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 78 毫秒
1.
The nutritionally versatile soil bacterium Acinetobacter baylyi ADP1 copes with salt stress by the accumulation of compatible solutes, a strategy that is widespread in nature. This bacterium synthesizes the sugar alcohol mannitol de novo in response to osmotic stress. In a previous study, we identified MtlD, a mannitol‐1‐phosphate dehydrogenase, which is essential for mannitol biosynthesis and which catalyses the first step in mannitol biosynthesis, the reduction of fructose‐6‐phosphate (F‐6‐P) to the intermediate mannitol‐1‐phosphate (Mtl‐1‐P). Until now, the identity of the second enzyme, the phosphatase that catalyses the dephosphorylation of Mtl‐1‐P to mannitol, was elusive. Here we show that MtlD has a unique sequence among known mannitol‐1‐phosphate dehydrogenases with a haloacid dehalogenase (HAD)‐like phosphatase domain at the N‐terminus. This domain is indeed shown to have a phosphatase activity. Phosphatase activity is strictly Mg2+ dependent. Nuclear magnetic resonance analysis revealed that purified MtlD catalyses not only reduction of F‐6‐P but also dephosphorylation of Mtl‐1‐P. MtlD of A. baylyi is the first bifunctional enzyme of mannitol biosynthesis that combines Mtl‐1‐P dehydrogenase and phosphatase activities in a single polypeptide chain. Bioinformatic analysis revealed that the bifunctional enzyme is widespread among Acinetobacter strains but only rarely present in other phylogenetic tribes.  相似文献   

2.
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   

3.
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   

4.
The mtl operon of Klebsiella pneumoniae KAY2026 (formerly Aerobacter aerogenes 1033-5P14) was shown to contain as the promoter-proximal gene mtlA, encoding a D-mannitol-specific enzyme II transporter (IICBA(Mtl)). This gene is followed by mtlD, coding for a mannitol-1-phosphate dehydrogenase (MtlD, 382 amino acid residues), and mtlR (MtlR, 195 amino acid residues) coding for a putative repressor, gene mtlR overlaps the termination codon of mtlD. The DNA and protein sequences are highly similar to the corresponding genes (81% identical bp) and proteins (79-85% identical amino acids) of Escherichia coli K-12. A truncated form of MtlD lacking the 162 C-terminal amino acid residues still shows 10% dehydrogenase activity which may explain the controversy in the literature concerning the properties of mannitol-phosphate and other medium-length dehydrogenases.  相似文献   

5.
Vibrio cholerae utilizes mannitol through an operon of the phosphoenolpyruvate-dependent phosphotransferase (PTS) type. A gene, mtlD, encoding mannitol-1-phosphate dehydrogenase was identified within the 3.9 kb mannitol operon of V. cholerae. The mtlD gene was cloned from V. cholerae O395, and the recombinant enzyme was functionally expressed in E. coli as a 6×His-tagged protein and purified to homogeneity. The recombinant protein is a monomer with a molecular mass of 42.35 kDa. The purified recombinant MtlD reduced fructose 6-phosphate (F6P) using NADH as a cofactor with a K(m) of 1.54 +/- 0.1 mM and V(max) of 320.8 +/- 7.81 micronmol/min/mg protein. The pH and temperature optima for F6P reduction were determined to be 7.5 and 37°C, respectively. Using quantitative real-time PCR analysis, mtlD was found to be constitutively expressed in V. cholerae, but the expression was up-regulated when grown in the presence of mannitol. The MtlD expression levels were not significantly different between V. cholerae O1 and non-O1 strains.  相似文献   

