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 46 毫秒
1.
The metabolism of glucose by nongrowing cells of Lactococcus lactis strain FI7851, constructed from the wild-type L. lactis strain MG1363 by disruption of the lactate dehydrogenase (ldh) gene [Gasson, M.J., Benson, K., Swindel, S. & Griffin, H. (1996) Lait 76, 33-40] was studied in a noninvasive manner by 13C-NMR. The kinetics of the build-up and consumption of the pools of intracellular intermediates mannitol 1-phosphate, fructose 1,6-bisphosphate, 3-phosphoglycerate, and phosphoenolpyruvate as well as the utilization of [1-13C]glucose and formation of products (lactate, acetate, mannitol, ethanol, acetoin, 2,3-butanediol) were monitored in vivo with a time resolution of 30 s. The metabolism of glucose by the parental wild-type strain was also examined for comparison. A clear shift from typical homolactic fermentation (parental strain) to a mixed acid fermentation (lactate dehdydrogenase deficient; LDHd strain) was observed. Furthermore, high levels of mannitol were transiently produced and metabolized once glucose was depleted. Mannitol 1-phosphate accumulated intracellularly up to 76 mM concentration. Mannitol was formed from fructose 6-phosphate by the combined action of mannitol-1-phosphate dehydrogenase and phosphatase. The results show that the formation of mannitol 1-phosphate by the LDHd strain during glucose catabolism is a consequence of impairment in NADH oxidation caused by a highly reduced LDH activity, the transient production of mannitol 1-phosphate serving as a regeneration pathway for NAD+ regeneration. Oxygen availability caused a drastic change in the pattern of intermediates and end-products, reinforcing the key-role of the fulfilment of the redox balance. The flux control coefficients for the step catalysed by mannitol-1-phosphate dehydrogenase were calculated and the implications in the design of metabolic engineering strategies are discussed.  相似文献   

2.
Mannitol is a sugar polyol claimed to have health-promoting properties. A mannitol-producing strain of Lactococcus lactis was obtained by disruption of two genes of the phosphoenolpyruvate (PEP)-mannitol phosphotransferase system (PTSMtl). Genes mtlA and mtlF were independently deleted by double-crossover recombination in strain L. lactis FI9630 (a food-grade lactate dehydrogenase-deficient strain derived from MG1363), yielding two mutant (ΔldhΔmtlA and ΔldhΔmtlF) strains. The new strains, FI10091 and FI10089, respectively, do not possess any selection marker and are suitable for use in the food industry. The metabolism of glucose in nongrowing cell suspensions of the mutant strains was characterized by in vivo 13C-nuclear magnetic resonance. The intermediate metabolite, mannitol-1-phosphate, accumulated intracellularly to high levels (up to 76 mM). Mannitol was a major end product, one-third of glucose being converted to this hexitol. The double mutants, in contrast to the parent strain, were unable to utilize mannitol even after glucose depletion, showing that mannitol was taken up exclusively by PEP-PTSMtl. Disruption of this system completely blocked mannitol transport in L. lactis, as intended. In addition to mannitol, approximately equimolar amounts of ethanol, 2,3-butanediol, and lactate were produced. A mixed-acid fermentation (formate, ethanol, and acetate) was also observed during growth under controlled conditions of pH and temperature, but mannitol production was low. The reasons for the alteration in the pattern of end products under nongrowing and growing conditions are discussed, and strategies to improve mannitol production during growth are proposed.  相似文献   

3.
4.
5.
A mannitol phosphotransferase system (PTS) was identified in Bacillus stearothermophilus by in vitro complementation with Escherichia coli EI, HPr, and IIA(Mtl). Degenerate primers based on regions of high amino acid similarity in the E. coli and Staphylococcus carnosus EII(Mt1) were used to develop a digoxigenin-labeled probe by PCR. Using this probe, we isolated three overlapping DNA fragments totaling 7.2 kb which contain the genes mtlA, mtlR, mtlF, and mtlD, encoding the mannitol IICB,a regulator, IIA, and a mannitol-1-phosphate dehydrogenase, respectively. The mtl4 gene consists of 1,413 bp coding for a 471-amino-acid protein with a calculated mass of 50.1 kDa. The amino acid sequence shows high similarity with the sequence of IICB(Mtl) of S. carnosus and the IICB part of the IICBA(Mtl)s of E. coli and B. subtilis. The enzyme could be functionally expressed in E. coli by placing it behind the strong tac promoter. The rate of thermal inactivation at 60 degrees C of B. stearothermophilus HCB(Mt1) expressed in E. coli was two times lower than that of E. coli IICB(Mtl). IICB(Mtl) in B. stearothermophilus is maximally active at 85 degrees C and thus very thermostable. The enzyme was purified on Ni-nitrilotriacetic acid resin to greater than 95% purity after six histidines were fused to the C-terminal part of the transporter.  相似文献   

