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 140 毫秒
1.
A monomeric NADP-dependent isocitrate dehydrogenase from the multicellular prokaryote Streptomyces avermitilis MA-4680 (SaIDH) was heteroexpressed in Escherichia coli, and the His-tagged enzyme was further purified to homogeneity. The molecular weight of SaIDH was about 80 kDa which is typical for monomeric isocitrate dehydrogenases. Structure-based sequence alignment reveals that the deduced amino acid sequence of SaIDH shows high sequence identity with known momomeric isocitrate dehydrogenase, and the coenzyme, substrate and metal ion binding sites are completely conserved. The optimal pH and temperature of SaIDH were found to be pH 9.4 and 45°C, respectively. Heat-inactivation studies showed that heating for 20 min at 50°C caused a 50% loss in enzymatic activity. In addition, SaIDH was absolutely specific for NADP+ as electron acceptor. Apparent K m values were 4.98 μM for NADP+ and 6,620 μM for NAD+, respectively, using Mn2+ as divalent cation. The enzyme performed a 33,000-fold greater specificity (k cat/K m) for NADP+ than NAD+. Moreover, SaIDH activity was entirely dependent on the presence of Mn2+ or Mg2+, but was strongly inhibited by Ca2+ and Zn2+. Taken together, our findings implicate the recombinant SaIDH is a divalent cation-dependent monomeric isocitrate dehydrogenase which presents a remarkably high cofactor preference for NADP+.  相似文献   

2.
An integrated study of cell growth characteristics, enzyme activities and protein expression patterns was carried out to investigate how the central metabolism of Escherichia coli changes upon knockout of the isocitrate dehydrogenase (ICDH) gene (icd) in the tricarboxylic acid cycle. Deletion of the icd gene led to reduced specific growth rate and reduced specific glucose consumption rate. The reduced specific growth rate in the icd mutant was due mainly to the lower intracellular ATP/ADP ratio as well as to the lower NADPH/NADP+ ratio compared with those in the parent strain. However, the specific carbon dioxide evolution rate was found to be higher in the icd mutant strain compared to the parent E. coli. This may be due to the higher activity of 6-phosphogluconate dehydrogenase, phosphoenol pyruvate carboxykinase and NADP+-dependent malic enzymes. The glyoxylate pathway was also utilized, as evidenced by the significant upregulation of isocitrate lyase and malate synthase activity in the icd mutant E. coli. The appearance of the glyoxylate pathway caused lower acetate production. Of 21 proteins showing altered expression levels, 17 were successfully identified with the aid of MALDI-TOF mass spectrometry. The results showed that the abolition of ICDH activity significantly affected the respiratory system and electron transport chain, as evidenced by the significant downregulation of proteins encoded by the genes nuoE, nuoH, cydA and cyoA in icd mutant E. coli compared to the parent.  相似文献   

3.
Isocitrate dehydrogenase (IDH) is one of the key enzymes in the citric acid cycle, which involves in providing energy and biosynthetic precursors for metabolism. Here, we report for the first time the enzymatic characterization of a monomeric NADP+-dependent IDH from Streptomyces lividans TK54 (SlIDH). The icd gene (GenBank database accession number EU661252) encoding IDH was cloned and overexpressed in Escherichia coli. The molecular mass of SlIDH was about 80 kDa, typical of a monomeric NADP-IDH, and showed high amino acid sequence identity with known monomeric IDHs. The optimal activity of the 6His-tagged SlIDH was found at pH values 8.5 (Mn2+) and 9.0 (Mg2+), and the optimal temperature was around 46 °C. Heat-inactivation studies showed that about 50% SlIDH activity was preserved at 38 °C after 20 min of incubation. The recombinant SlIDH displayed a 62,000-fold (kcat/Km) preference for NADP+ over NAD+ with Mn2+, and a 85,000-fold greater specificity for NADP+ than NAD+ with Mg2+. Therefore, SlIDH is a divalent cation-dependent monomeric IDH with remarkably high coenzyme preference for NADP+.  相似文献   

