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.
Fermentation of xylose by Klebsiella pneumoniae (ATCC 8724) producers meso and nonmeso 2,3-butaneodiol. The enzyme Kinetic of 2,3-butanediol stereoisomer formation from acetone is currently under study in our laboratory. Modeling of these kinetics requires resolution of meso and racemic 2,3-butanediol and positive identification of these resolved components. We report their resolution by aqueous liquid chromatography on both an analytical and a preparative scale. The resolved stereoisomer were identified by a combination of gas chromatography, gas chromatography/mass spectroscopy, 13C-NMR spectroscopy, optical activity, and, melting points of the m-dinitrobenzoyl eaters of meso and racemic 2,3-butanediol. An aqueous liquid chromatographic technique for resolving and qualifying major components of a butanediol fermentation mixture in 40 min is presented.  相似文献   

2.
The meso-2,3-butanediol dehydrogenase (meso-BDH) from S. marcescens H30 is responsible for converting acetoin into 2,3-butanediol during sugar fermentation. Inactivation of the meso-BDH encoded by budC gene does not completely abolish 2,3-butanediol production, which suggests that another similar enzyme involved in 2,3-butanediol formation exists in S. marcescens H30. In the present study, a glycerol dehydrogenase (GDH) encoded by gldA gene from S. marcescens H30 was expressed in Escherichia coli BL21(DE3), purified and characterized for its properties. In vitro conversion indicated that the purified GDH could catalyze the interconversion of (3S)-acetoin/meso-2,3-butanediol and (3R)-acetoin/(2R,3R)-2,3-butanediol. (2S,3S)-2,3-Butanediol was not a substrate for the GDH at all. Kinetic parameters of the GDH enzyme showed lower K m value and higher catalytic efficiency for (3S/3R)-acetoin in comparison to those for (2R,3R)-2,3-butanediol and meso-2,3-butanediol, implying its physiological role in favor of 2,3-butanediol formation. Maximum activity for reduction of (3S/3R)-acetoin and oxidations of meso-2,3-butanediol and glycerol was observed at pH 8.0, while it was pH 7.0 for diacetyl reduction. The enzyme exhibited relative high thermotolerance with optimum temperature of 60 °C in the oxidation–reduction reactions. Over 60 % of maximum activity was retained at 70 °C. Additionally, the GDH activity was significantly enhanced for meso-2,3-BD oxidation in the presence of Fe2+ and for (3S/3R)-acetoin reduction in the presence of Mn2+, while several cations inhibited its activity, particularly Fe2+ and Fe3+ for (3S/3R)-acetoin reduction. The properties provided potential application for single configuration production of acetoin and 2,3-butanediol .  相似文献   

3.
The functional reconstruction of acetoin and meso-2,3-butanediol (meso-2,3-BD) biosynthetic pathways in Escherichia coli have been explored systematically. Pathway construction involved the in vsivo screening of prospective pathway isozymes of yeast and bacterial origin. After substantial engineering of the host background to increase pyruvate availability, E. coli YYC202(DE3) ldhA( ilvC( expressing ilvBN from E. coli and aldB from L. lactis (encoding acetolactate synthase and acetolactate decarboxylase activities, respectively) was able to produce up to 870 mg/L acetoin, with no coproduction of diacetyl observed. These strains were also found to produce small quantities of meso-2,3-BD, suggesting the existence of endogenous 2,3-BD dehydrogenase activity. Finally, the coexpression of bdh1 from S. cerevisiae, encoding 2,3-BD dehydrogenase, in this strain resulted in the production of up to 1120 mg/L meso-2,3-BD, with glucose a yield of 0.29 g/g. While disruption of the native lactate biosynthesis pathway increased pyruvate precursor availability to this strain, increased availability of NADH for acetoin reduction to meso-2,3-BD was found to be the most important consequence of ldhA deletion.  相似文献   