6.
To obtain a mannitol-producing Lactococcus lactis strain, the mannitol 1-phosphate dehydrogenase gene (mtlD) from Lactobacillus plantarum was overexpressed in a wild-type strain, a lactate dehydrogenase(LDH)-deficient strain, and a strain with reduced phosphofructokinase activity. High-performance liquid chromatography and (13)C nuclear magnetic resonance analysis revealed that small amounts (<1%) of mannitol were formed by growing cells of mtlD-overexpressing LDH-deficient and phosphofructokinase-reduced strains, whereas resting cells of the LDH-deficient transformant converted 25% of glucose into mannitol. Moreover, the formed mannitol was not reutilized upon glucose depletion. Of the metabolic-engineering strategies investigated in this work, mtlD-overexpressing LDH-deficient L. lactis seemed to be the most promising strain for mannitol production.  相似文献   

7.
Mannitol (Mtl), not previously reported as an intracellular component of bacteria, although it has been found as an extracellular end product of anaerobic carbohydrate metabolism, accumulated within strains of all 10 staphylococcal species tested after aerobic incubation of washed cell suspensions in phosphate-buffered 1% glucose for 2 h. Phenol extracts of the cells, before and after incubation, were analyzed for Mtl content by periodate utilization and paper chromatography and for Mtl 1-phosphate content, with Mtl 1-phosphate dehydrogenase. In Staphylococcus aureus Towler, the content of Mtl increased from a 0-h value of less than 2.4 to 16 mumol/g (dry weight) after incubation, and the level of Mtl 1-phosphate increased from a 0-h value of 1 to 8 mumol/g. The identification of Mtl was confirmed as the per-O-acetyl ester by gas-liquid chromatography and as the per-O-methyl ether by mass spectrometry. Also tested were 5 additional S. aureus strains and 32 coagulase-negative staphylococcal strains. All strains accumulated Mtl, even those strains that could not utilize exogenous Mtl during aerobic growth, usually in the range 4 to 25 mumol/g. Furthermore, three strains accumulated very high Mtl levels. Bacteria from several other genera were tested, and some were found to accumulate low to moderate levels of Mtl under similar incubation conditions. The metabolic conversion of glucose to intracellular Mtl, probably via Mtl 1-phosphate, is a common feature of staphylococci and also occurs in some other bacteria.  相似文献   

8.
To achieve high mannitol production by Lactococcus lactis, the mannitol 1-phosphatase gene of Eimeria tenella and the mannitol 1-phosphate dehydrogenase gene mtlD of Lactobacillus plantarum were cloned in the nisin-dependent L. lactis NICE overexpression system. As predicted by a kinetic L. lactis glycolysis model, increase in mannitol 1-phosphate dehydrogenase and mannitol 1-phosphatase activities resulted in increased mannitol production. Overexpression of both genes in growing cells resulted in glucose-mannitol conversions of 11, 21, and 27% by the L. lactis parental strain, a strain with reduced phosphofructokinase activity, and a lactate dehydrogenase-deficient strain, respectively. Improved induction conditions and increased substrate concentrations resulted in an even higher glucose-to-mannitol conversion of 50% by the lactate dehydrogenase-deficient L. lactis strain, close to the theoretical mannitol yield of 67%. Moreover, a clear correlation between mannitol 1-phosphatase activity and mannitol production was shown, demonstrating the usefulness of this metabolic engineering approach.  相似文献   

9.
The distribution of carbon flux at the pyruvate node was investigated in Lactococcus lactis under anaerobic conditions with mutant strains having decreased lactate dehydrogenase activity. Strains previously selected by random mutagenesis by H. Boumerdassi, C. Monnet, M. Desmazeaud, and G. Corrieu (Appl. Environ. Microbiol. 63, 2293-2299, 1997) were found to have single punctual mutations in the ldh gene and presented a high degree of instability. The strain L. lactis JIM 5711 in which lactate dehydrogenase activity was diminished to less than 30% of the wild type maintained homolactic metabolism. This was due to an increase in the intracellular pyruvate concentration, which ensures the maintained flux through the lactate dehydrogenase. Pyruvate metabolism was linked to the flux limitation at the level of glyceraldehyde-3-phosphate dehydrogenase, as previously postulated for the parent strain (C. Garrigues, P. Loubière, N. D. Lindley, and M. Cocaign-Bousquet (1997) J. Bacteriol. 179, 5282-5287, 1997). However, a strain (L. lactis JIM 5954) in which the ldh gene was interrupted reoriented pyruvate metabolism toward mixed metabolism (production of formate, acetate, and ethanol), though the glycolytic flux was not strongly diminished. Only limited production of acetoin occurred despite significant overflow of pyruvate. Intracellular metabolite profiles indicated that the in vivo glyceraldehyde-3-phosphate dehydrogenase activity was no longer flux limiting in the Deltaldh strain. The shift toward mixed acid fermentation was correlated with the lower intracellular trioses phosphate concentration and diminished allosteric inhibition of pyruvate formate lyase.  相似文献   