4.
The Escherichia coli NADP+-dependent isocitrate dehydrogenase (IDH; EC 1.1.1.42), encoded by an icd gene, is a tricarboxylic acid (TCA) cycle enzyme responsible for the oxidative decarboxylation of isocitrate to α-ketoglutarate. In order to examine how the icd gene expression is regulated, an icd-lacZ reporter fusion was constructed. While the icd gene was induced in exponential growth phase, it was repressed in stationary growth phase. Genetic inactivation of an rpoS gene, whose product is an alternative sigma factor, induced the icd gene expression approximately 4.8 times more in the stationary phase and the IDH enzyme activity in the rpoS mutant was 3.2 times higher than that in the wild type, indicating that the RpoS factor acts as a negative regulator of the icd gene expression in the stationary phase.  相似文献   

5.
Intracellular NADPH/NADP+ ratio in cells grown on various production media with different carbon and nitrogen sources had a positive correlation with the thymidine production. To improve thymidine production in a previously engineered E. coli strain, NAD+ kinase was overexpressed in it resulting in the NADPH/NADP+ ratio shifting from 0.184 to 0.267. The [NADH + NADP+]/[NAD+ + NADPH] ratio was, however, not significantly altered. In jar fermentation, 740 mg thymidine l−1 was produced in parental strain, while 940 mg l−1 of thymidine was produced in NAD+ kinase-expressing strain.  相似文献   

6.
The gene encoding isocitrate dehydrogenase (IDH) of Methylococcus capsulatus (McIDH) was cloned and overexpressed in Escherichia coli. The purified enzyme was NAD+-dependent with a thermal optimum for activity at 55–60°C and an apparent midpoint melting temperature (T m) of 70°C. Analytical ultracentrifugation (AUC) revealed a homotetrameric state, and McIDH thus represents the first homotetrameric NAD+-dependent IDH that has been characterized. Based on a structural alignment of McIDH and homotetrameric homoisocitrate dehydrogenase (HDH) from Thermus thermophilus (TtHDH), we identified the clasp-like domain of McIDH as a likely site for tetramerization. McIDH showed moreover, higher sequence identity (48%) to TtHDH than to previously characterized IDHs. Putative NAD+-IDHs with high sequence identity (48–57%) to McIDH were however identified in a variety of bacteria showing that NAD+-dependent IDHs are indeed widespread within the domain, Bacteria. Phylogenetic analysis including these new sequences revealed a close relationship with eukaryal allosterically regulated NAD+-IDH and the subfamily III of IDH was redefined to include bacterial NAD+- and NADP+-dependent IDHs. This apparent relationship suggests that the mitochondrial genes encoding NAD+-IDH are derived from the McIDH-like IDHs.  相似文献   

7.
The in vivo flux through the oxidative branch of the pentose phosphate pathway (oxPPP) in Penicillium chrysogenum was determined during growth in glucose/ethanol carbon-limited chemostat cultures, at the same growth rate. Non-stationary 13C flux analysis was used to measure the oxPPP flux. A nearly constant oxPPP flux was found for all glucose/ethanol ratios studied. This indicates that the cytosolic NADPH supply is independent of the amount of assimilated ethanol. The cofactor assignment in the model of van Gulik et al. (Biotechnol Bioeng 68(6):602–618, 2000) was supported using the published genome annotation of P. chrysogenum. Metabolic flux analysis showed that NADPH requirements in the cytosol remain nearly the same in these experiments due to constant biomass growth. Based on the cytosolic NADPH balance, it is known that the cytosolic aldehyde dehydrogenase in P. chrysogenum is NAD +  dependent. Metabolic modeling shows that changing the NAD + -aldehyde dehydrogenase to NADP + -aldehyde dehydrogenase can increase the penicillin yield on substrate.  相似文献   

