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.
A xyloglucan-specific endo-1,4β-glucanase (XcXGHA) from Xanthomonas citri pv. mangiferaeindicae has been cloned, expressed in Escherichia coli, purified and characterised. The XcXGHA enzyme belongs to CAZy family GH74 and has catalytic site residues conserved with other xyloglucanases in this family. At its optimal reaction conditions, pH 7.0 and 40 °C, the enzyme has a k cat/K M value of 2.2?×?107 min?1 M?1 on a tamarind seed xyloglucan substrate. XcXGHA is relatively stable within a broad pH range (pH 4–9) and up to 50 °C (t 1/2, 50 °C of 74 min). XcXGHA is proven to be xyloglucan-specific, and a glycan microarray study verifies that XcXGHA catalyses cleavage of xyloglucan extracted from both monocot and dicot plant species. The enzyme catalyses hydrolysis of tamarind xyloglucan in a unique way by cleaving XXXG into XX and XG (X is xylosyl-substituted glucose; G is unsubstituted glucose), is able to degrade more complex xyloglucans and notably is able to cleave near more substituted xyloglucan motifs such as L [i.e. α-l-Fucp-(1?→?2)-β-d-Galp-(1?→?2)-α-d-Xylp-(1?→?6)-β-d-Glcp]. LC-MS/MS analysis of product profiles of tamarind xyloglucan which had been catalytically degraded by XcXGHA revealed that XcXGHA has specificity for X in subsite ?1. The 3D model suggests that XcXGHA consists of two seven-bladed β-propeller domains with the catalytic center formed by the interface of these two domains, which is conserved in xyloglucanases in the GH74 family. However, the XcXGHA has two amino acids (D264 and R472) that differ from the conserved residues of other GH74 xyloglucanases. These two amino acids were predicted to be located on the opposite side of the active site pocket, facing each other and forming a closing surface above the active site pocket. These two amino acids may contribute to the unique substrate specificity of the XcXGHA enzyme.  相似文献   

2.
A novel protodioscin-(steroidal saponin)-glycoside hydrolase, named protodioscin-glycosidase-1 (PGase-1), was purified and characterized from the Aspergillus oryzae strain. The molecular mass of this enzyme was determined to be about 55 kDa based on SDS-polyacrylamide gel electrophoresis. PGase-1 was able to hydrolyze the terminal 26-O-β-d-glucopyranoside of protodioscin (furostanoside) to produce dioscin (spirostanoside), and then further hydrolyze the terminal 3-O-(1?→?4)-α-l-rhamnopyranoside of dioscin to form progenin III. However, PGase-1 could hardly hydrolyze the 3-O-(1?→?2)-α-l-rhamnopyranoside of progenin III, 3-O-β-d-glucoside of trillin, and the 1-O-glycosides of ophiopogonin D (steroidal saponin). In addition, PGase-1 also could hydrolyze the α-d-galactopyranoside, β-d-glucopyranoside, and β-d-galactopyranoside of p-nitrophenyl-glycosides, but the enzyme could not hydrolyze the α-d-mannopyranoside, α-l-arabinopyranoside, α-d-glucopyranoside, β-d-xylopyranoside, and α-l-rhamnopyranoside of p-nitrophenyl-glycosides. These new properties of PGase-1 were significantly different from those of previously described steroidal saponin-glycosidases and the glycosidases currently described in Enzyme Nomenclature by the NC-IUBMB. The gene (termed as pgase-1) encoding PGase-1 was cloned, sequenced, and expressed in Pichia pastoris GS115. The complete nucleotide sequence of pgase-1 consists of 1,725 bp. The recombinant PGase-1 from recombinant P. pastoris GS115 strain also showed the activity hydrolyzing glycosides of steroidal saponins which was similar to that of the wild-type PGase-1 from A. oryzae. The PGase-1 gene is highly similar to Aspergilli α-amylase (EC 3.2.1.1), and PGase-1 should be classified as glycoside hydrolase family 13 by the method of gene sequence-based classification. But the enzyme properties of PGase-1 are different from those of α-amylase in this family.  相似文献   

