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.
Succinic acid is not the dominant fermentation product from glucose in wild-type Escherichia coli W1485. To reduce byproduct formation and increase succinic acid accumulation, pyruvate formate-lyase and lactate dehydrogenase, encoded by pflB and ldhA genes, were inactivated. However, E. coli NZN111, the ldhA and pflB deletion strain, could not utilize glucose anaerobically due to the block of NAD(+) regeneration. To restore glucose utilization, overexpression of nicotinic acid phosphoribosyltransferase, a rate limiting enzyme of NAD(H) synthesis encoded by the pncB gene, resulted in a significant increase in cell mass and succinic acid production. Furthermore, the results indicated a significant increase in NAD(H) pool size, and decrease in the NADH/NAD(+) ratio from 0.64 to 0.13, in particular, the concentration of NAD(+) increased 6.2-fold during anaerobic fermentation. In other words, the supply of enough NAD(+) for NADH oxidation by regulation of NAD(H) salvage synthesis mechanism could improve the cell growth and glucose utilization anaerobically. In addition, the low NADH/NAD(+) ratio also change the metabolite distribution during the dual-phase fermentation. As a result, there was a significant increase in succinic acid production, and it is provided further evidence that regulation of NAD(H) pool and NADH/NAD(+) ratio was very important for succinic acid production.  相似文献   

2.
Liang LY  Liu RM  Ma JF  Chen KQ  Jiang M  Wei P 《Biotechnology letters》2011,33(12):2439-2444
Escherichia coli NZN111 is a double mutant with inactivated lactate dehydrogenase and pyruvate formate-lyase. It cannot utilize glucose anaerobically because of its unusually high intracellular NADH/NAD(+) ratio. We have now constructed a recombinant strain, E. coli NZN111/pTrc99a-mdh, which, during anaerobic fermentation, produced 4.3 g succinic acid l(-1) from 13.5 g glucose l(-1). The NADH/NAD(+) ratio decreased from 0.64 to 0.26. Furthermore, dual-phase fermentation (aerobic growth followed by anaerobic phase) resulted in enhanced succinic acid production and reduced byproduct formation. The yield of succinic acid from glucose during the anaerobic phase was 0.72 g g(-1), and the productivity was 1.01 g l(-1) h(-1).  相似文献   

3.
It is generally known that cofactors play a major role in the production of different fermentation products. This paper is part of a systematic study that investigates the potential of cofactor manipulations as a new tool for metabolic engineering. The NADH/NAD+ cofactor pair plays a major role in microbial catabolism, in which a carbon source, such as glucose, is oxidized using NAD+ and producing reducing equivalents in the form of NADH. It is crucially important for continued cell growth that NADH be oxidized to NAD+ and a redox balance be achieved. Under aerobic growth, oxygen is used as the final electron acceptor. While under anaerobic growth, and in the absence of an alternate oxidizing agent, the regeneration of NAD+ is achieved through fermentation by using NADH to reduce metabolic intermediates. Therefore, an increase in the availability of NADH is expected to have an effect on the metabolic distribution. We have previously investigated a genetic means of increasing the availability of intracellular NADH in vivo by regenerating NADH through the heterologous expression of an NAD(+)-dependent formate dehydrogenase and have demonstrated that this manipulation provoked a significant change in the final metabolite concentration pattern both anaerobically and aerobically (Berríos-Rivera et al., 2002, Metabolic engineering of Escherichia coli: increase of NADH availability by overexpressing an NAD(+)-dependent formate dehydrogenase, Metabolic Eng. 4, 217-229). The current work explores further the effect of substituting the native cofactor-independent formate dehydrogenase (FDH) by an NAD(+)-dependent FDH from Candida boidinii on the NAD(H/+) levels, NADH/NAD+ ratio, metabolic fluxes and carbon-mole yields in Escherichia coli under anaerobic chemostat conditions. Overexpression of the NAD(+)-dependent FDH provoked a significant redistribution of both metabolic fluxes and carbon-mole yields. Under anaerobic chemostat conditions, NADH availability increased from 2 to 3 mol NADH/mol glucose consumed and the production of more reduced metabolites was favored, as evidenced by a dramatic increase in the ethanol to acetate ratio and a decrease in the flux to lactate. It was also found that the NADH/NAD+ ratio should not be used as a sole indicator of the oxidation state of the cell. Instead, the metabolic distribution, like the Et/Ac ratio, should also be considered because the turnover of NADH can be fast in an effort to achieve a redox balance.  相似文献   