8.
In most living organisms, isocitrate dehydrogenases (IDHs) convert isocitrate into ɑ-ketoglutarate (ɑ-KG). Phylogenetic analyses divide the IDH protein family into two subgroups: types I and II. Based on cofactor usage, IDHs are either NAD+-specific (NAD-IDH) or NADP+-specific (NADP-IDH); NADP-IDH evolved from NAD-IDH. Type I IDHs include NAD-IDHs and NADP-IDHs; however, no type II NAD-IDHs have been reported to date. This study reports a novel type II NAD-IDH from the marine bacterium Congregibacter litoralis KT71 (ClIDH, GenBank accession no. EAQ96042). His-tagged recombinant ClIDH was produced in Escherichia coli and purified; the recombinant enzyme was NAD+-specific and showed no detectable activity with NADP+. The K m values of the enzyme for NAD+ were 262.6±7.4 μM or 309.1±11.2 μM with Mg2+ or Mn2+ as the divalent cation, respectively. The coenzyme specificity of a ClIDH Asp487Arg/Leu488His mutant was altered, and the preference of the mutant for NADP+ was approximately 24-fold higher than that for NAD+, suggesting that ClIDH is an NAD+-specific ancestral enzyme in the type II IDH subgroup. Gel filtration and analytical ultracentrifugation analyses revealed the homohexameric structure of ClIDH, which is the first IDH hexamer discovered thus far. A 163-amino acid segment of CIIDH is essential to maintain its polymerization structure and activity, as a truncated version lacking this region forms a non-functional monomer. ClIDH was dependent on divalent cations, the most effective being Mn2+. The maximal activity of purified recombinant ClIDH was achieved at 35°C and pH 7.5, and a heat inactivation experiment showed that a 20-min incubation at 33°C caused a 50% loss of ClIDH activity. The discovery of a NAD+-specific, type II IDH fills a gap in the current classification of IDHs, and sheds light on the evolution of type II IDHs.  相似文献   

9.
Biosynthesis of guanosine 5′-diphosphate-l-fucose (GDP-l-fucose) requires NADPH as a reducing cofactor. In this study, endogenous NADPH regenerating enzymes such as glucose-6-phosphate dehydrogenase (G6PDH), isocitrate dehydrogenase (Icd), and NADP+-dependent malate dehydrogenase (MaeB) were overexpressed to increase GDP-l-fucose production in recombinant Escherichia coli. The effects of overexpression of each NADPH regenerating enzyme on GDP-l-fucose production were investigated in a series of batch and fed-batch fermentations. Batch fermentations showed that overexpression of G6PDH was the most effective for GDP-l-fucose production. However, GDP-l-fucose production was not enhanced by overexpression of G6PDH in the glucose-limited fed-batch fermentation. Hence, a glucose feeding strategy was optimized to enhance GDP-l-fucose production. Fed-batch fermentation with a pH-stat feeding mode for sufficient supply of glucose significantly enhanced GDP-l-fucose production compared with glucose-limited fed-batch fermentation. A maximum GDP-l-fucose concentration of 235.2 ± 3.3 mg l−1, corresponding to a 21% enhancement in the GDP-l-fucose production compared with the control strain overexpressing GDP-l-fucose biosynthetic enzymes only, was achieved in the pH-stat fed-batch fermentation of the recombinant E. coli overexpressing G6PDH. It was concluded that sufficient glucose supply and efficient NADPH regeneration are crucial for NADPH-dependent GDP-l-fucose production in recombinant E. coli.  相似文献   

10.
We studied the physiological effect of the interconversion between the NAD(H) and NADP(H) coenzyme systems in recombinant Saccharomyces cerevisiae expressing the membrane-bound transhydrogenase from Escherichia coli. Our objective was to determine if the membrane-bound transhydrogenase could work in reoxidation of NADH to NAD+ in S. cerevisiae and thereby reduce glycerol formation during anaerobic fermentation. Membranes isolated from the recombinant strains exhibited reduction of 3-acetylpyridine-NAD+ by NADPH and by NADH in the presence of NADP+, which demonstrated that an active enzyme was present. Unlike the situation in E. coli, however, most of the transhydrogenase activity was not present in the yeast plasma membrane; rather, the enzyme appeared to remain localized in the membrane of the endoplasmic reticulum. During anaerobic glucose fermentation we observed an increase in the formation of 2-oxoglutarate, glycerol, and acetic acid in a strain expressing a high level of transhydrogenase, which indicated that increased NADPH consumption and NADH production occurred. The intracellular concentrations of NADH, NAD+, NADPH, and NADP+ were measured in cells expressing transhydrogenase. The reduction of the NADPH pool indicated that the transhydrogenase transferred reducing equivalents from NADPH to NAD+.  相似文献   