6.
Mannitol metabolism in fungi is thought to occur through a mannitol cycle first described in 1978. In this cycle, mannitol 1-phosphate 5-dehydrogenase (EC 1.1.1.17) was proposed to reduce fructose 6-phosphate into mannitol 1-phosphate, followed by dephosphorylation by a mannitol 1-phosphatase (EC 3.1.3.22) resulting in inorganic phosphate and mannitol. Mannitol would be converted back to fructose by the enzyme mannitol dehydrogenase (EC 1.1.1.138). Although mannitol 1-phosphate 5-dehydrogenase was proposed as the major biosynthetic enzyme and mannitol dehydrogenase as a degradative enzyme, both enzymes catalyze their respective reverse reactions. To date the cycle has not been confirmed through genetic analysis. We conducted enzyme assays that confirmed the presence of these enzymes in a tobacco isolate of Alternaria alternata. Using a degenerate primer strategy, we isolated the genes encoding the enzymes and used targeted gene disruption to create mutants deficient in mannitol 1-phosphate 5-dehydrogenase, mannitol dehydrogenase, or both. PCR analysis confirmed gene disruption in the mutants, and enzyme assays demonstrated a lack of enzymatic activity for each enzyme. GC-MS experiments showed that a mutant deficient in both enzymes did not produce mannitol. Mutants deficient in mannitol 1-phosphate 5-dehydrogenase or mannitol dehydrogenase alone produced 11.5 and 65.7 %, respectively, of wild type levels. All mutants grew on mannitol as a sole carbon source, however, the double mutant and mutant deficient in mannitol 1-phosphate 5-dehydrogenase grew poorly. Our data demonstrate that mannitol 1-phosphate 5-dehydrogenase and mannitol dehydrogenase are essential enzymes in mannitol metabolism in A. alternata, but do not support mannitol metabolism operating as a cycle.  相似文献   

7.
The alpha-aminoadipate pathway for the biosynthesis of lysine is present only in fungi and euglena. Until now, this unique metabolic pathway has never been investigated in the opportunistic fungal pathogens Candida albicans, Cryptococcus neoformans, and Aspergillus fumigatus. Five of the eight enzymes (homocitrate synthase, homoisocitrate dehydrogenase, alpha-aminoadipate reductase, saccharopine reductase, and saccharopine dehydrogenase) of the alpha-aminoadipate pathway and glucose-6-phosphate dehydrogenase, a glycolytic enzyme used as a control, were demonstrated in wild-type cells of these organisms. All enzymes were present in Saccharomyces cerevisiae and the pathogenic organisms except C. neoformans 32608 serotype C, which exhibited no saccharopine reductase activity. The levels of enzyme activity varied considerably from strain to strain. Variation among organisms was also observed for the control enzyme. Among the pathogens, C. albicans exhibited much higher homocitrate synthase, homoisocitrate dehydrogenase, and alpha-aminoadipate reductase activities. Seven lysine auxotrophs of C. albicans and one of Candida tropicalis were characterized biochemically to determine the biochemical blocks and gene-enzyme relationships. Growth responses to alpha-aminoadipate- and lysine-supplemented media, accumulation of alpha-aminoadipate semialdehyde, and the lack of enzyme activity revealed that five of the mutants (WA104, WA153, WC7-1-3, WD1-31-2, and A5155) were blocked at the alpha-aminoadipate reductase step, two (STN57 and WD1-3-6) were blocked at the saccharopine dehydrogenase step, and the C. tropicalis mutant (X-16) was blocked at the saccharopine reductase step. The cloned LYS1 gene of C. albicans in the recombinant plasmid YpB1078 complemented saccharopine dehydrogenase (lys1) mutants of S. cerevisiae and C. albicans. The Lys1+ transformed strains exhibited significant saccharopine dehydrogenase activity in comparison with untransformed mutants. The cloned LYS1 gene has been localized on a 1.8-kb HindIII DNA insert of the recombinant plasmid YpB1041RG1. These results established the gene-enzyme relationship in the second half of the alpha-aminoadipate pathway. The presence of this unique pathway in the pathogenic fungi could be useful for their rapid detection and control.  相似文献   