4.
Metabolic engineering studies have generally focused on manipulating enzyme levels through either the amplification, addition, or deletion of a particular pathway. However, with cofactor-dependent production systems, once the enzyme levels are no longer limiting, cofactor availability and the ratio of the reduced to oxidized form of the cofactor can become limiting. Under these situations, cofactor manipulation may become crucial in order to further increase system productivity. Although it is generally known that cofactors play a major role in the production of different fermentation products, their role has not been thoroughly and systematically studied. However, cofactor manipulations can potentially become a powerful tool for metabolic engineering. Nicotinamide adenine dinucleotide (NAD) functions as a cofactor in over 300 oxidation-reduction reactions and regulates various enzymes and genetic processes. The NADH/NAD+ cofactor pair plays a major role in microbial catabolism, in which a carbon source, such as glucose, is oxidized using NAD+ producing reducing equivalents in the form of NADH. It is crucially important for continued cell growth that NADH be oxidized to NAD+ and a redox balance be achieved. Under aerobic growth, oxygen is used as the final electron acceptor. While under anaerobic growth, and in the absence of an alternate oxidizing agent, the regeneration of NAD+ is achieved through fermentation by using NADH to reduce metabolic intermediates. Therefore, an increase in the availability of NADH is expected to have an effect on the metabolic distribution. This paper investigates a genetic means of manipulating the availability of intracellular NADH in vivo by regenerating NADH through the heterologous expression of an NAD(+)-dependent formate dehydrogenase. More specifically, it explores the effect on the metabolic patterns in Escherichia coli under anaerobic and aerobic conditions of substituting the native cofactor-independent formate dehydrogenase (FDH) by and NAD(+)-dependent FDH from Candida boidinii. The over-expression of the NAD(+)-dependent FDH doubled the maximum yield of NADH from 2 to 4 mol NADH/mol glucose consumed, increased the final cell density, and provoked a significant change in the final metabolite concentration pattern both anaerobically and aerobically. Under anaerobic conditions, the production of more reduced metabolites was favored, as evidenced by a dramatic increase in the ethanol-to-acetate ratio. Even more interesting is the observation that during aerobic growth, the increased availability of NADH induced a shift to fermentation even in the presence of oxygen by stimulating pathways that are normally inactive under these conditions.  相似文献   

5.
Guo T  Kong J  Zhang L  Zhang C  Hu S 《PloS one》2012,7(4):e36296
Lactococcus lactis is a well-studied bacterium widely used in dairy fermentation and capable of producing metabolites with organoleptic and nutritional characteristics. For fine tuning of the distribution of glycolytic flux at the pyruvate branch from lactate to diacetyl and balancing the production of the two metabolites under aerobic conditions, a constitutive promoter library was constructed by randomizing the promoter sequence of the H(2)O-forming NADH oxidase gene in L. lactis. The library consisted of 30 promoters covering a wide range of activities from 7,000 to 380,000 relative fluorescence units using a green fluorescent protein as reporter. Eleven typical promoters of the library were selected for the constitutive expression of the H(2)O-forming NADH oxidase gene in L. lactis, and the NADH oxidase activity increased from 9.43 to 58.17-fold of the wild-type strain in small steps of activity change under aerobic conditions. Meanwhile, the lactate yield decreased from 21.15 ± 0.08 mM to 9.94 ± 0.07 mM, and the corresponding diacetyl production increased from 1.07 ± 0.03 mM to 4.16 ± 0.06 mM with the intracellular NADH/NAD(+) ratios varying from 0.711 ± 0.005 to 0.383 ± 0.003. The results indicated that the reduced pyruvate to lactate flux was rerouted to the diacetyl with an almost linear flux variation via altered NADH/NAD(+) ratios. Therefore, we provided a novel strategy to precisely control the pyruvate distribution for fine tuning of the lactate and diacetyl production through promoter engineering in L. lactis. Interestingly, the increased H(2)O-forming NADH oxidase activity led to 76.95% lower H(2)O(2) concentration in the recombinant strain than that of the wild-type strain after 24 h of aerated cultivation. The viable cells were significantly elevated by four orders of magnitude within 28 days of storage at 4°C, suggesting that the increased enzyme activity could eliminate H(2)O(2) accumulation and prolong cell survival.  相似文献   