10.
To enhance ethanol production in Thermoanaerobacterium aotearoense, the lactate dehydrogenase (ldh) gene, which is responsible for lactic acid production in a key branch pathway, was successfully disrupted via homologous recombination. ldh-up and ldh-down were designed and amplified based on JW/SL-YS485-AY 278026, and they were subsequently used as homologous fragments with an inserted erythromycin resistance gene to construct the targeted vector based on pBLUESCRIPT II SK(+). Southern hybridization and PCR-based assay definitely confirmed that the ldh gene in the Δldh mutant was disrupted by the insertion of the erythromycin resistance gene. Compared with the wild type, the Δldh mutant exhibited increases of 31.0% and 31.4% in cell yield under glucose and xylose cultivation, respectively, probably because knocking out the ldh gene results in increased acetate and ATP levels. Knockout of lactate dehydrogenase produced 2.37- and 2.1-fold increases in the yield of ethanol (mole/mole substrate) under glucose and xylose cultivation, respectively. Moreover, no lactic acid was detected in Δldh mutant fermentation mixtures (detection limit of HPLC: 0.5 mM), but lactic acid was readily detected for growth of the wild-type strain on both glucose and xylose, with final concentrations up to 59.24 mM and 56.06 mM, respectively. The success of this process thoroughly demonstrates the methodological possibility of gene knockout through homologous recombination in Thermoanaerobacterium.  相似文献   

11.
12.
Candida magnoliae HH-01, a yeast strain that is currently used for the industrial production of mannitol, has the highest mannitol production ever reported for a mannitol-producing microorganism. However, when the fructose concentration exceeds 150 g/L, the volumetric mannitol production rate decreases because of a lag in mannitol production, and the yield decreases as a result of the formation of side products. In fed-batch culture, the volumetric production rate and mannitol yield from fructose vary substantially with the fructose concentration and are maximal at a controlled fructose concentration of 50 g/L. In continuous feeding experiments, the maximum mannitol yield was 85% (g/g) at a glucose/fructose feeding ratio of 1/20. A high glucose concentration in the production phase resulted in the formation of ethanol followed by a decrease in yield and productivity. NAD(P)H-dependent mannitol dehydrogenase was purified to homogeneity from C. magnoliae. In vitro, mannitol dehydrogenase was inhibited by increasing ethanol concentration. Mannitol product was also found to be inhibitory with a K(i) of 183 mM. Under optimum conditions, a final mannitol production of 213 g/L was obtained from 250 g fructose/L after 110 h.  相似文献   

13.
A series of mutant strains of Lactococcus lactis were constructed with lactate dehydrogenase (LDH) activities ranging from below 1% to 133% of the wild-type activity level. The mutants with 59% to 133% of lactate dehydrogenase activity had growth rates similar to the wild-type and showed a homolactic pattern of fermentation. Only after lactate dehydrogenase activity was reduced ninefold compared to the wild-type was the growth rate significantly affected, and the ldh mutants started to produce mixed-acid products (formate, acetate, and ethanol in addition to lactate). Flux control coefficients were determined and it was found that lactate dehydrogenase exerted virtually no control on the glycolytic flux at the wild-type enzyme level and also not on the flux catalyzed by the enzyme itself, i.e. on the lactate production. As expected, the flux towards the mixed-acid products was strongly enhanced in the strain deleted for lactate dehydrogenase. What is more surprising is that the enzyme had a strong negative control ( CLDHJF1 =-1.3) on the flux to formate at the wild-type level of lactate dehydrogenase. Furthermore, we showed that L. lactis has limited excess of capacity of lactate dehydrogenase, only 70% more than needed to catalyze the lactate flux in the wild-type cells.  相似文献   

14.
We have examined the metabolic consequences of knocking out the two ldh genes in Lactobacillus plantarum using 13C nuclear magnetic resonance. Unlike its wild-type isogenic progenitor, which produced lactate as the major metabolite under all conditions tested, ldh null strain TF103 mainly produced acetoin. A variety of secondary end products were also found, including organic acids (acetate, succinate, pyruvate, and lactate), ethanol, 2,3-butanediol, and mannitol.  相似文献   

