Enzyme‐loaded particles could be easily purified by centrifugation. Total conversion of 4'‐chloroacetophenone to (S)‐4‐chloro‐α‐methylbenzyl alcohol could be accomplished using enzyme‐loaded particles, catalytic amounts of NAD+ and formate as substrates for FDH. Chiral GC‐MS analysis revealed that immobilized ADH retained enantioselectivity with 99 % enantiomeric excess. In conclusion, this strategy may become a cost‐effective alternative to coupled reactions using purified enzymes.  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