6.
Under anaerobic conditions, Propionibacterium pentosaceum reduces nitrate to nitrite until nitrate is exhausted from the medium when nitrite is converted into N2 or N2O. In the presence of nitrate, fermentation patterns for lactate, glycerol and pyruvate were different from those obtained during anaerobic growth without an inorganic electron acceptor. In the presence of these substrates, a drastic decrease in propionate formation was observed, some pyruvate accumulated during growth with lactate, and acetate was produced from glycerol. Acetate production from lactate and pyruvate was not influenced by the presence of nitrate. Furthermore, CO2 was produced by citric acid cycle activity. The fermentation pattern during nitrite reduction resembled that of P. pentosaceum grown anaerobically without an inorganic electron acceptor. Nitrits has a toxic effect, since bacteria inoculated into a medium with 9 mM-nitrite failed to grow. The cytochrome spectrum of anaerobically grown P. pentosaceum was similar with and without nitrate. In membrane fractions of bacteria grown anaerobically with nitrate, cytochrome b functioned in the transfer of electrons from lactate, glycerol I-phosphate and NADH to nitrate. Molar growth yeilds were increased in the presence of nitrate, indicating an increased production of ATP. This could be explained by citric acid cycle activity, and by ocidative phosphorylation coupled to nitrate reduction. Assuming that I mol ATP is formed in the electron transfer from lactate or glycerol I-phosphate to nitrate, and that 2 mol ATP are formed in the electron transfer from NADH to nitrate, YATP values (g dry wt bacteria/mol ATP) were obtained of between 5-0 and 12-6. The higher YATP values were similar to those obtained during anaerobic growth without an inorganic electron acceptor. This supports the assumptions about the efficiency of oxidative phosphorylation for electron transport to nitrate. Low YAPT values were found when high concentrations of nitrite (15 to 50 mM) accumulated, and were probably due to the toxic effect of nitrite.  相似文献   

7.
Escherichia coli overexpressing a NAD(+)-dependent formate dehydrogenase (FDH) from Candida boidinii was grown in chemostat culture on various carbon sources at 0.05 h(-1) dilution rate, under anaerobic conditions using defined medium and compared to a control without the heterologous FDH pathway. Metabolic fluxes, NADH/NAD(+) ratios and NAD(H/(+)) levels were determined under a range of intracellular NADH availability. The effect of NADH manipulation on the distribution of metabolic fluxes in E. coli was assessed under steady-state conditions. The heterologous FDH pathway converts 1 mol of formate into 1 mol of NADH and carbon dioxide, in contrast with the native FDH where no cofactor involvement is present. Previously, we found that this NADH regeneration system doubled the maximum yield of NADH from 2 to 4 mol NADH/mol glucose consumed and reached 4.6 mol NADH/mol of substrate when sorbitol was used as a carbon source in a complex medium. In the current study, it was found that higher NADH yields and NADH/NAD(+) ratios were achieved with our in vivo NADH regeneration system compared to a control lacking the new FDH pathway in the three carbon sources (glucose, gluconate and sorbitol) examined suggesting a more reduced intracellular environment. The total NAD(H/(+)) amounts were very similar for all the combinations studied. It was also found that the ethanol to acetate ratio increased with increased NADH availability. This ratio increased from 1.05 for the control strain in glucose to 9.45 for the strain expressing the heterologous NAD(+)-dependent FDH in sorbitol.  相似文献   