4.
Two 2,3-butanediol dehydrogenases (enzymes 1 and 2; molecular weight of each, 170,000) have been partially purified from Lactococcus lactis subsp. lactis (Streptococcus diacetylactis) D10 and shown to have reductase activity with either diacetyl or acetoin as the substrate. However, the reductase activity with 10 mM diacetyl was far greater for both enzymes (7.0- and 4.7-fold for enzymes 1 and 2, respectively) than with 10 mM acetoin as the substrate. In contrast, when acetoin and diacetyl were present together, acetoin was the preferred substrate for both enzymes, with enzyme 1 showing the more marked preference for acetoin. meso-2,3-Butanediol was the only isomeric product, with enzyme 1 independent of the substrate combinations. For enzyme 2, both the meso and optical isomers of 2,3-butanediol were formed with acetoin as the substrate, but only the optical isomers were produced with diacetyl as the substrate. With batch cultures of strain D10 at or near the point of citrate exhaustion, the main isomers of 2,3-butanediol present were the optical forms. If the pH was sufficiently high (>pH 5), acetoin reduction occurred over time and was followed by diacetyl reduction, and meso-2,3-butanediol became the predominant isomer. Interconversion of the optical isomers into the meso isomer did occur. The properties of 2,3-butanediol dehydrogenases are consistent with diacetyl and acetoin removal and the appearance of the isomers of 2,3-butanediol.  相似文献   

5.
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   

6.
P Boutron 《Cryobiology》1992,29(3):347-358
A 2,3-butanediol containing 96.7% (w/w) racemic mixture of the levo and dextro isomers and only 3.1% (w/w) of the meso isomer (called 2,3-butanediol 97% dl) has been used for the cryoprotection of red blood cells. The erythrocytes were cooled to -196 degrees C at rates between 2 and 3500 degrees C/min, followed by slow or rapid warming. Up to 20% (w/w) of this polyalcohol, only the classical peak of survival is observed, as with up to 20% (w/w) 1,2-propanediol or 1,3-butanediol. Twenty percent 2,3-butanediol 97% dl can protect red blood cells very efficiently. The maximum survival, of 90%, as with 20% glycerol, is a little lower than with 20% 1,2-propanediol and higher than with 20% 1,3-butanediol. Fifteen percent 2,3-butanediol protects fewer red blood cells than 15% glycerol or 1,2-propanediol, with a maximum survival of about 80%. The best cryoprotection by 30% 2,3-butanediol 97% dl is obtained at the slowest cooling and warming rates, where survival approaches 90%. After a minimum, an increase of survival is observed at the fastest cooling rates, which would correspond to complete vitrification. These rates are lower than with 30%, 1,2-propanediol or 1,3-butanediol, in agreement with the higher glass-forming tendency of 2,3-butanediol 97% dl solutions. In agreement with the remarkable physical properties of its aqueous solutions, the present experiments also suggest that 2,3-butanediol containing mainly the levo and dextro isomers could be a very useful cryoprotectant for organ cryopreservation. However, it would perhaps be better to use it in combination with other cryoprotectants, since it is a little more toxic than glycerol or 1,2-propanediol at high concentrations.  相似文献   

7.
By way of investigating possible mechanisms for the abiotic amplification of small enantiomeric excesses (e.e.'s) in almost racemic mixtures of amino acid enantiomers, we have undertaken a quantitative study of the base-initiated partial polymerization of leucine and valineN-carboxy-anhydride (NCA) enantiomer mixtures containing known excesses of both theR- andS-forms. Polymerization to the extent ofca. 50% of leucine NCA having an 8–70% e.e. of either theR- orS-enantiomer led to an e.e. enhancement in the polymer, which contained a 12–84% e.e. of that enantiomer which predominated in the original monomer NCA. A corresponding decrease in the e.e. of the initially predominant enantiomer was noted in the unpolymerized residue from each reaction. Polymerization to the extent of 25–50% of mixtures of valine NCA enantiomers containing a 12–13% e.e. of eitherR- orS-isomer led to polymers showing a 7–8%decrease in the e.e. of the initially predominant enantiomer, and to an increase of its e.e. in the unpolymerized residue. These divergent results, the latter of which is quite novel, are compared with earlier qualitative results in the literature and are discussed briefly from the viewpoint of both mechanism and the amplification of optical activity.A portion of this material was presented at a symposium on The Origins of Optical Activity in Nature, Chemical Institute of Canada, University of British Columbia, Vancouver, B.C., June 5, 1979.  相似文献   