8.
Enzyme IIIMtl is part of the mannitol phosphotransferase system of Enterococcus faecalis. It is phosphorylated in a reaction sequence requiring enzyme I and heat-stable phosphocarrier protein (HPr). The phospho group is transferred from enzyme IIIMtl to enzyme IIMtl, which then catalyzes the uptake and concomitant phosphorylation of mannitol. The internalized mannitol-1-phosphate is oxidized to fructose-6-phosphate by mannitol-1-phosphate dehydrogenase. In this report we describe the cloning of the mtlF and mtlD genes, encoding enzyme IIIMtl and mannitol-1-phosphate dehydrogenase of E. faecalis, by a complementation system designed for cloning of gram-positive phosphotransferase system genes. The complete nucleotide sequences of mtlF, mtlD, and flanking regions were determined. From the gene sequences, the primary translation products are deduced to consist of 145 amino acids (enzyme IIIMtl) and 374 amino acids (mannitol-1-phosphate dehydrogenase). Amino acid sequence comparison confirmed a 41% similarity of E. faecalis enzyme IIIMtl to the hydrophilic enzyme IIIMtl-like portion of enzyme IIMtl of Escherichia coli and 45% similarity to enzyme IIIMtl of Staphylococcus carnosus. The putative N-terminal NAD+ binding domain of mannitol-1-phosphate dehydrogenase of E. faecalis shows a high degree of similarity with the N terminus of E. coli mannitol-1-phosphate dehydrogenase (T. Davis, M. Yamada, M. Elgort, and M. H. Saier, Jr., Mol. Microbiol. 2:405-412, 1988) and the N-terminal part of the translation product of S. carnosus mtlD, which was also determined in this study. There is 40% similarity between the dehydrogenases of E. faecalis and E. coli over the whole length of the enzymes. The organization of mannitol-specific genes in E. faecalis seems to be similar to the organization in S. carnosus. The open reading frame for enzyme IIIMtl E. faecalis is followed by a stem-loop structure, analogous to a typical Rho-independent terminator. We conclude that the mannitol-specific genes are organized in an operon and that the gene order is mtlA orfX mtlF mtlD.  相似文献   

9.
The solution structure of the cytoplasmic B domain of the mannitol (Mtl) transporter (II(Mtl)) from the mannitol branch of the Escherichia coli phosphotransferase system has been solved by multidimensional NMR spectroscopy with extensive use of residual dipolar couplings. The ordered IIB(Mtl) domain (residues 375-471 of II(Mtl)) consists of a four-stranded parallel beta-sheet flanked by two helices (alpha(1) and alpha(3)) on one face and helix alpha(2) on the opposite face with a characteristic Rossmann fold comprising two right-handed beta(1)alpha(1)beta(2) and beta(3)alpha(2)beta(4) motifs. The active site loop is structurally very similar to that of the eukaryotic protein tyrosine phosphatases, with the active site cysteine (Cys-384) primed in the thiolate state (pK(a) < 5.6) for nucleophilic attack at the phosphorylated histidine (His-554) of the IIA(Mtl) domain through stabilization by hydrogen bonding interactions with neighboring backbone amide groups at positions i + 2/3/4 from Cys-384 and with the hydroxyl group of Ser-391 at position i + 7. Modeling of the phosphorylated state of IIB(Mtl) suggests that the phosphoryl group can be readily stabilized by hydrogen bonding interactions with backbone amides in the i + 2/4/5/6/7 positions as well as with the hydroxyl group of Ser390 at position i + 6. Despite the absence of any significant sequence identity, the structure of IIB(Mtl) is remarkably similar to the structures of bovine protein tyrosine phosphatase (which contains two long insertions relative to IIB(Mtl)) and the cytoplasmic B component of enzyme II(Chb), which fulfills an analogous role to IIB(Mtl) in the N,N'-diacetylchitobiose branch of the phosphotransferase system. All three proteins utilize a cysteine residue in the nucleophilic attack of a phosphoryl group covalently bound to another protein.  相似文献   