8.
The involvement of nicotinamide adenine nucleotides (NAD(+), NADH) in the regulation of glycolysis in Lactococcus lactis was investigated by using (13)C and (31)P NMR to monitor in vivo the kinetics of the pools of NAD(+), NADH, ATP, inorganic phosphate (P(i)), glycolytic intermediates, and end products derived from a pulse of glucose. Nicotinic acid specifically labeled on carbon 5 was synthesized and used in the growth medium as a precursor of pyridine nucleotides to allow for in vivo detection of (13)C-labeled NAD(+) and NADH. The capacity of L. lactis MG1363 to regenerate NAD(+) was manipulated either by turning on NADH oxidase activity or by knocking out the gene encoding lactate dehydrogenase (LDH). An LDH(-) deficient strain was constructed by double crossover. Upon supply of glucose, NAD(+) was constant and maximal (approximately 5 mm) in the parent strain (MG1363) but decreased abruptly in the LDH(-) strain both under aerobic and anaerobic conditions. NADH in MG1363 was always below the detection limit as long as glucose was available. The rate of glucose consumption under anaerobic conditions was 7-fold lower in the LDH(-) strain and NADH reached high levels (2.5 mm), reflecting severe limitation in regenerating NAD(+). However, under aerobic conditions the glycolytic flux was nearly as high as in MG1363 despite the accumulation of NADH up to 1.5 mm. Glyceraldehyde-3-phosphate dehydrogenase was able to support a high flux even in the presence of NADH concentrations much higher than those of the parent strain. We interpret the data as showing that the glycolytic flux in wild type L. lactis is not primarily controlled at the level of glyceraldehyde-3-phosphate dehydrogenase by NADH. The ATP/ADP/P(i) content could play an important role.  相似文献   

9.
《BBA》1987,893(3):386-397
Three NAD(P)H dehydrogenases were found and purified from a soluble fraction of cells of the purple non-sulfur bacterium Rhodobacter capsulatus, strain B10. Molecular mass of NAD(P)H, NADPH and NADH dehydrogenases are 67 000 (4 · 18 000), 35 000 and 39 000, and the isoelectric points are 4.6, 4.3 and 4.5, respectively. NAD(P)H dehydrogenase is characterized by a higher sensitivity to quinacrine, NADPH dehydrogenase by its sensitivity to p-chloromercuribenzoate and NADH dehydrogenase by its sensitivity to sodium arsenite. In contrast to the other two enzymes, NAD(P)H dehydrogenase is capable of oxidizing NADPH as well as NADH, but the ratio of their oxidation rates depends on the pH. All NAD(P)H dehydrogenases reacted with ferricyanide, 2,6-dichlorophenolindophenol, benzoquinone and naphthoquinone, but did not exhibit transhydrogenase, reductase or oxidase activity. Moreover, NADH dehydrogenase was also capable of reducing FAD and FMN. NAD(P)H and NADH dehydrogenases possessed cytochrome-c reductase activity, which was stimulated by menadione and ubiquinone Q1. The activity of NAD(P)H and NADH dehydrogenases depended on culture-growth conditions. The activity of NAD(P)H dehydrogenase from cells grown under chemoheterotrophic aerobic conditions was the lowest and it increased notably under photoheterotrophic anaerobic conditions upon lactate or malate growth limitation. The activity of NADH dehydrogenase was higher from the cells grown under photoheterotrophic anaerobic conditions upon nitrate growth limitation and under chemoheterotrophic aerobic conditions. NADPH dehydrogenase synthesis dependence on R. capsulatus growth conditions was insignificant.  相似文献   

10.
大肠杆菌NZN111厌氧发酵的主要产物为丁二酸,是发酵生产丁二酸的潜力菌株。但是由于敲除了乳酸脱氢酶的编码基因 (ldhA) 和丙酮酸甲酸裂解酶的编码基因 (pflB),导致辅酶NADH/NAD+不平衡,厌氧条件下不能利用葡萄糖生长代谢。构建烟酸转磷酸核糖激酶的重组菌Escherichia coli NZN111/pTrc99a-pncB,在厌氧摇瓶发酵过程中通过添加0.5 mmol/L的烟酸、0.3 mmol/L的IPTG诱导后重组菌的烟酸转磷酸核糖激酶 (Nicotinic acid phosphor  相似文献   

