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.
Using a conserved sequence motif, a new gene (YAL060W) of the MDR family has been identified in Saccharomyces cerevisiae. The expressed protein was a stereoespecific (2R,3R)-2,3-butanediol dehydrogenase (BDH). The best substrates were (2R,3R)-2,3-butanediol for the oxidation and (3R/3S)-acetoin and 1-hydroxy-2-propanone for the reduction reactions. The enzyme is extremely specific for NAD(H) as cofactor, probably because the presence of Glu223 in the cofactor binding site, instead of the highly conserved Asp223. BDH is inhibited competitively by 4-methylpyrazole with a Ki of 34 μM. Yeast could grow on 2,3-butanediol or acetoin as a sole energy and carbon sources, and a 3.6-fold increase in BDH activity was observed when cells were grown in 2,3-butanediol, suggesting a role of the enzyme in 2,3-butanediol metabolism. However, the disruption of the YAL060W gene was not lethal for the yeast under laboratory conditions, and the disrupted strain could also grow in 2,3-butanediol and acetoin. This suggests that other enzymes, in addition to BDH, can also metabolize 2,3-butanediol in yeast.  相似文献   

2.
The completion of the Saccharomyces cerevisiae genome project in 1996 showed that almost 60% of the potential open reading frames of the genome had no experimentally determined function. Using a conserved sequence motif present in the zinc-containing medium-chain alcohol dehydrogenases, we found several potential alcohol dehydrogenase genes with no defined function. One of these, YAL060W, was overexpressed using a multicopy inducible vector, and its protein product was purified to homogeneity. The enzyme was found to be a homodimer that, in the presence of NAD(+), but not of NADP, could catalyze the stereospecific oxidation of (2R,3R)-2, 3-butanediol (K(m) = 14 mm, k(cat) = 78,000 min(-)(1)) and meso-butanediol (K(m) = 65 mm, k(cat) = 46,000 min(-)(1)) to (3R)-acetoin and (3S)-acetoin, respectively. It was unable, however, to further oxidize these acetoins to diacetyl. In the presence of NADH, it could catalyze the stereospecific reduction of racemic acetoin ((3R/3S)- acetoin; K(m) = 4.5 mm, k(cat) = 98,000 min(-)(1)) to (2R,3R)-2,3-butanediol and meso-butanediol, respectively. The substrate stereospecificity was determined by analysis of products by gas-liquid chromatography. The YAL060W gene product can therefore be classified as an NAD-dependent (2R,3R)-2,3-butanediol dehydrogenase (BDH). S. cerevisiae could grow on 2,3-butanediol as the sole carbon and energy source. Under these conditions, a 3. 5-fold increase in (2R,3R)-2,3-butanediol dehydrogenase activity was observed in the total cell extracts. The isoelectric focusing pattern of the induced enzyme coincided with that of the pure BDH (pI 6.9). The disruption of the YAL060W gene was not lethal for the yeast under laboratory conditions. The disrupted strain could also grow on 2,3-butanediol, although attaining a lesser cell density than the wild-type strain. Taking into consideration the substrate specificity of the YAL060W gene product, we propose the name of BDH for this gene. The corresponding enzyme is the first eukaryotic (2R, 3R)-2,3-butanediol dehydrogenase characterized of the medium-chain dehydrogenase/reductase family.  相似文献   

3.
Zhang GL  Wang CW  Li C 《Biotechnology letters》2012,34(8):1519-1523
The budC gene encoding the meso-2,3-BDH from Klebsiella pneumoniae XJ-Li was expressed in E. coli BL21 (DE3) pLys. Hypothetical amino acid sequence alignments revealed that the enzyme belongs to the short chain dehydrogenase/reductase family. After purification and refolding, the recombinant enzyme had activities of 218 U/mg for reduction of acetoin and 66 U/mg for oxidation of meso-2,3-butanediol. Highest activities were at pH 8.0 and 9.0 respectively. These are higher than other meso-2,3-butanediol dehydrogenases from K. pneumoniae. The low K (m) value (0.65 mM) for acetoin indicated that the enzyme can easily reduce acetoin to meso-2,3-butanediol. There were no significant activities towards 2R,3R-2,3-butanediol, 1,4-butanediol and 2S,3S-2,3-butanediol, suggesting that the enzyme has a high stereospecificity for the meso-dihydric alcohol.  相似文献   