3.
A recombinant putative N-acyl-d-glucosamine 2-epimerase from Dictyoglomus turgidum was identified as a cellobiose 2-epimerase by evaluating its substrate specificity. The purified enzyme was a 46?kDa monomer with a specific activity of 16.8?μmol?min?1?mg?1 for cellobiose. The epimerization activity was maximal at pH 7.0 and 70?°C with a half-life of 55?h. The isomerization of the glucose at the reducing end of β-1,4- and α-1,4-linked gluco-oligosaccharides to a fructose moiety by the enzyme took place after the epimerization of the glucose to a mannose moiety. The enzyme converted cellobiose to 12.8?% 4-O-β-d-glucopyranosyl-d-mannose and 54.6?% 4-O-β-d-glucopyranosyl-d-fructose as an equilibrium and converted lactose to 12.8?% epilactose and 54.3?% lactulose.  相似文献   

4.
d-galactose is an attractive substrate for bioconversion. Herein, Escherichia coli was metabolically engineered to convert d-galactose into d-galactonate, a valuable compound in the polymer and cosmetic industries. d-galactonate productions by engineered E. coli strains were observed in shake flask cultivations containing 2 g L?1 d-galactose. Engineered E. coli expressing gld coding for galactose dehydrogenase from Pseudomonas syringae was able to produce 0.17 g L?1 d-galactonate. Inherent metabolic pathways for assimilating both d-galactose and d-galactonate were blocked to enhance the production of d-galactonate. This approach finally led to a 7.3-fold increase with d-galactonate concentration of 1.24 g L?1 and yield of 62.0 %. Batch fermentation in 20 g L?1 d-galactose of E. coli ?galK?dgoK mutant expressing the gld resulted in 17.6 g L?1 of d-galactonate accumulation and highest yield of 88.1 %. Metabolic engineering strategy developed in this study could be useful for industrial production of d-galactonate.  相似文献   

5.
The d,d-transpeptidase activity of Penicillin Binding Proteins (PBPs) is essential to maintain cell wall integrity. PBPs catalyze the final step of the peptidoglycan synthesis by forming 4 → 3 cross-links between two peptide stems. Recently, a novel β-lactam resistance mechanism involving l,d-transpeptidases has been identified in Enterococcus faecium and Mycobacterium tuberculosis. In this resistance pathway, the classical 4 → 3 cross-links are replaced by 3 → 3 cross-links, whose formation are catalyzed by the l,d-transpeptidases. To date, only one class of the entire β-lactam family, the carbapenems, is able to inhibit the l,d-transpeptidase activity. Nevertheless, the specificity of this inactivation is still not understood. Hence, the study of this new transpeptidase family is of considerable interest in order to understand the mechanism of the l,d-transpeptidases inhibition by carbapenems. In this context, we present herein the backbone and side-chain 1H, 15N and 13C NMR assignment of the l,d-transpeptidase from Bacillus subtilis (LdtBs) in the apo and in the acylated form with a carbapenem, the imipenem.  相似文献   

6.
A possible role of arabinogalactan proteins in control of shoot regeneration from stem explants of two citrus cultivars, Carrizo citrange and ‘Duncan’ grapefruit, was investigated. Treatment of explants with (β-d-Glc)3 Yariv phenylglycoside, able to bind specifically to AGPs, led to a decrease of cumulative regeneration potential of both Carrizo citrange and ‘Duncan’ grapefruit. For Carrizo, lower cumulative regeneration potential on (β-d-Glc)3 Yariv phenylglycoside-treated explants was the result of both lower number of shoots on the explants that had shoots (explant regeneration potential) and decreased percentage of explants with shoots. In the case of ‘Duncan’, treatment with (β-d-Glc)3 Yariv phenylglycoside reduced cumulative regeneration potential only by lowering the percentage of explants with shoots, but it did not affect the number of shoots on the explants with shoots. Citrus explants treated with (α-d-Man)3 Yariv phenylglycoside, which does not bind AGPs, responded similarly to untreated explants. Transformability of cells on the cut ends of explants was also lower for both cultivars following the treatment of explants with (β-d-Glc)3 Yariv phenylglycoside. Our data suggest that arabinogalactan proteins play important role in processes controlling differentiation and genetic transformation of citrus cells by Agrobacterium.  相似文献   