8.
Microalgal biomass was hydrolyzed using a solid acid catalyst with the aid of liquid acid. The use of solid acid as the main catalyst instead of liquid acid was to omit subsequent neutralization and/or desalination steps, which are commonly required in using the resulting hydrolysates for microbial fermentation. The hydrolysis of 10 g/L of lipid-extracted Chlorella vulgaris containing 12.2% carbohydrates using 7.6 g/L Amberlyst 36 and 0.0075 N nitric acid at 150°C resulted in 1.08 g/L of mono-sugars with a yield of 88.5%. For hydrolysis of higher concentrations of the biomass over 10 g/L, the amount of Amberlyst 36 needed to be increased in proportion to the biomass concentration to maintain similar levels of hydrolysis performance. Increasing the solid acid concentration protected the surface of the solid acid from being severely covered by cell debris during the reaction. A hydrolysate of lipid-extracted C. vulgaris 50 g/L was used, with no post-treatment of desalination, for the cultivation of Klebsiella oxytoca producing 2,3-butanediol. Cell growth in the hydrolysate was found to be almost the same as in the conventional medium with the same monosaccharide composition, confirming its fermentation compatibility. It was noticeable that the yield of 2,3-butanediol with the hydrolysate was observed to be 2.6 times higher than that with the conventional medium. © 2018 American Institute of Chemical Engineers Biotechnol. Prog., 35: e2729, 2019  相似文献   

9.
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   

10.
NAD-dependent butanediol dehydrogenase (Bdh1p) from Saccharomyces cerevisiae reversibly transforms acetoin to 2,3-butanediol in a stereospecific manner. Deletion of BDH1 resulted in an accumulation of acetoin and a diminution of 2,3-butanediol in two S. cerevisiae strains under two different growth conditions. The concentrations of (2R,3R)-2,3-butanediol are mostly dependent on Bdh1p activity, while those of (meso)-2,3-butanediol are also influenced by the activity of NADP(H)-dependent oxidoreductases. One of them has been purified and shown to be d-arabinose dehydrogenase (Ara1p), which converts (R/S)-acetoin to meso-2,3-butanediol and (2S,3S)-2,3-butanediol. Deletion of BDH2, a gene adjacent to BDH1, whose encoded protein is 51% identical to Bdh1p, does not significantly alter the levels of acetoin or 2,3-butanediol in comparison to the wild-type strain. Furthermore, we have expressed Bdh2p with a histidine tag and have shown it to be inactive toward 2,3-butanediol. A whole-genome expression analysis with microarrays demonstrates that BDH1 and BDH2 are reciprocally regulated.Acetoin and 2,3-butanediol are minor products generated by Saccharomyces cerevisiae during alcohol fermentation. Their sensory impacts on wine are poorly documented. Acetoin may affect the wine bouquet, although its perception threshold in wine is relatively high, around 150 mg/liter (21, 31). On the other hand, 2,3-butanediol is odorless (33) and cannot be expected to appreciably affect the sensory quality of wine. However, the compound may contribute to the wine body (28).Acetaldehyde, pyruvate, and α-acetolactate are the main precursors of acetoin in S. cerevisiae. Acetoin can be formed from acetaldehyde and/or pyruvate through an anomalous reaction of pyruvate decarboxylase. Thus, although its main activity is to irreversibly decarboxylate pyruvate to acetaldehyde, it can also catalyze carbon-carbon bond formation, yielding acetoin from pyruvate and/or acetaldehyde (2, 4). In addition, α-acetolactate would produce acetoin through its nonenzymatic decarboxylation to diacetyl and subsequent reduction to acetoin through the action of several NADH- and NADPH-dependent oxidoreductases (12). However, the situation is more complex in wine fermentation, where other yeasts and bacteria display supplementary enzymatic activities capable of producing both acetoin and 2,3-butanediol (1, 27).We have previously characterized a butanediol dehydrogenase (Bdh1p) as a medium-chain dehydrogenase/reductase (MDR) that can reversibly transform R-acetoin and S-acetoin to (2R,3R)-2,3-butanediol and meso-2,3-butanediol, respectively, in a NAD(H)-dependent reaction (10). BDH2 is a gene adjacent to BDH1 whose uncharacterized protein product (Bdh2p) shares 51% sequence identity with Bdh1p. To evaluate the in vivo roles of Bdh1p and Bdh2p, we compared the levels of several extracellular metabolites in cultures of wild-type and deficient strains. The results show that, although Bdh1p is the main enzyme in 2,3-butanediol production [essentially the (2R,3R)-2,3-butanediol stereoisomer], some meso-2,3-butanediol is still produced by the bdh1Δ strains. We have characterized Ara1p as an oxidoreductase that can reduce racemic acetoin to meso-2,3-butanediol and (2S,3S)-2,3-butanediol in the presence of NADPH.Furthermore, we have overexpressed Bdh2p with a histidine tag at its carboxyl terminus and have shown it to be inactive toward acetoin and 2,3-butanediol. A microarray study indicated that BDH1 and BDH2 are reciprocally regulated under the conditions studied.  相似文献   