4.
Bacillus subtilis mutants were obtained after the wild strain JNA 3-10 was mutagenized by UV irradiation coupled with diethyl sulfate. A visual filter assay was employed for the qualitative identification of 2,3-butanediol dehydrogenase (BDH) blocked B. subtilis. Selected mutants were tested for the activities of acetoin reductase (AR) and BDH. According to further batch fermentation, one mutant named JNA-UD-6 that produced 24.3 % more acetoin than JNA 3-10 with the corresponding byproducts of 2,3-butanediol decreased by 39.8 % was isolated. A nonsense mutation (p.Tyr118X) that precluded the synthesis of a full-length functional AR/BDH within the bdhA gene of JNA-UD-6 was detected. Acetoin production of JNA-UD-6 was further improved to about 53.9 g/L in a 5-L fermentor with 150 g/L glucose consumed. However,a small amount of 2,3-butanediol was found in late phase of JNA-UD-6 fermentation, and it was due to the existence of a putative gene that encoding a minor AR. This work proved a strategy to efficiently breeding an acetoin high producing strain by traditional mutation methods.  相似文献   

5.
Bacillus subtilis is capable of producing 2,3-butanediol from acetoin by fermentation, but to date, the gene encoding the enzyme responsible, acetoin reductase/2,3-butanediol dehydrogenase (AR/BDH), has remained unknown. A search of the B. subtilis genome database with the amino acid sequences of functional AR/BDHs from Saccharomyces cerevisiae and Bacillus cereus resulted in the identification of a highly similar protein encoded by the B. subtilis ydjL gene. A knockout strain carrying a ydjL::cat insertion mutation was constructed, which (i) abolished 2,3-butanediol production in early stationary phase, (ii) produced no detectable AR or BDH activity in vitro, and (iii) accumulated the precursor acetoin in early stationary phase. The ydjL::cat mutation also affected the kinetics of lactate but not acetate production during stationary-phase cultivation with glucose under oxygen limitation. A very small amount of 2,3-butanediol was detected in very-late-stationary-phase (96-hour) cultures of the ydjL::cat mutant, suggesting the existence of a second gene encoding a minor AR activity. From the data, it is proposed that the major AR/BDH-encoding gene ydjL be renamed bdhA.  相似文献   

6.
As a byproduct of yeast valine metabolism during fermentation, diacetyl can produce a buttery aroma in wine. However, high diacetyl concentrations generate an aromatic off-flavor and poor quality in wine. 2,3-Butanediol dehydrogenase encoded by BDH1 can catalyze the two reactions of acetoin from diacetyl and 2,3-butanediol from acetoin. BDH2 is a gene adjacent to BDH1, and these genes are regulated reciprocally. In this study, BDH1 and BDH2 were overexpressed in Saccharomyces uvarum to reduce the diacetyl production of wine either individually or in combination. Compared with those in the host strain WY1, the diacetyl concentrations in the recombinant strains WY1-1 with overexpressed BDH1, WY1-2 with overexpressed BDH2 alone, and WY1-12 with co-overexpressed BDH1 and BDH2 were decreased by 39.87, 33.42, and 46.71%, respectively. BDH2 was only responsible for converting diacetyl into acetoin, but not for the metabolic pathway of acetoin to 2,3-butanediol in S. uvarum. This study provided valuable insights into diacetyl reduction in wine.  相似文献   

7.
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   

8.
A (2R,3R)-2,3-butanediol dehydrogenase (BDH99::67) from Paenibacillus polymyxa ATCC 12321 was functionally characterized. The genetic characteristics of BDH99::67 are completely different from those of meso- and (2S,3S)-2,3-butanediol dehydrogenases. The results showed that BDH99::67 belongs to the medium-chain dehydrogenase/reductase superfamily and not to the short-chain dehydrogenase/reductase superfamily, to which meso- and (2S,3S)-2,3-butanediol dehydrogenases belong.  相似文献   