7.
The catalytic amino acid residues of the β-d-xylosidase (EC 3.2.1.37; GH43), from Thermobifida fusca TM51 (TfBXyl43), were investigated by direct chemical modifications. The pH dependence curves of the kinetic parameters (kcat and kcat/KM) gave pK values for the free enzyme (5.55 ± 0.19; 6.44 ± 0.19), and pK values of for the enzyme–substrate complex (4.85 ± 0.23; 7.60 ± 0.28) respectively, by using an artificial substrate p-nitrophenyl-β-d-xylopyranoside (pNP-Xyl). The detailed inhibition studies demonstrated well the hydrophobic character of the glycon binding site. Carbodiimide-mediated chemical modifications of the enzyme with 1-ethyl-3-(3-dimethylaminopropyl) carbodiimide (EDAC) in the presence of glycine methyl ester supports the conclusion that a carboxylate residue can be fundamental in the catalytic process. We have also synthesized and tested N-bromoacetyl-β-d-xylopyranosylamine (NBAXA) for TfBXyl43 as an affinity label, which also inactivated the enzyme irreversible. The pH dependence studies in both cases of inactivation revealed that the modified group is the catalytic proton donor (NBAXA pKA = 6.68 ± 0,1; EDAC pKA = 7.42 ± 0.22) which displays its essential role in the hydrolytic process. The β-d-xylopyranosylazide as competitive inhibitor protected the enzyme in all cases against the inactivation, suggesting that the chemical modification which has an impact on the activity took place in the active center. Changing of the enzyme conformation was followed by CD spectroscopy, as a result of the NBAXA inactivation. Our study is important because to our knowledge no similar investigations were made in the case of an inverting β-d-xylosidase.  相似文献   

8.
It has long been believed that amino acids comprising proteins of all living organisms are only of the l-configuration, except for Gly. However, peptidyl d-amino acids were observed in hydrolysates of soluble high molecular weight fractions extracted from cells or tissues of various organisms. This strongly suggests that significant amounts of d-amino acids are naturally present in usual proteins. Thus we analyzed the d-amino acid contents of His-tag-purified β-galactosidase and human urocortin, which were synthesized by Escherichia coli grown in controlled synthetic media. After acidic hydrolysis for various times at 110°C, samples were derivatized with 4-fluoro-7-nitro-2, 1, 3-benzoxadiazole (NBD-F) and separated on a reverse-phase column followed by a chiral column into d- and l-enantiomers. The contents of d-enantiomers of Ala, Leu, Phe, Val, Asp, and Glu were determined by plotting index d/(d + l) against the incubation time for hydrolysis and extrapolating the linear regression line to 0 h to eliminate the effect of racemization of amino acids during the incubation. Significant contents of d-amino acids were reproducibly detected, the d-amino acid profile being specific to an individual protein. This finding indicated the likelihood that d-amino acids are in fact present in the purified proteins. On the other hand, the d-amino acid contents of proteins were hardly influenced by the addition of d- or l-amino acids to the cultivation medium, whereas intracellular free d-amino acids sensitively varied according to the extracellular conditions. The origin of these d-amino acids detected in proteins was discussed.  相似文献   

9.
We performed sensory evaluations on 141 bottles of sake and analyzed the relationship between the d-amino acid concentrations, and the taste of the sake using principal component analysis, which yielded seven principal components (PC1–7) that explained 100 % of the total variance in the data. PC1, which explains 33.6 % of the total variance, correlates most positively with strong taste and most negatively with balanced tastes. PC2, which explains 54.4 % of the total variance, correlates most positively with a sweet taste and most negatively with bitter and sour tastes. Sakes brewed with “Kimoto yeast starter” and “Yamahaimoto” had high scores for PC1 and PC2, and had strong taste in comparison with sakes brewed with “Sokujo-moto”. When present at concentrations below 50 μM, d-Ala did not affect the PC1 score, but all the sakes showed a high PC1 score, when the d-Ala was above 100 μM. Similar observations were found for the d-Asp and d-Glu concentrations with regard to PC1, and the threshold concentrations of d-Asp and d-Glu that affected the taste were 33.8 and 33.3 μM, respectively. Certain bacteria present in sake, especially lactic acid bacteria, produce d-Ala, d-Asp and d-Glu during storage, and these d-amino acids increased the PC1 score and produced a strong taste (Nojun). When d- and l-Ala were added to the sakes, the value for the umami taste in the sensory evaluation increased, with the effect of d-Ala being much stronger than that of l-Ala. The addition of 50–5,000 μM dl-Ala did not effect on the aroma of the sakes at all.  相似文献   