11.
The activity of pure calf-liver and Escherichia coli thioredoxin reductases decreased drastically in the presence of NADPH or NADH, while NADP+, NAD+ and oxidized E. coli thioredoxin activated both enzymes significantly, particularly the bacterial one. The loss of activity under reducing conditions was time-dependent, thus suggesting an inactivation process: in the presence of 0.24 mM NADPH the half-lives for the E. coli and calf-liver enzymes were 13.5 and 2 min, respectively. Oxidized E. coli thioredoxin fully protected both enzymes from inactivation, and also promoted their complete reactivation after only 30 min incubation at 30° C. Lower but significant protection and reactivation was also observed with NADP+ and NAD+. EDTA protected thioredoxin reductase from NADPH inactivation to a great degree, thus indicating the participation of metals in the process; EGTA did not protect the enzyme from redox inactivation. Thioredoxin reductase was extensively inactivated by NADPH under aerobic and anaerobic conditions, thus excluding the participation of O2 or oxygen active species in redox inactivation. The loss of thioredoxin reductase activity promoted by NADPH was much faster and complete in the presence of NAD+ glycohydrolase, thus suggesting that inactivation was related to full reduction of the redox-active disulfide. Those results indicate that thioredoxin reductase activity can be modulated in bacteria and mammals by the redox status of NADP(H) and thioredoxin pools, in a similar way to glutathione reductase. This would considerably expand the regulatory potential of the thioredoxin-thioredoxin reductase system with the enzyme being self-regulated by its own substrate, a regulatory protein.Abbreviations DTNB 5,5-dithiobis(2-nitrobenzoate) - EGTA Ethylenglycoltetraacetic Acid - TNB 5-thio-2-nitrobenzoate - Trx Thioredoxin - Trx(SH)2 Reduced Thioredoxin - Trx-S2 Oxidized Thioredoxin  相似文献   

12.
Previously, we constructed a glycerol oxidative pathway-deficient mutant strain of Klebsiella pneumoniae by inactivation of glycerol dehydrogenase (dhaD) to eliminate by-product synthesis during production of 1,3-propanediol (1,3-PD) from glycerol. Although by-product formation was successfully blocked in the resultant strain, the yield of 1,3-PD was not enhanced, probably because dhaD disruption resulted in insufficient regeneration of the cofactor NADH essential for the activity of 1,3-PD oxidoreductase (DhaT). To improve cofactor regeneration, in the present study we overexpressed an NAD+-dependent aldehyde dehydrogenase in the recombinant strain. To this end, an aldehyde dehydrogenase AldHk homologous to E. coli AldH but with NAD+-dependent propionaldehyde dehydrogenase activity was identified in K. pneumoniae. Functional analysis revealed that the substrate specificity of AldHk embraced various aldehydes including propionaldehyde, and that NAD+ was preferred over NADP+ as a cofactor. Overexpression of AldHk in the glycerol oxidative pathway-deficient mutant AK/pVOTHk resulted in a 3.6-fold increase (0.57 g l−1 to 2.07 g l−1) in the production of 3-hydroxypropionic acid (3-HP), and a 1.1-fold enhancement (8.43 g l−1 to 9.65 g l−1) of 1,3-PD synthesis, when glycerol was provided as the carbon source, compared to the levels synthesized by the control strain (AK/pVOT). Batch fermentation using AK/pVOTHk showed a significant increase (to 70%, w/w) in conversion of glycerol to the reductive metabolites, 1,3-PD and 3-HP, with no production of by-products except acetate.  相似文献   