10.
The mannitol permease (EII(Mtl)) from Escherichia coli couples mannitol transport to phosphorylation of the substrate. Renewed topology prediction of the membrane-embedded C domain suggested that EII(Mtl) contains more membrane-embedded segments than the six proposed previously on the basis of a PhoA fusion study. Cysteine accessibility was used to confirm this notion. Since cysteine 384 in the cytoplasmic B domain is crucial for the phosphorylation activity of EII(Mtl), all cysteine mutants contained this activity-linked cysteine residue in addition to those introduced for probing the membrane topology of the protein. To distinguish between the activity-linked cysteine and the probed cysteine, either trypsin was used to specifically digest the two cytoplasmic domains (A and B), thereby removing Cys384, or Cys384 was protected by phosphorylation from alkylation by N-ethylmaleimide (NEM). Our data show that upon phosphorylation EII(Mtl) undergoes major conformational changes, whereby residues in the putative first cytoplasmic loop become accessible to NEM. Other residues in this loop were accessible to NEM in intact cells and inside-out membrane vesicles, but cysteine residues at these positions only reacted with the membrane-impermeable sulfhydryl reagent from the periplasmic side of the protein. These and other results suggest that the predicted loop between TM2 and TM3 may fold back into the membrane and form part of the translocation path.  相似文献   

11.
Mannitol dehydrogenase, a mannitol:mannose 1-oxidoreductase, constitutes the first enzymatic step in the catabolism of mannitol in nonphotosynthetic tissues of celery (Apium graveolens L.). Endogenous regulation on the enzyme activity in response to environmental cues is critical in modulating tissue concentration of mannitol, which, importantly, contribute to stress tolerance of celery. The enzyme was purified to homogeneity from celery suspension cultures grown on D-mannitol as the carbon source. Mannitol dehydrogenase was purified 589-fold to a specific activity of 365 mumol h-1 mg-1 protein with a 37% yield of enzyme activity present in the crude extract. A highly efficient and simple purification protocol was developed involving polyethylene glycol fractionation, diethylaminoethyl-anion-exchange chromatography, and NAD-agarose affinity chromatography using NAD gradient elution. Sodium dodecylsulfate gel electrophoresis of the final preparation revealed a single 40-kD protein. The molecular mass of the native protein was determined to be approximately 43 kD, indicating that the enzyme is a monomer. Polyclonal antibodies raised against the enzyme inhibited enzymatic activity of purified mannitol dehydrogenase. Immunoblots of crude protein extracts from mannitol-grown celery cells and sink tissues of celery, celeriac, and parsley subjected to sodium dodecyl sulfate gel electrophoresis showed a single major immuno-reactive 40-kD protein.  相似文献   

12.
Reversal of the Mannitol-Sorbitol Diauxie in Escherichia coli   总被引:16,自引:12,他引:4  
In Escherichia coli K-12 the proteins involved in the dissimilation of mannitol and sorbitol are specified by two separate gene clusters. The mannitol cluster appears to consist of a regulatory gene mtlC, a gene mtlA coding an enzyme II complex of the phosphoenolpyruvate phosphotransferase system, and a gene mtlD coding a mannitol-1-phosphate dehydrogenase. Three corresponding genes, sblC, sblA, and sblD, exist for the sorbitol pathway. In both pathways the hexitol captured from the medium and delivered into the cytoplasm as a phosphorylated compound is dehydrogenated to fructose-6-phosphate. The enzyme II complex for sorbitol is able to catalyze the phosphorylation also of mannitol if this substrate is present at high concentrations. Consequently mtlA(-) mutants lacking the enzyme II complex for mannitol can grow on mannitol either if the sorbitol phosphorylating system is preinduced by sorbitol or if mtlA is suppressed by a mutation of sblC to constitutivity. In wild-type cells, the induction of the enzymes in the mannitol pathway and dissimilation of the substrate are not prevented by glucose. The sorbitol system, however, is sensitive to glucose and to mannitol as well. In the suppressed strains (mtlA(-), sblC(c)) in which mannitol is utilized through the sorbitol enzyme, glucose becomes effective in restraining the consumption of mannitol, causing a definite diauxie. Moreover, in a mixture of mannitol and sorbitol, the latter is utilized preferentially. This reversal of normal diauxic pattern is consequent to the fact that the enzyme II complex for sorbitol has relatively poor affinity for mannitol.  相似文献   