10.
Twelve different amino acids were each substituted for threonine-654 in a cloned glucansucrase from Leuconostoc mesenteroides NRRL B-1118. Both the native and the cloned enzyme with threonine at position 654 produced a water-insoluble glucan containing approximately 44 mol% 1,3-disubstituted α-d-glucopyranosyl units and 29 mol% 1,6-disubstituted α-d-glucopyranosyl units. Several substitutions yielded an enzyme that produced an increased percentage of 1,3-disubstituted α-d-glucopyranosyl units, with corresponding decreases in 1,6-disubstituted α-d-glucopyranosyl units. Only one substitution, tyrosine, resulted in a significant increase in the percentage of 1,6-disubstituted α-d-glucopyranosyl units, with a concomitant increase in glucan yield. The mutated enzymes that produced the highest levels of 1,3-disubstituted α-d-glucopyranosyl units were also significantly activated by the addition of dextran, but glucan yields were also lower in these mutants.  相似文献   

11.
Basically the peptidoglycan of Myxobater AL-1 consists of alternating β-1,4-linked N-acetylglucosamic-N-acetylmuramic acid chains. After splitting the aminosugar backbone with a specific algal enzyme three subunits arise: a monomer, a dimer and a trimer. Investigation of the monomer with specific enzymes and comparison of the degradation products to standards derived from other bacterial peptidoglycans suggest the following structure of the monomer peptide: l-alanyl-d-glutamic-l-meso-diaminopimelic-d-alanine. A d-alanyl-d-meso-diaminopimelic acid bond is the bridgebond between the peptides of the subunits.  相似文献   

12.
Whereas an abundance of literature is available on the occurrence of common proteinogenic amino acids (AAs) in edible fruits of the date palm (Phoenix dactylifera L.), recent reports on non-proteinogenic (non-coded) AAs and amino components are scarce. With emphasis on these components we have analyzed total hydrolysates of twelve cultivars of date fruits using automated ion-exchange chromatography, HPLC employing a fluorescent aminoquinolyl label, and GC–MS of total hydrolysates using the chiral stationary phases Chirasil®-L-Val and Lipodex® E. Besides common proteinogenic AAs, relatively large amounts of the following non-proteinogenic amino acids were detected: (2S,5R)-5-hydroxypipecolic acid (1.4–4.0 g/kg dry matter, DM), 1-aminocyclopropane-1-carboxylic acid (1.3–2.6 g/kg DM), γ-amino-n-butyric acid (0.5–1.2 g/kg DM), (2S,4R)-4-hydroxyproline (130–230 mg/kg DM), l-pipecolic acid (40–140 mg/kg DM), and 2-aminoethanol (40–160 mg/kg DM) as well as low or trace amounts (<70 mg/kg DM) of l-ornithine, 5-hydroxylysine, β-alanine, and in some samples (<20 mg/kg DM) of (S)-β-aminoisobutyric acid and (<10 mg/kg DM) l-allo-isoleucine. In one date fruit, traces of α-aminoadipic acid could be determined. Enantiomeric analysis of 6 M DCl/D2O hydrolysates of AAs using chiral capillary gas chromatography–mass spectrometry revealed the presence of very low amounts of d-Ala, d-Asp, d-Glu, d-Ser and d-Phe (1.2–0.4 %, relative to the corresponding l-enantiomers), besides traces (0.2–1 %) of other d-AAs. The possible relevance of non-proteinogenic amino acids in date fruits is briefly addressed.  相似文献   

13.
Selected eight yeast strains isomerized-glycero-d-galacto-heptose tod-manno-heptulose. The conversion is 7–10%. Under identical conditions, the reverse isomerization ofd-manno-heptulose tod-glycero-d-galacto-heptose ord-glycero-d-talo-heptose does not take place.  相似文献   

14.
The tetrasaccharide 2-(p-trifluoroacetamidophenyl)ethylO-α-l-fucopyranosyl-(1–3)-O-(2-acetamido-2-deoxy-β-d-glucopyranosyl)-(1–3)-O-β-d-galactopyranosyl-(1–4)-β-d-glucopyranoside was synthesized from thioglycoside intermediates. The key step was a methyl triflate promoted glycosidation of a lactose-derived 3′,4′-diol with a disaccharide thioglycoside to give a β(1–3)-linked tetrasaccharide derivative in 67% yield.  相似文献   

15.
A recombinant l-fucose isomerase from Caldicellulosiruptor saccharolyticus was purified as a single 68 kDa band with an activity of 76 U mg?1. The molecular mass of the native enzyme was 204 kDa as a trimer. The maximum activity for l-fucose isomerization was at pH 7 and 75°C in the presence of 1 mM Mn2+. Its half-life at 70°C was 6.1 h. For aldose substrates, the enzyme displayed activity in decreasing order for l-fucose, with a k cat of 11,910 min?1 and a K m of 140 mM, d-arabinose, d-altrose, and l-galactose. These aldoses were converted to the ketoses l-fuculose, d-ribulose, d-psicose, and l-tagatose, respectively, with 24, 24, 85, 55% conversion yields after 3 h.  相似文献   