11.
The nifJ gene codes for pyruvate:ferredoxin oxidoreductase (PFOR), which reduces ferredoxin during fermentative catabolism of pyruvate to acetyl-coenzyme A (acetyl-CoA). A nifJ knockout mutant was constructed that lacks one of two pathways for the oxidation of pyruvate in the cyanobacterium Synechococcus sp. strain PCC 7002. Remarkably, the photoautotrophic growth rate of this mutant increased by 20% relative to the wild-type (WT) rate under conditions of light-dark cycling. This result is attributed to an increase in the quantum yield of photosystem II (PSII) charge separation as measured by photosynthetic electron turnover efficiency determined using fast-repetition-rate fluorometry (F(v)/F(m)). During autofermentation, the excretion of acetate and lactate products by nifJ mutant cells decreased 2-fold and 1.2-fold, respectively. Although nifJ cells displayed higher in vitro hydrogenase activity than WT cells, H(2) production in vivo was 1.3-fold lower than the WT level. Inhibition of acetate-CoA ligase and pyruvate dehydrogenase complex by glycerol eliminated acetate production, with a resulting loss of reductant and a 3-fold decrease in H(2) production by nifJ cells compared to WT cells. Continuous electrochemical detection of dissolved H(2) revealed two temporally resolved phases of H(2) production during autofermentation, a minor first phase and a major second phase. The first phase was attributed to reduction of ferredoxin, because its level decreased 2-fold in nifJ cells. The second phase was attributed to glycolytic NADH production and decreased 20% in nifJ cells. Measurement of the intracellular NADH/NAD(+) ratio revealed that the reductant generated by PFOR contributing to the first phase of H(2) production was not in equilibrium with bulk NADH/NAD(+) and that the second phase corresponded to the equilibrium NADH-mediated process.  相似文献   

12.
过量表达苹果酸脱氢酶对大肠杆菌NZN111产丁二酸的影响   总被引:2,自引:1,他引:1  
大肠杆菌NZN111是敲除了乳酸脱氢酶的编码基因 (ldhA) 和丙酮酸-甲酸裂解酶的编码基因 (pflB) 的工程菌,厌氧条件下由于辅酶NAD(H) 的不平衡导致其丧失了代谢葡萄糖的能力。构建了苹果酸脱氢酶的重组菌大肠杆菌NZN111/pTrc99a-mdh,在厌氧摇瓶发酵过程中通过0.3 mmol/L的IPTG诱导后重组菌的苹果酸脱氢酶 (Malate dehydrogenase,MDH) 酶活较出发菌株提高了14.8倍,NADH/NAD+的比例从0.64下降到0.26,同时NAD+和NADH浓度分别  相似文献   

13.
Barley (Hordeum vulgare L.) has NADH-specific and NAD(P)H-bispecific nitrate reductase isozymes. Four isogenic lines with different nitrate reductase isozyme combinations were used to determine the role of NADH and NAD(P)H nitrate reductases on nitrate transport and assimilation in barley seedlings. Both nitrate reductase isozymes were induced by nitrate and were required for maximum nitrate assimilation in barley seedlings. Genotypes lacking the NADH isozyme (Az12) or the NAD(P)H isozyme (Az70) assimilated 65 or 85%, respectively, as much nitrate as the wild type. Nitrate assimilation by genotype (Az12;Az70) which is deficient in both nitrate reductases, was only 13% of the wild type indicating that the NADH and NAD(P)H nitrate reductase isozymes are responsible for most of the nitrate reduction in barley seedlings. For all genotypes, nitrate assimilation rates in the dark were about 55% of the rates in light. Hypotheses that nitrate reductase has direct or indirect roles in nitrate uptake were not supported by this study. Induction of nitrate transporters and the kinetics of net nitrate uptake were the same for all four genotypes indicating that neither nitrate reductase isozyme has a direct role in nitrate uptake in barley seedlings.  相似文献   