13.
Redox interconversion of glutathione reductase was studiedin situ withS. cerevisiae. The enzyme was more sensitive to redox inactivation in 24 hour-starved cells than in freshly-grown ones. While 5 μM NADPH or 100 μM NADH caused 50% inactivation in normal cells in 30 min, 0.75 μM NADPH or 50 μM NADH promoted a similar effect in starved cells. GSSG reactivated the enzyme previously inactivated by NADPH, ascertaining that the enzyme was subjected to redox interconversion. Low EDTA concentrations fully protected the enzyme from NADPH inactivation, thus confirming the participation of metals in such a process. Extensive inactivation was obtained in permeabilized cells incubated with glucose-6-phosphate or 6-phosphogluconate, in agreement with the very high specific activities of the corresponding dehydrogenases. Some inactivation was also observed with malate, L-lactate, gluconate or isocitrate in the presence of low NADP+ concentrations. The inactivation of yeast glutathione reductase has also been studiedin vivo. The activity decreased to 75% after 2 hours of growth with glucono-δ-lactone as carbon source, while NADPH rose to 144% and NADP+ fell to 86% of their initial values. Greater changes were observed in the presence of 1.5 μM rotenone: enzymatic activity descended to 23% of the control value, while the NADH/NAD+ and NADPH/NADP+ ratios rose to 171% and 262% of their initial values, respectively. Such results indicate that the lowered redox potential of the pyridine nucleotide pool existing when glucono-δ-lactone is oxidized promotesin vivo inactivation of glutathione reductase.  相似文献   

14.
Isocitrate deyhdrogenase (IDH) is a reversible enzyme in the tricarboxylic acid cycle that catalyzes the NAD(P)+-dependent oxidative decarboxylation of isocitrate to α-ketoglutarate (αKG) and the NAD(P)H/CO2-dependent reductive carboxylation of αKG to isocitrate. The IDH gene from Streptococcus mutans was fused with the icd gene promoter from Escherichia coli to initiate its expression in the glutamate auxotrophic strain E. coli Δicd::kanr of which the icd gene has been replaced by kanamycin resistance gene. The expression of S. mutans IDH (SmIDH) may restore the wild-type phenotype of the icd-defective strain on minimal medium without glutamate. The molecular weight of SmIDH was estimated to be 70 kDa by gel filtration chromatography, suggesting a homodimeric structure. SmIDH was divalent cation-dependent and Mn2+ was found to be the most effective cation. The optimal pH of SmIDH was 7.8 and the maximum activity was around 45°C. SmIDH was completely NAD+ dependent and its apparent K m for NAD+ was 137 μM. In order to evaluate the role of the putative phosphorylation site at Ser102 in catalysis, two “stably phosphorylated” mutants were constructed by converting Ser102 into Glu102 or Asp102 in SmIDH to mimick a constitutively phosphorylated state. Meanwhile, the functional roles of another four amino acids (threonine, glycine, alanine and tyrosine) containing variant size of side chains were investigated. The replacement of Asp102 or Glu102 totally inactivated the enzyme, while the S102T, S102G, S102A and S102Y mutants decreased the affinity to isocitrate and only retained 16.0%, 2.8%, 3.3% and 1.1% of the original activity, respectively. These results reveal that Ser102 plays important role in substrate binding and is required for the enzyme function. Also, Ser102 in SmIDH is a potential phosphorylation site, indicating that the ancient NAD-dependent IDHs might be the underlying origin of “phosphorylation mechanism” used by their bacterial NADP-dependent homologs.  相似文献   

15.
Thermoplasma acidophilum utilizes l-rhamnose as a sole carbon source. To determine the metabolic pathway of l-rhamnose in Archaea, we identified and characterized l-rhamnose dehydrogenase (RhaD) in T. acidophilum. Ta0747P gene, which encodes the putative T. acidophilum RhaD (Ta_RhaD) enzyme belonging to the short-chain dehydrogenase/reductase family, was expressed in E. coli as an active enzyme catalyzing the oxidation of l-rhamnose to l-rhamnono-1,4-lactone. Analysis of catalytic properties revealed that Ta_RhaD oxidized l-rhamnose, l-lyxose, and l-mannose using only NADP+ as a cofactor, which is different from NAD+/NADP+-specific bacterial RhaDs and NAD+-specific eukaryal RhaDs. Ta_RhaD showed the highest activity toward l-rhamnose at 60 °C and pH 7. The K m and k cat values were 0.46 mM, 1,341.3 min−1 for l-rhamnose and 0.1 mM, 1,027.2 min−1 for NADP+, respectively. Phylogenetic analysis indicated that branched lineages of archaeal RhaD are quite distinct from those of Bacteria and Eukarya. This is the first report on the identification and characterization of NADP+-specific RhaD.  相似文献   