9.
2,3-butanediol dehydrogenase (BDH, EC 1.1.1.76) also known as acetoin reductase (AR, EC 1.1.1.4) is the key enzyme converting acetoin (AC) into 2,3-butanediol (BD) and undertaking the irreversible conversion of diacetyl to acetoin in various microorganisms. The existence of three BDHs (R,R-, meso-, and S,S-BDH) product different BD isomers. Catalyzing mechanisms of meso- and S,S-BDH have been understood with the assistance of their X-ray crystal structures. However, the lack of structural data for R,R-BDH restricts the integral understanding of the catalytic mechanism of BDHs. In this study, we successfully crystallized and solved the X-ray crystal structure of Bacillus subtilis R,R-BDH. A zinc ion was found locating in the catalytic center and coordinated by Cys37, His70 and Glu152, helping to stabilize the chiral substrates observed in the predicted molecular docking model. The interaction patterns of different chiral substrates in the molecular docking model explained the react priority measured by the enzyme activity assay of R,R-BDH. Site-directed mutation experiments determined that the amino acids Cys37, Thr244, Ile268 and Lys340 are important in the catalytically active center. The structural information of R,R-BDH presented in this study accomplished the understanding of BDHs catalytic mechanism and more importantly provides useful guidance for the directional engineering of R,R-BDH to obtain high-purity monochiral BD and AC.  相似文献   

10.
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   

11.
2,3-Butanediol dehydrogenase (BDH) catalyzes the NAD-dependent redox reaction between acetoin and 2,3-butanediol. There are three types of homologous BDH, each stereospecific for both substrate and product. To establish how these homologous enzymes possess differential stereospecificities, we determined the crystal structure of l-BDH with a bound inhibitor at 2.0 Å. Comparison with the inhibitor binding mode of meso-BDH highlights the role of a hydrogen-bond from a conserved Trp residue192. Site-directed mutagenesis of three active site residues of meso-BDH, including Trp190, which corresponds to Trp192 of l-BDH, converted its stereospecificity to that of l-BDH. This result confirms the importance of conserved residues in modifying the stereospecificity of homologous enzymes.  相似文献   

12.
Cloning of the Alcaligenes eutrophus alcohol dehydrogenase gene   总被引:7,自引:6,他引:1       下载免费PDF全文
Mutants of Alcaligenes eutrophus which are altered with respect to the utilization of 2,3-butanediol and acetoin were isolated after transposon mutagenesis. The suicide vehicle pSUP5011 was used to introduce the drug resistance transposable element Tn5 into A. eutrophus. Kanamycin-resistant transconjugants of the 2,3-butanediol-utilizing parent strains CF10141 and AS141 were screened for mutants impaired in the utilization of 2,3-butanediol or acetoin. Eleven mutants were negative for 2,3-butanediol but positive for acetoin; they were unable to synthesize active fermentative alcohol dehydrogenase protein (class 1). Forty mutants were negative for 2,3-butanediol and for acetoin (class 2). Tn5-mob was also introduced into a Smr derivative of the 2,3-butanediol-nonutilizing parent strain H16. Of about 35,000 transconjugants, 2 were able to grow on 2,3-butanediol. Both mutants synthesized the fermentative alcohol dehydrogenase constitutively (class 3). The Tn5-labeled EcoRI fragments of genomic DNA of four class 1 and two class 3 mutants were cloned from a cosmid library. They were biotinylated and used as probes for the detection of the corresponding wild-type fragments in a lambda L47 and a cosmid gene bank. The gene which encodes the fermentative alcohol dehydrogenase in A. eutrophus was cloned and localized to a 2.5-kilobase (kb) SalI fragment which is located within a 11.5-kb EcoRI-fragment. The gene was heterologously expressed in A. eutrophus JMP222 and in Pseudomonas oxalaticus. The insertion of Tn5-mob in class 3 mutants mapped near the structural gene for alcohol dehydrogenase on the same 2.5-kb SalI fragment.  相似文献   

13.

Background

Previously, a safe strain, Bacillus amyloliquefaciens B10-127 was identified as an excellent candidate for industrial-scale microbial fermentation of 2,3-butanediol (2,3-BD). However, B. amyloliquefaciens fermentation yields large quantities of acetoin, lactate and succinate as by-products, and the 2,3-BD yield remains prohibitively low for commercial production.