16.
Bifidobacterium longum NRRL B-41409 l-arabinose isomerase (l-AI) was overexpressed in Lactococcus lactis using a phosphate depletion inducible expression system. The resting L. lactis cells harboring the B. longum l-AI were used for production of d-tagatose from d-galactose in the presence of borate buffer. Multivariable analysis suggested that high pH, temperature and borate concentration favoured the conversion of d-galactose to d-tagatose. Almost quantitative conversion (92 %) was achieved at 20 g L?1 substrate and at 37.5 °C after 5 days. The d-tagatose production rate of 185 g L?1 day?1 was obtained at 300 g L?1 galactose, at 1.15 M borate, and at 41 °C during 10 days when the production medium was changed every 24 h. There was no significant loss in productivity during ten sequential 24 h batches. The initial d-tagatose production rate was 290 g L?1 day?1 under these conditions.  相似文献   

17.
d-Tagatose 3-epimerase family enzymes can efficiently catalyze the epimerization of free keto-sugars, which could be used for d-psicose production from d-fructose. In previous studies, all optimum pH values of these enzymes were found to be alkaline. In this study, a d-psicose 3-epimerase (DPEase) with neutral pH optimum from Clostridium bolteae (ATCC BAA-613) was identified and characterized. The gene encoding the recombinant DPEase was cloned and expressed in Escherichia coli. In order to characterize the catalytic properties, the recombinant DPEase was purified to electrophoretic homogeneity using nickel-affinity chromatography. Ethylenediaminetetraacetic acid was shown to inhibit the enzyme activity completely; therefore, the enzyme was identified as a metalloprotein that exhibited the highest activity in the presence of Co2+. Although the DPEase demonstrated the most activity at a pH ranging from 6.5 to 7.5, it exhibited optimal activity at pH 7.0. The optimal temperature for the recombinant DPEase was 55 °C, and the half-life was 156 min at 55 °C. Using d-psicose as the substrate, the apparent K m, k cat, and catalytic efficiency (k cat/K m) were 27.4 mM, 49 s?1, and 1.78 s?1 mM?1, respectively. Under the optimal conditions, the equilibrium ratio of d-fructose to d-psicose was 69:31. For high production of d-psicose, 216 g/L d-psicose could be produced with 28.8 % turnover yield at pH 6.5 and 55 °C. The recombinant DPEase exhibited weak-acid stability and thermostability and had a high affinity and turnover for the substrate d-fructose, indicating that the enzyme was a potential d-psicose producer for industrial production.  相似文献   

18.
The effects of sodium, potassium, sugar inhibitors, and membrane potential on 3H-d-glucose uptake by hepatopancreatic epithelial brush border membrane vesicles (BBMV) of the Atlantic marine shrimp, Litopenaeus setiferus, were investigated. Brush border membrane vesicles were prepared using a MgCl2/EGTA precipitation method and uptake experiments were conducted using a high speed filtration technique. 3H-d-Glucose uptake was stimulated by both sodium and potassium and these transport rates were almost doubled in the presence of an inside-negative-induced membrane potential. Kinetics of 3H-d-glucose influx were hyperbolic functions of both external Na+ or K+, and an induced membrane potential increased influx J max and lowered Km in both salts. 3H-d-Glucose influx versus [glucose] in both Na+ or K+ media also displayed Michaelis–Menten properties that were only slightly affected by induced membrane potential. Phloridzin was a poor inhibitor of 0.5 mM 3H-d-glucose influx, requiring at least 5 mM in NaCl and 10 mM in KCl to significantly reduce hexose transport. Several sugars (d-galactose, α-methyl-d-gluco-pyranoside, unlabeled d-glucose, d-fructose, and d-mannose) were used at 75 mM as potential inhibitors of 0.1 mM 3H-d-glucose influx. Only unlabeled d-glucose, d-fructose, and d-mannose significantly (p < 0.05) reduced labeled glucose transport. An additional experiment using increasing concentrations of d-mannose (0, 10, 25, 75, and 100 mM) showed this hexose to be an effective inhibitor of 0.1 mM 3H-d-glucose uptake at concentrations of 75 mM and higher. As a whole these results suggest that 3H-d-glucose transport by hepatopancreatic BBMV occurs by a carrier system that is able to use both Na+ and K+ as drivers, is enhanced by membrane potential, is relatively refractory to phloridzin, and is only inhibited by itself, d-fructose, and d-mannose. These properties are similar to those exhibited by the mammalian SLC5A9/SGLT4 transporter, suggesting that an invertebrate analogue of this protein may occur in shrimp.  相似文献   