13.
A Varma  K J Kwon-Chung 《Gene》1999,232(2):155-163
The GPD gene encoding glyceraldehyde-3-phosphate dehydrogenase was isolated from Cryptococcus neoformans, a heterobasidiomycetous yeast that is pathogenic to humans. The gene contains 11 introns, differing from the conserved intron positions found in the GPD genes of Basidiomycetes. The predicted amino-acid sequence of this gene is extremely similar to that reported from GPD proteins of other basidiomycetes. The promoter region of the C. neoformans GPD gene was similar to those of other basidiomycetes. Plasmid constructs containing up to 1600 base pairs upstream of the native GPD open reading frame were used to express either the native URA5 gene in a ura5 mutant or the heterologous hphI gene (a bacterial gene that confers resistance to the aminoglycoside hygromycin) in a wild-type strain of C. neoformans. Transformation frequencies resulting from the plasmid-borne Gpdp::URA5 gene were at levels similar to those of the native URA5, which suggested that all the sequences necessary for proper expression were present. Transformation frequencies using the Gpdp::hphI gene constructs were poor. However, addition of DNA sequences flanking the 3'-end of an native C. neoformans gene significantly improved the transformation frequencies resulting from the expression of the heterologous hphI gene.  相似文献   

14.
Mannitol kinase and mannitol-1-phosphate dehydrogenase activities were detected in two Micromonospora isolates. The presence of these enzyme activities indicates that mannitol is catabolized first to mannitol-1-phosphate and then to fructose-6-phosphate. Mannitol-oxidizing enzymes were also surveyed in representative species of four other genera of actinomycetes. Mannitol-1-phosphate dehydrogenase was detected in cell-free extracts of Streptomyces lactamdurans. In contrast, cell-free extracts of Mycobacterium smegmatis, Nocardia erythrophila, Streptomyces lavendulae, and Actinoplanes missouriensis contained mannitol dehydrogenase activity but no detectable mannitol-1-phosphate dehydrogenase activity. The mannitol dehydrogenase activities in the latter species support the operation of a pathway for catabolism of mannitol that involves the oxidation of mannitol to fructose, followed by phosphorylation to fructose-6-phosphate.  相似文献   

15.
Mannitol kinase and mannitol-1-phosphate dehydrogenase activities were detected in two Micromonospora isolates. The presence of these enzyme activities indicates that mannitol is catabolized first to mannitol-1-phosphate and then to fructose-6-phosphate. Mannitol-oxidizing enzymes were also surveyed in representative species of four other genera of actinomycetes. Mannitol-1-phosphate dehydrogenase was detected in cell-free extracts of Streptomyces lactamdurans. In contrast, cell-free extracts of Mycobacterium smegmatis, Nocardia erythrophila, Streptomyces lavendulae, and Actinoplanes missouriensis contained mannitol dehydrogenase activity but no detectable mannitol-1-phosphate dehydrogenase activity. The mannitol dehydrogenase activities in the latter species support the operation of a pathway for catabolism of mannitol that involves the oxidation of mannitol to fructose, followed by phosphorylation to fructose-6-phosphate.  相似文献   

16.
Mutants of Escherichia coli K-12 defective in the mannitol-specific enzyme II complex of the phosphoenolpyruvate phosphotransferase system (PTS) or lacking mannitol-1-phosphate dehydrogenase have been isolated. These mutants fail only to grow on mannitol. Growth of the dehydrogenase-negative mutant on casein hydrolysate can be abruptly inhibited by exposure to mannitol. A mutant with constitutive expression of both of these enzymes has also been isolated. All three mutations are clustered in a region represented at min 71 of the Taylor map. In a mutant with less than 5% of the activity of enzyme I of the PTS, both the enzyme II complex and the dehydrogenase remain inducible by mannitol. In the mutant defective in the enzyme II complex, mannitol is able to induce the dehydrogenase. Thus, mannitol, rather than its phosphorylated product, seems to be the inducer.  相似文献   