11.

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   

12.
Nutritive value of 4 glycol esters, i.e. ethanediol diacetate, 1,2-propanediol diacetate, 1,3-butanediol diacetate and 1,3-butanediol dioctylate, was estimated biologically by feeding the esters to growing chicks and rats. Energy in the esters taken by both chicks and rats was well utilized, though feed intake of the diets containing the esters at high level tended to decrease. Bitter taste of the esters was suspected to be related to low appetite. The acetates were somewhat volatile and released free acetic acid in the diet during storage. These properties of the acetates makes their use for dietary energy source difficult in practical condition.  相似文献   

13.
Production of 2,3-butanediol by Bacillus subtilis takes place in late-log or stationary phase, depending on the expression of bdhA gene encoding acetoin reductase, which converts acetoin to 2,3-butanediol. The present work focuses on the development of a strain of B. subtilis for enhanced production of 2,3-butanediol in early log phase of growth cycle. For this, the bdhA gene was expressed under the control of P alsSD promoter of AlsSD operon for acetoin fermentation which served the substrate for 2,3-butanediol production. Addition of acetic acid in the medium induced the production of 2,3-butanediol by 2-fold. Two-step aerobic–anaerobic fermentation further enhanced 2,3-butanediol production by 4-fold in comparison to the control parental strain. Thus, addition of acetic acid and low dissolved oxygen in the medium are involved in activation of bdhA gene expression from P alsSD promoter in early log phase. Under the conditions tested in this work, the maximum production of 2,3-butanediol, 2.1 g/l from 10 g/l glucose, was obtained at 24 h. Furthermore, under the optimized microaerophilic condition, the production of 2,3-butanediol improved up to 6.1 g/l and overall productivity increased by 6.7-fold to 0.4 g/l h in the engineered strain compared to that in the parental control.  相似文献   

14.
Bacillus licheniformis produced 2,3-butanediol from glucose with an optimum yield of 47 g/100 g glucose after 72 h of growth on a peptone/beef extract medium containing 2% (w/v) glucose at pH 6.0 and 37°C. This yield of 2,3-butanediol was higher than those previously reported forKlebsiella oxytoca (37 g/100 g glucose) andBacillus polymyxa (24 g/100 glucose).  相似文献   