Methodology/Principal Findings

In the 2,3-butanediol metabolic pathway, glyceraldehyde-3-phosphate dehydrogenase (GAPDH) catalyzes the conversion of 3-phosphate glyceraldehyde to 1,3-bisphosphoglycerate, with concomitant reduction of NAD+ to NADH. In the same pathway, 2,3-BD dehydrogenase (BDH) catalyzes the conversion of acetoin to 2,3-BD with concomitant oxidation of NADH to NAD+. In this study, to improve 2,3-BD production, we first over-produced NAD+-dependent GAPDH and NADH-dependent BDH in B. amyloliquefaciens. Excess GAPDH reduced the fermentation time, increased the 2,3-BD yield by 12.7%, and decreased the acetoin titer by 44.3%. However, the process also enhanced lactate and succinate production. Excess BDH increased the 2,3-BD yield by 16.6% while decreasing acetoin, lactate and succinate production, but prolonged the fermentation time. When BDH and GAPDH were co-overproduced in B. amyloliquefaciens, the fermentation time was reduced. Furthermore, in the NADH-dependent pathways, the molar yield of 2,3-BD was increased by 22.7%, while those of acetoin, lactate and succinate were reduced by 80.8%, 33.3% and 39.5%, relative to the parent strain. In fed-batch fermentations, the 2,3-BD concentration was maximized at 132.9 g/l after 45 h, with a productivity of 2.95 g/l·h.

Conclusions/Significance

Co-overexpression of bdh and gapA genes proved an effective method for enhancing 2,3-BD production and inhibiting the accumulation of unwanted by-products (acetoin, lactate and succinate). To our knowledge, we have attained the highest 2,3-BD fermentation yield thus far reported for safe microorganisms.  相似文献   

14.
Acetoin reductase (ACR) catalyzes the conversion of acetoin to 2,3-butanediol. Under certain conditions, Clostridium acetobutylicum ATCC 824 (and strains derived from it) generates both d- and l-stereoisomers of acetoin, but because of the absence of an ACR enzyme, it does not produce 2,3-butanediol. A gene encoding ACR from Clostridium beijerinckii NCIMB 8052 was functionally expressed in C. acetobutylicum under the control of two strong promoters, the constitutive thl promoter and the late exponential adc promoter. Both ACR-overproducing strains were grown in batch cultures, during which 89 to 90% of the natively produced acetoin was converted to 20 to 22 mM d-2,3-butanediol. The addition of a racemic mixture of acetoin led to the production of both d-2,3-butanediol and meso-2,3-butanediol. A metabolic network that is in agreement with the experimental data is proposed. Native 2,3-butanediol production is a first step toward a potential homofermentative 2-butanol-producing strain of C. acetobutylicum.  相似文献   

15.
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   

16.
Acetoin reductase catalyzes the production of 2,3-butanediol from acetoin. The gene encoding the acetoin reductase of Klebsiella pneumoniae CG21 was cloned and expressed in Escherichia coli and Clostridium acetobutylicum ATCC 824. The nucleotide sequence of the gene encoding the enzyme was determined to be 768 bp long. Expression of the K. pneumoniae acetoin reductase gene in E. coli revealed that the enzyme has a molecular mass of about 31,000 Da based on sodium dodecyl sulfate polyacrylamide gel electrophoresis analysis. The K. pneumoniae acetoin reductase gene was cloned into a clostridial/E. coli shuttle vector, and expression of the gene resulted in detectable levels of acetoin reductase activity in both E. coli and C. acetobutylicum. While acetoin, the natural substrate of acetoin reductase, is a typical product of fermentation by C. acetobutylicum, 2,3-butanediol is not. Analysis of culture supernatants by gas chromatography revealed that introduction of the K. pneumoniae acetoin reductase gene into C. acetobutylicum was not sufficient for 2,3-butanediol production even though the cultures were producing acetoin. 2,3-Butanediol was produced by cultures of C. acetobutylicum containing the gene only when commercial acetoin was added. Journal of Industrial Microbiology & Biotechnology (2001) 27, 220–227. Received 12 September 2000/ Accepted in revised form 26 June 2001  相似文献   