17.
以蔗糖为底物利用重组大肠杆菌合成甘露醇   总被引:1,自引:0,他引:1  
【目的】异型发酵乳酸菌可利用胞内产生的甘露醇脱氢酶将果糖高效转化为甘露醇,但果糖作为底物相对昂贵,不利于工业化生产。为了降低生产成本,必须选择廉价的底物。蔗糖相对便宜,并且大量存在于自然界中,能够被重组大肠杆菌利用产生甘露醇。蔗糖水解酶(Sucrose hydrolase)和甘露醇脱氢酶(Mannitol dehydrogenase)是发酵生产甘露醇中催化蔗糖转化成甘露醇的关键酶,构建蔗糖水解酶和甘露醇脱氢酶共表达菌株并进行相关研究是本文的主旨。【方法】利用PCR方法分别从植物乳杆菌(Lactobacillus plantarum)和布氏乳杆菌(Lactobacillus buchneri)基因组DNA中获得sac A和mdh基因,得到大小分别为1 502 bp和1 032 bp的目的基因,经序列分析后将其连接到表达载体p ET-28a(+)上,得到重组表达载体p ET28a-sac A-mdh。将重组质粒转化到大肠杆菌BL21(DE3)中,并用SDS-PAGE分析目的蛋白的表达情况并测定其酶活。【结果】SDS-PAGE显示表达蛋白的大小亚基分子量分别为55.1 k D和37.8 k D,与预期分子量一致,实现sac A和mdh基因的表达。蔗糖水解酶和甘露醇脱氢酶酶活分别为25.78 U/m L和14.56 U/m L。对重组菌株BL21(DE3)/p ET28a-sac A-mdh进行发酵条件优化,甘露醇质量浓度达到45.19 g/L,总糖转化率为37.66%。【结论】与乳酸菌利用蔗糖发酵生产甘露醇相比,产量提高了6倍,且具有发酵周期短、稳定性高等优点,菌株的成功构建为甘露醇工业化生产奠定了基础。  相似文献   

18.
Mutant strains of Pseudomonas aeruginosa PAO were isolated on the basis of their inability to utilize mannitol as sole carbon source for growth. Four linkage groups (I through IV) among these mutant strains were resolved by two-factor crosses using the general transducing phage F116, and the strains appeared to contain point mutations as evidenced by ability to give rise to spontaneous revertants with wild phenotype on mannitol minimal agar. Group I strains were affected only in ability to grow on mannitol; all were deficient in inducible mannitol dehydrogenase activity, and all but one were deficient in inducible mannitol transport activity. Fructokinase was induced in group I strains and in wild-type bacteria during growth in the presence of mannitol but not fructose, indicating the presence of a pathway specific for endogenously generated fructose. Cells grown on fructose contained phosphoenolpyruvate:fructose-1-phosphotransferase activity, and mannitol-grown cells contained a lower level of this activity. Group II mutants were deficient in constitutive phosphoglucoisomerase, failed to grow on mannitol, grew very slowly on glycerol and fructose, but grew normally on glucose and gluconate. Group III strains were deficient in both nicotinamide adenine dinucleotide- and nicotinamide adenine dinucleotide phosphate-linked glucose-6-phosphate dehydrogenase activities that reside in a single enzyme species. 6-Phosphogluconate appeared to be the inductive effector for this enzyme, which was not required for aerobic growth on glucose or gluconate. A single mannitol-negative mutant in group IV also failed to grow on glycerol and glucose, but no biochemical lesion was identified.  相似文献   

19.
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   

20.
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