15.
ImmobilizedCandida antarctica lipase was used to catalyze the separation of ketoprofen into its components by means of esterification followed by the enzymatic hydrolysis of the ester product. In this study, ketoprofen underwent esterification to ethanol in the presence of isooctane. When the reaction was complete, 58.3% of the ketoprofen had been transformed into an ester. The ketoprofen remaining in solution after the reation was complete consisted primarily of itsS-enantiomer (83.0%), while the 59.4% of the ketoprofen component of the ester consisted of itsR-enantiomer. We then subjected the ester product to enzymatic hydrolysis in the presence of the same enzyme and produced a ketoprofen product rich in theR-enantiomer; 77% of this product consisted of theR-enantiomer when 50% of the ester had been hydrolyzed, and 90% of it consisted of theR-enantiomer when 30% of the ester had been hydrolyzed. By contrast, theR-enantiomer levels only reached approximately 42 and 65%, respectively, when 50 and 30% of the racemic ester was hydrolyzed under the same conditions.  相似文献   

16.
Anaerobic nitrite production (thein vivo NO3-R activity) in an incubation medium lacking exogenous nitrate but containing 0.5%n-propanol and 0.1% Triton X-100 showed higher correlation (y - ax b) with the level of endogenous nitrate inPisum sativum L. leaves than thein vitro nitrate reductase activity. Thein vivo NO3-R activity correlated well with thein vitro activity up to the 50 ppm NO3-N level of endogenous nitrate. The ratioin vivo: in vitro activity slightly decreased with increasing level of endogenous nitrate in leaf tissue.  相似文献   

17.
Summary Hemicellulose-rich fractions from several agricultural residues were converted to 2,3-butanediol by a combined enzymatic hydrolysis and fermentation process. Culture filtrates from Trichoderma harzianum E58 were used to hydrolyze the substrates while Klebsiella pneumoniae fermented the liberated sugars to 2,3-butanediol. Approximately 50–60% of a 5% (w/v) xylan preparation could be hydrolyzed and quantitatively converted to 2,3-butanediol using this procedure. Although enzymatic hydrolysis was optimal at pH 5.0 and 50° C, the combined hydrolysis and fermentation was most efficient at pH 6.5 and 30° C. Combined hydrolysis and fermentation resulted in butanediol levels that were 20–40% higher than could be obtained with a separate hydrolysis and fermentation process. The hemicellulose-rich water-soluble fractions obtained from a variety of steam-exploded agricultural residues could be readily used by the combined hydrolysis and fermentation approach resulting in butanediol yields of 0.4–0.5 g/g of reducing sugar utilized.  相似文献   

18.
Large-scale preparation of (R)-(–)-1,3-butanediol (R-BDO), an important chiral synthon, from the racemate by Candida parapsilosis IFO 1396 was investigated. We found that ethanol accumulated during culture enhanced the secondary alcohol oxido-reduction activity of cells. Large-scale preparation of R-BDO was done using a fermentor. 3092 g of R-BDO was obtained from the racemate by the use of this strain with 94.0% enantiomeric excess.  相似文献   

19.
Biochemical 2,3-butanediol is a renewable material with the potential to be used as an alternative fuel. However, in the lack of an effective separation process has limited its industrial application. In this paper, an effective process was achieved to separate 2,3-butanediol by reactive-extraction. Acetaldehyde and cyclohexane were chosen as the reactant and extractant, respectively. Ion-exchange resin HZ732 was used as the catalyst. Reaction equilibrium and a kinetic study on the reaction between 2,3-butanediol and acetaldehyde were investigated to provide basic data for process development. The reaction enthalpy and activation energy of reaction of 2,3-butanediol and acetaldehyde were ?30.05 ± 1.62 KJ/mol and 45.29 ± 2.89 KJ/mol, respectively. Feasible conditions were obtained as follows: operating temperature = 20°C, acetaldehyde: 2,3-butanediol = 0.5:1 (w/w), cyclohexane: fermentation broth = 0.5:1 (w/w), catalyst amount = 100 g/L, stirring rate = 500 rpm and three-stage counter-current extraction method was used. Under these conditions, the total yield rate of 2,3-butanediol from fermentation broth was over 90% and the mass fraction of 2,3-butanediol in the final product reached 99%.  相似文献   

20.
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   

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