15.
Mutants of Escherichia coli K-12 were isolated which lack the normal phosphotransferase system-dependent catabolic pathway for D-mannitol (Mtl). In some mutants the pts genes for the general proteins enzyme I and histidine protein of the phosphoenolpyruvate-dependent carbohydrate phosphotransferase systems were deleted. Other mutants expressed truncated mannitol-specific enzymes II (II(Mtl)) which lacked the IIA(Mtl) or IIBA(Mtl) domain(s), and the mtlA genes originated either from E. coli K-12 or from Klebsiella pneumoniae 1033-5P14. The dalD gene from Klebsiella oxytoca M5a1 was cloned on single-copy plasmids and transformed into the strains described above. This gene encodes an NAD-dependent D-arabinitol dehydrogenase (DalD) which converts D-arabinitol into D-xylulose and also converts D-mannitol into D-fructose. The different strains were used to isolate mutations which allow efficient transport of mannitol through the nonphosphorylated II(Mtl) complexes by selecting for growth on this polyhydric alcohol. More than 40 different mutants were analyzed to determine their ability to grow on mannitol, as well as their ability to bind and transport free mannitol and, after restoration of the missing domain(s), their ability to phosphorylate mannitol. Four mutations were identified (E218A, E218V, H256P, and H256Y); all of these mutations are located in the highly conserved loop 5 of the IIC membrane-bound transporter, and two are located in its GIHE motif. These mutations were found to affect the various functions in different ways. Interestingly, in the presence of all II(Mtl) variants, whether they were in the truncated form or in the complete form, in the phosphorylated form or in the nonphosphorylated form, and in the wild-type form or in the mutated form, growth occurred on the low-affinity analogue D-arabinitol with good efficiency, while only the uncoupled mutated forms transported mannitol at a high rate.  相似文献   

16.
17.
《Trends in biotechnology》2023,41(6):745-749
Mannitol is a readily accessible component of seaweed with higher energy content than glucose, making it a promising feedstock for biomanufacturing. Microorganisms have been engineered for converting mannitol into various bioproducts. Microbial strain discovery and synthetic biology approach will advance biomanufacturing using mannitol and other complex components of marine biomass.  相似文献   

18.
The Mtl gene from Cryptococcus neoformans, which confers the ability of Saccharomyces cerevisiae Sc4l YJO to grow on mannitol with substantial NAD-dependent mannitol dehydrogenase activity, was identified. Purifications and characterizations of this enzyme show that it is found in polyploid strain BB1, and the peptide sequence of the enzyme helped identify the saccharomyces gene encoding this mannitol dehydrogenase activity. On the other hand, the Mtl gene of C. neoformans encodes a 346-amino-acid protein which is not mannitol dehydrogenase but a regulatory element which is active in a heterologous fungus.  相似文献   

19.
Mutants of Escherichia coli deficient in the fermentative NAD-linked lactate dehydrogenase (ldh) have been isolated. These mutants showed no growth defects under anaerobic conditions unless present together with a defect in pyruvate formate lyase (pfl). Double mutants (pfl ldh) were unable to grow anaerobically on glucose or other sugars even when supplemented with acetate, whereas pfl mutants can do so. The ldh mutation was found to map at 30.5 min on the E. coli chromosome. The ldh mutant FMJ39 showed no detectable lactate dehydrogenase activity and produced no lactic acid from glucose under anaerobic conditions as estimated by in vivo nuclear magnetic resonance measurements. We also found that in wild-type strains the fermentative lactate dehydrogenase was conjointly induced by anaerobic conditions and an acidic pH. Despite previous findings that phosphate concentrations affect the proportion of lactic acid produced during fermentation, we were unable to find any intrinsic effect of phosphate on lactate dehydrogenase activity, apart from the buffering effect of this ion.  相似文献   

20.
We have isolated a Kluyveromyces lactis mutant unable to grow on all respiratory carbon sources with the exception of lactate. Functional complementation of this mutant led to the isolation of KlSDH1, the gene encoding the flavoprotein subunit of the succinate dehydrogenase (SDH) complex, which is essential for the aerobic utilization of carbon sources. Despite the high sequence conservation of the SDH genes in Saccharomyces cerevisiae and K. lactis, they do not have the same relevance in the metabolism of the two yeasts. In fact, unlike SDH1, KlSDH1 was highly expressed under both fermentative and nonfermentative conditions. In addition to this, but in contrast with S. cerevisiae, K. lactis strains lacking KlSDH1 were still able to grow in the presence of lactate. In these mutants, oxygen consumption was one-eighth that of the wild type in the presence of lactate and was normal with glucose and ethanol, indicating that the respiratory chain was fully functional. Northern analysis suggested that alternative pathway(s), which involves pyruvate decarboxylase and the glyoxylate cycle, could overcome the absence of SDH and allow (i) lactate utilization and (ii) the accumulation of succinate instead of ethanol during growth on glucose.  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