14.
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   

15.
The amount of Vitreoscilla hemoglobin (VHb) expression was modulated over a broad range with an isopropyl-beta-D-thiogalactopyranoside- (IPTG-) inducible plasmid, and the consequences on microaerobic Escherichia coli physiology were examined in glucose fed-batch cultivations. The effect of IPTG induction on growth under oxygen-limited conditions was most visible during late fed-batch phase where the final cell density increased initially linearly with increasing VHb concentrations, ultimately saturating at a 2.7-fold increase over the VHb-negative (Vhb(-)) control. During the same growth phase, the specific excretions of fermentation by-products, acetate, ethanol, formate, lactate, and succinate from the culture expressing the highest amount of VHb were reduced by 25%, 49%, 68%, 72%, and 50%, respectively, relative to the VHb(-) control. During the exponential growth phase, VHb exerted a positive but smaller control on growth rate, growth yield, and respiration. Varying the amount of VHb from 0 to 3.8 mumol/g dry cell weight (DCW) increased the specific growth rate, the growth yield, and the oxygen consumption rate by 33%, 35%, and 60%, respectively. Increasing VHb concentration to 3.8 mumol/g DCW suppressed the rate of carbon dioxide evolution in the exponential phase by 30%. A metabolic flux distribution analysis incorporating data from these cultivations discloses that VHb(+) cells direct a larger fraction of glucose toward the pentose phosphate pathway and a smaller fraction of carbon through the tricarboxylic acid cycle from acetyl coenzyme A. The overall nicotinamide adenine dinucleotide [NAD(P)H] flux balance indicates that VHb-expressing cells generate a net NADH flux by the NADH/NADPH transhydrogenase while the VHb(-) cells yield a net NADPH flux under the same growth conditions. Flux distribution analysis also reveals that VHb(+) cells have a smaller adenosine triphosphate (ATP) synthesis rate from substrate-level phosphorylation but a larger overall ATP production rate under microaerobic conditions. The thermodynamic efficiency of growth, based on reducing equivalents generated per unit of biomass produced, is greater for VHb(+) cells. (c) 1996 John Wiley & Sons, Inc.  相似文献   

16.
The aim of this work was to understand the steps controlling the biotransformation of trimethylammonium compounds into L(-)-carnitine by Escherichia coli. The high-cell density reactor steady-state levels of carbon source (glycerol), biotransformation substrate (crotonobetaine), acetate (anaerobiosis product) and fumarate (as an electron acceptor) were pulsed by increasing them fivefold. Following the pulse, the evolution of the enzyme activities involved in the biotransformation process of crotonobetaine into L(-)-carnitine (crotonobetaine hydration), in the synthesis of acetyl-CoA (ACS: acetyl-CoA synthetase and PTA: ATP: acetate phosphotransferase) and in the distribution of metabolites for the tricarboxylic acid (ICDH: isocitrate dehydrogenase) and glyoxylate (ICL: isocitrate lyase) cycles was monitored. In addition, the levels of carnitine, the cell ATP content and the NADH/NAD(+) ratio were measured in order to assess the importance and participation of these energetic coenzymes in the catabolic system. The results provided an experimental demonstration of the important role of the glyoxylate shunt during biotransformation and the need for high levels of ATP to maintain metabolite transport and biotransformation. Moreover, the results obtained for the NADH/NAD(+) pool indicated that it is correlated with the biotransformation process at the NAD(+) regeneration and ATP production level in anaerobiosis. More importantly, a linear correlation between the NADH/NAD(+) ratio and the levels of the ICDH and ICL (carbon and electron flows) and the PTA and ACS (acetate and ATP production and acetyl-CoA synthesis) activity levels was assessed. The main metabolic pathway operating during cell metabolic perturbation with a pulse of glycerol and acetate in the high-cell density membrane reactor was that related to ICDH and ICL, both regulating the carbon metabolism, together with PTA and ACS enzymes (regulating ATP production).  相似文献   