17.
Eight representative strains of Alcaligenes eutrophus, two strains of Alcaligenes hydrogenophilus and three strains of Paracoccus denitrificans were examined for their ability to use different alcohols and acetoin as a carbon source for growth. A. eutrophus strains N9A, H16 and derivative strains were unable to grow on ethanol or on 2,3-butanediol. Alcohol-utilizing mutants derived from these strains, isolated in this study, can be categorized into two major groups: Type I-mutants represented by strain AS1 occurred even spontaneously and were able to grow on 2,3-butanediol (t d=2.7–6.4 h) and on ethanol (t d=15–50 h). The fermentative alcohol dehydrogenase was present on all substrates tested, indicating that this enzyme in vivo is able to oxidize 2,3-butanediol to acetoin which is a good substrate for wild type strains. Type II-mutants represented by strain AS4 utilize ethanol as a carbon source for growth (t d=3–9 h) but do not grow on butanediol. In these mutants the fermentative alcohol dehydrogenase is only present in cells cultivated under conditions of restricted oxygen supply, but a different NAD-dependent alcohol dehydrogenase is present in ethanol grown cells. Cells grown on ethanol, acetoin or 2,3-butanediol synthesized in addition two proteins exhibiting NAD-dependent acetaldehyde dehydrogenase activity and acetate thiokinase. An acylating acetaldehyde dehydrogenase (EC 1.2.1.10) was not detectable. Applying the colistin- and pin point-technique for mutant selection to strain AS1, mutants, which lack the fermentative alcohol dehydrogenase even if cultivated under conditions of restricted oxygen supply, were isolated; the growth pattern served as a readily identifiable phenotypic marker for the presence or absence of this enzyme.  相似文献   

18.
The capacity to produce 2,3-butanediol by 90 strains of four different species of wine yeasts (Kloeckera apiculata, Saccharomyces cerevisiae, Saccharomycodes ludwigii, Zygosaccharomyces bailii) was tested in grape must by automated multiple development HPTLC. The total amount of 2,3-butanediol produced varied between 23mg l–1 and 857.7mg l–1 according to the yeast species. S. cerevisiae and Z. bailii behaved similarly, producing elevated amounts of 2,3-butanediol. K. apiculata and Sc. ludwigii, in contrast, were low producers. When considerable amounts of 2,3-butanediol were found, little acetoin was present; the amounts of butanediol and acetoin were characteristic of the individual species.  相似文献   

19.
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   

20.
Aerobacter (Enterobacter) aerogenes wild type and three mutants deficient in the formation of acetoin and 2,3-butanediol were grown in a glucose minimal medium. Culture densities, pH, and diacetyl, acetoin, and 2,3-butanediol levels were recorded. The pH in wild-type cultures dropped from 7.0 to 5.8, remained constant while acetoin and 2,3-butanediol were formed, and increased to pH 6.5 after exhaustion of the carbon source. More 2,3-butanediol than acetoin was formed initially, but after glucose exhaustion reoxidation to acetoin occurred. The three mutants differed from the wild type in yielding acid cultures (pH below 4.5). The wild type and one of the mutants were grown exponentially under aerobic and anaerobic conditions with the pH fixed at 7.0, 5.8, and 5.0, respectively. Growth rates decreased with decreasing pH values. Aerobically, this effect was weak, and the two strains were affected to the same degree. Under anaerobic conditions, the growth rates were markedly inhibited at a low pH, and the mutant was slightly more affected than the wild type. Levels of alcohol dehydrogenase were low under all conditions, indicating that the enzyme plays no role during exponential growth. The levels of diacetyl (acetoin) reductase, lactate dehydrogenase, and phosphotransacetylase were independent of the pH during aerobic growth of the two strains. Under anaerobic conditions, the formation of diacetyl (acetoin) reductase was pH dependent, with much higher levels of the enzyme at pH 5.0 than at pH 7.0. Lactate dehydrogenase and phosphotransacetylase revealed the same pattern of pH-dependent formation in the mutant, but not in the wild type.  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