16.
The malic enzyme-encoding cDNA (GQ372891) from the oleaginous yeast Lipomyces starkeyi AS 2.1560 was isolated, which has an 1719-bp open reading frame flanked by a 290-bp 5′ untranslated sequence and a 92-bp 3′ untranslated sequence. The proposed gene, LsME1, encoded a protein with 572 amino acid residues. The protein presented 58% sequence identity with the malic enzymes from Yarrowia lipolytica CLIB122 and Aspergillus fumigatus Af293. The LsME1 gene was cloned into the vector pMAL-p4x to express a fusion protein (MBP-LsME1) in Escherichia coli TB1. The fusion protein was purified and then cleaved by Factor Xa to give the recombinant LsME1. This purified enzyme took either NAD+ or NADP+ as the coenzyme but preferred NAD+. The K m values for malic acid, NAD+ and NADP+ were 0.85 ± 0.05 mM, 0.34 ± 0.08 mM, and 7.4 ± 0.32 mM, respectively, at pH 7.3.  相似文献   

17.
Methanobacterium thermoautotrophicum (strain Marburg) was found to contain two malate dehydrogenases, which were partially purified and characterized. One was specific for NAD+ and catalyzed the dehydrogenation of malate at approximately one-third of the rate of oxalacetate reduction, and the other could equally well use NAD+ and NADP+ as coenzyme and catalyzed essentially only the reduction of oxalacetate. Via the N-terminal amino acid sequences, the encoding genes were identified in the genome of M. thermoautotrophicum (strain ΔH). Comparison of the deduced amino acid sequences revealed that the two malate dehydrogenases are phylogenetically only distantly related. The NAD+-specific malate dehydrogenase showed high sequence similarity to l-malate dehydrogenase from Methanothermus fervidus, and the NAD(P)+-using malate dehyrogenase showed high sequence similarity to l-lactate dehydrogenase from Thermotoga maritima and l-malate dehydrogenase from Bacillus subtilis. A function of the two malate dehydrogenases in NADPH:NAD+ transhydrogenation is discussed. Received: 29 December 1997 / Accepted: 4 March 1998  相似文献   

18.
19.
20.
Isocitrate dehydrogenase (IDH) activities were measured in mitochondria isolated from aerial parts of 21-day-old spruce (Picea abies L. Karst.) seedlings. Mitochondria were purified by two methods, involving continuous and discontinuous Percoll gradients. Whatever the method of purification, the mitochondrial outer membrane was about 69% intact, and the mitochondria contained very low cytosolic, chloroplastic and peroxisomal contaminations. Nevertheless, as judged by the recovery of fumarase activity, purification on a continuous 28% Percoll gradient gave the best yield in mitochondria, which exhibited a high degree of inner membrane intactness (91%). The purified mitochondria oxidized succinate and malate with good respiratory control and ADP/O ratios. The highest oxidation rate was obtained with succinate as substrate, and malate oxidation was improved (+ 60%) by addition of exogenous NAD+. Experiments using standard respiratory chain inhibitors indicated that, in spruce mitochondria, the alternative pathway was present. Both NAD+-isocitrate dehydrogenase (EC 1.1.1.41) and NADP+-isocitrate dehydrogenase (EC 1.1.1.42) were present in the mitochondrial matrix fraction, and NAD+-IDH activity was about 2-fold higher than NADP+-IDH activity. The NAD+-IDH showed sigmoidal kinetics in response to isocitrate and standard Michaelis-Menten kinetics for NAD+ and Mg2+. The NADP+-IDH, in contrast, displayed lower Km values. For NAD+-IDH the pH optimum was at 7.4, whereas NADP+-IDH exhibited a broad pH optimum between 8.3 and 9. In addition, NAD+-IDH was more thermolabile. Adenine nucleotides and 2-oxoglutarate were found to inhibit NAD(P)+-IDH activities only at high concentrations.  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