19.
In this study, yeasts associated with lignocellulosic materials in Brazil, including decaying wood and sugarcane bagasse, were isolated, and their ability to produce xylanolytic enzymes was investigated. A total of 358 yeast isolates were obtained, with 198 strains isolated from decaying wood and 160 strains isolated from decaying sugarcane bagasse samples. Seventy-five isolates possessed xylanase activity in solid medium and were identified as belonging to nine species: Candida intermedia, C. tropicalis, Meyerozyma guilliermondii, Scheffersomyces shehatae, Sugiyamaella smithiae, Cryptococcus diffluens, Cr. heveanensis, Cr. laurentii and Trichosporon mycotoxinivorans. Twenty-one isolates were further screened for total xylanase activity in liquid medium with xylan, and five xylanolytic yeasts were selected for further characterization, which included quantitative analysis of growth in xylan and xylose and xylanase and β-d-xylosidase activities. The yeasts showing the highest growth rate and cell density in xylan, Cr. laurentii UFMG-HB-48, Su. smithiae UFMG-HM-80.1 and Sc. shehatae UFMG-HM-9.1a, were, simultaneously, those exhibiting higher xylanase activity. Xylan induced the highest level of (extracellular) xylanase activity in Cr. laurentii UFMG-HB-48 and the highest level of (intracellular, extracellular and membrane-associated) β-d-xylosidase activity in Su. smithiae UFMG-HM-80.1. Also, significant β-d-xylosidase levels were detected in xylan-induced cultures of Cr. laurentii UFMG-HB-48 and Sc. shehatae UFMG-HM-9.1a, mainly in extracellular and intracellular spaces, respectively. Under xylose induction, Cr. laurentii UFMG-HB-48 showed the highest intracellular β-d-xylosidase activity among all the yeast tested. C. tropicalis UFMG-HB 93a showed its higher (intracellular) β-d-xylosidase activity under xylose induction and higher at 30 °C than at 50 °C. This study revealed different xylanolytic abilities and strategies in yeasts to metabolise xylan and/or its hydrolysis products (xylo-oligosaccharides and xylose). Xylanolytic yeasts are able to secrete xylanolytic enzymes mainly when induced by xylan and present different strategies (intra- and/or extracellular hydrolysis) for the metabolism of xylo-oligosaccharides. Some of the unique xylanolytic traits identified here should be further explored for their applicability in specific biotechnological processes.  相似文献   

20.
We previously demonstrated the simplicity of oxygen-deprived Corynebacterium glutamicum to produce d-lactate, a primary building block of next-generation biodegradable plastics, at very high optical purity by introducing heterologous D-ldhA gene from Lactobacillus delbrueckii. Here, we independently evaluated the effects of overexpressing each of genes encoding the ten glycolytic enzymes on d-lactate production in C. glutamicum. We consequently show that while the reactions catalyzed by glucokinase (GLK), glyceraldehyde 3-phosphate dehydrogenase (GAPDH), phosphofructokinase (PFK), triosephosphate isomerase (TPI), and bisphosphate aldolase had positive effects on d-lactate productivity by increasing 98, 39, 15, 13, and 10 %, respectively, in 10 h reactions in minimal salts medium, the reaction catalyzed by pyruvate kinase had large negative effect by decreasing 70 %. The other glycolytic enzymes did not affect d-lactate productivity when each of encoding genes was overexpressed. It is noteworthy that all reactions associated with positive effects are located upstream of glycerate-1,3-bisphosphate in the glycolytic pathway. The d-lactate yield also increased by especially overexpressing TPI encoding gene up to 94.5 %. Interestingly, overexpression of PFK encoding gene reduced the yield of succinate, one of the main by-products of d-lactate production, by 52 %, whereas overexpression of GAPDH encoding gene increased succinate yield by 26 %. Overexpression of GLK encoding gene markedly increased the yield of dihydroxyacetone and glycerol by 10- and 5.8-fold in exchange with decreasing the d-lactate yield. The effect of overexpressing glycolytic genes was also evaluated in 80 h long-term reactions. The variety of effects of overexpressing each of genes encoding the ten glycolytic enzymes on d-lactate production is discussed.  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