17.
18.
NAD(H) kinase catalyzes the phosphorylation of NAD(H) to form NADP(H) using ATP or inorganic polyphosphate as a phosphoryl donor. While the enzyme is conserved throughout prokaryotes and eukaryotes, remarkable differences in kinetic parameters including substrate preference, cation dependence, and physiological roles exist among the organisms. In the present study, we biochemically characterized NAD(H) kinase from the anaerobic/microaerophilic fermentative protozoan parasite Entamoeba histolytica, which lacks the conventional mitochondria capable of oxidative phosphorylation, leading to ATP. The kinetic properties of E. histolytica NAD(H) kinase recombinantly produced in Escherichia coli showed remarkable differences from those in bacteria and higher eukaryotes. Entamoeba NAD(H) kinase preferred NADH to NAD+ as the phosphoryl acceptor, utilized nucleoside triphosphates including ATP, GTP and deoxyATP, but not nucleoside di-, mono-phosphates, or inorganic polyphosphates, as the phosphoryl donor. To further understand the physiological roles in E. histolytica, we generated a stable transformant overexpressing NAD(H) kinase. Overexpression of NAD(H) kinase resulted in a 1.6–2 fold increase in the NADPH and NADP+ concentrations, a 40% reduction of the intracellular concentration of reactive oxygen species, and also led to increased tolerance toward hydrogen peroxide. These data, together with the essentially of NAD(H) kinase gene, underscore its significance as an NADP(H)-producing enzyme in this organism, and should help in designing of drugs targeting this enzyme.  相似文献   

19.
Han Q  Li G  Li J 《Biochimica et biophysica acta》2000,1523(2-3):246-253
A specific chorion peroxidase is present in Aedes aegypti and this enzyme is responsible for catalyzing chorion protein cross-linking through dityrosine formation during chorion hardening. Peroxidase-mediated dityrosine cross-linking requires H(2)O(2), and this study discusses the possible involvement of the chorion peroxidase in H(2)O(2) formation by mediating NADH/O(2) oxidoreduction during chorion hardening in A. aegypti eggs. Our data show that mosquito chorion peroxidase is able to catalyze pH-dependent NADH oxidation, which is enhanced in the presence of Mn(2+). Molecular oxygen is the electron acceptor during peroxidase-catalyzed NADH oxidation, and reduction of O(2) leads to the production of H(2)O(2), demonstrated by the formation of dityrosine in a NADH/peroxidase reaction mixture following addition of tyrosine. An oxidoreductase capable of catalyzing malate/NAD(+) oxidoreduction is also present in the egg chorion of A. aegypti. The cooperative roles of chorion malate/NAD(+)oxidoreductase and chorion peroxidase on generating H(2)O(2) with NAD(+) and malate as initial substrates were demonstrated by the production of dityrosine after addition of tyrosine to a reaction mixture containing NAD(+) and malate in the presence of both malate dehydrogenase fractions and purified chorion peroxidase. Data suggest that chorion peroxidase-mediated NADH/O(2) oxidoreduction may contribute to the formation of the H(2)O(2) required for chorion protein cross-linking mediated by the same peroxidase, and that the chorion associated malate dehydrogenase may be responsible for the supply of NADH for the H(2)O(2) production.  相似文献   

20.
Chlorophyll a fluorescence rise kinetics (from 50 μs to 1 s) were used to investigate the non-photochemical reduction of the plastoquinone (PQ) pool in osmotically broken spinach chloroplasts (Spinacia oleracea L.). Incubation of the chloroplasts in the presence of exogenous NADPH or NADH resulted in significant changes in the shape of the fluorescence transient reflecting an NAD(P)H-dependent accumulation of reduced PQ in the dark, with an extent depending on the concentration of NAD(P)H and the availability of oxygen; the dark reduction of the PQ pool was saturated at lower NAD(P)H concentrations and reached a higher level when the incubation took place under anaerobic conditions than when it occurred under aerobic conditions. Under both conditions NADPH was more effective than NADH in reducing PQ, however only at sub-saturating concentrations. Neither antimycin A nor rotenone were found to alter the effect of NAD(P)H. The addition of mercury chloride to the chloroplast suspension decreased the NAD(P)H-dependent dark reduction of the PQ pool, with the full inhibition requiring higher mercury concentrations under anaerobic than under aerobic conditions. This is the first time that this inhibitory role of mercury is reported for higher plants. The results demonstrate that in the dark the redox state of the PQ pool is regulated by the reduction of PQ via a mercury-sensitive NAD(P)H-PQ oxidoreductase and the reoxidation of reduced PQ by an O2-dependent pathway, thus providing additional evidence for the existence of a chlororespiratory electron transport chain in higher plant chloroplasts. This revised version was published online in June 2006 with corrections to the Cover Date.  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