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 31 毫秒
1.
Enterobacter aerogenes NBRC12010 was able to ferment glycerol to ethanol and hydrogen gas. Fermentation of glycerol ceased in the stationary phase of growth, and it was activated by electrochemical reactions using thionine as an electron transfer mediator from bacterial cells to an electrode. Using resting cells of E. aerogenes NBRC12010 in only citrate buffer solution, the cells did not consume glycerol at all, but they could metabolize glucose. These results suggest that the regulation of glycerol metabolism occurred at enzymatic steps before glycolysis. In E. aerogenes NBRC12010, glycerol was metabolized via glycerol dehydrogenase (GDH) and then dehydroxyacetone kinase. The GDH-catalyzed reaction mainly depended on the ratio of NAD+/NADH. At a NAD+/NADH ratio of nearly 1 or less, it was substantially suppressed and glycerol metabolism stopped. When the ratio was higher than 1, GDH was activated and glycerol was metabolized. Thus, the reaction of glycerol metabolism depended on the balance of cellular NAD+/NADH. Exogenous NADH was oxidized to NAD+ by electrochemical reactions with thionine. We proposed the activation mechanism of glycerol metabolism under electrochemical conditions.  相似文献   

2.
Saccharomyces cerevisiae maintains a redox balance under fermentative growth conditions by re-oxidizing NADH formed during glycolysis through ethanol formation. Excess NADH stimulates the synthesis of mainly glycerol, but also of other compounds. Here, we investigated the production of primary and secondary metabolites in S. cerevisiae strains where the glycerol production pathway was inactivated through deletion of the two glycerol-3-phosphate dehydrogenases genes (GPD1/GPD2) and replaced with alternative NAD+-generating pathways. While these modifications decreased fermentative ability compared to the wild-type strain, all improved growth and/or fermentative ability of the gpd1Δgpd2Δ strain in self-generated anaerobic high sugar medium. The partial NAD+ regeneration ability of the mutants resulted in significant amounts of alternative products, but at lower yields than glycerol. Compared to the wild-type strain, pyruvate production increased in most genetically manipulated strains, whereas acetate and succinate production decreased in all strains. Malate production was similar in all strains. Isobutanol production increased substantially in all genetically manipulated strains compared to the wild-type strain, whereas only mutant strains expressing the sorbitol producing SOR1 and srlD genes showed increases in isoamyl alcohol and 2-phenyl alcohol. A marked reduction in ethyl acetate concentration was observed in the genetically manipulated strains, while isobutyric acid increased. The synthesis of some primary and secondary metabolites appears more readily influenced by the NAD+/NADH availability. The data provide an initial assessment of the impact of redox balance on the production of primary and secondary metabolites which play an essential role in the flavour and aroma character of beverages.  相似文献   

3.
4.
Xanthine dehydrogenase AtXDH1 from Arabidopsis thaliana is a key enzyme in purine degradation where it oxidizes hypoxanthine to xanthine and xanthine to uric acid. Electrons released from these substrates are either transferred to NAD+ or to molecular oxygen, thereby yielding NADH or superoxide, respectively. By an alternative activity, AtXDH1 is capable of oxidizing NADH with concomitant formation of NAD+ and superoxide. Here we demonstrate that in comparison to the specific activity with xanthine as substrate, the specific activity of recombinant AtXDH1 with NADH as substrate is about 15-times higher accompanied by a doubling in superoxide production. The observation that NAD+ inhibits NADH oxidase activity of AtXDH1 while NADH suppresses NAD+-dependent xanthine oxidation indicates that both NAD+ and NADH compete for the same binding-site and that both sub-activities are not expressed at the same time. Rather, each sub-activity is determined by specific conditions such as the availability of substrates and co-substrates, which allows regulation of superoxide production by AtXDH1. Since AtXDH1 exhibits the most pronounced NADH oxidase activity among all xanthine dehydrogenase proteins studied thus far, our results imply that in particular by its NADH oxidase activity AtXDH1 is an efficient producer of superoxide also in vivo.  相似文献   

5.
Glycerol is a major by-product of ethanol fermentation by Saccharomyces cerevisiae and typically 2–3% of the sugar fermented is converted to glycerol. Replacing the NAD+-regenerating glycerol pathway in S. cerevisiae with alternative NADH reoxidation pathways may be useful to produce metabolites of biotechnological relevance. Under fermentative conditions yeast reoxidizes excess NADH through glycerol production which involves NADH-dependent glycerol-3-phosphate dehydrogenases (Gpd1p and Gpd2p). Deletion of these two genes limits fermentative activity under anaerobic conditions due to accumulation of NADH. We investigated the possibility of converting this excess NADH to NAD+ by transforming a double mutant (gpd1gpd2∆) with alternative oxidoreductase genes that might restore the redox balance and produce either sorbitol or propane-1,2-diol. All of the modifications improved fermentative ability and/or growth of the double mutant strain in a self-generated anaerobic high sugar medium. However, these strain properties were not restored to the level of the parental wild-type strain. The results indicate an apparent partial NAD+ regeneration ability and formation of significant amounts of the commodity chemicals like sorbitol or propane-1,2-diol. The ethanol yields were maintained between 46 and 48% of the sugar mixture. Other factors apart from the maintenance of the redox balance appeared to influence the growth and production of the alternative products by the genetically manipulated strains.  相似文献   

6.
In this article we compare the kinetic behavior toward pyridine nucleotides (NAD+, NADH) of NAD+-malic enzyme, pyruvate dehydrogenase, isocitrate dehydrogenase, α-ketoglutarate dehydrogenase, and glycine decarboxylase extracted from pea (Pisum sativum) leaf and potato (Solanum tuberosum) tuber mitochondria. NADH competitively inhibited all the studied dehydrogenases when NAD+ was the varied substrate. However, the NAD+-linked malic enzyme exhibited the weakest affinity for NAD+ and the lowest sensitivity for NADH. It is suggested that NAD+-linked malic enzyme, when fully activated, is able to raise the matricial NADH level up to the required concentration to fully engage the rotenone-resistant internal NADH-dehydrogenase, whose affinity for NADH is weaker than complex I.  相似文献   

7.
A new method for extracting pyridine nucleotides from tissue samples at room temperature that allows the simultaneous extraction of both the oxidized and reduced nucleotide when using a 70% buffered ethanol solution as the extractant has been developed. The extraction efficiencies for NAD+ and NADH were 91 and 102%, respectively. The extraction method was followed by a combined bioluminescent assay of both nucleotides. A bacterial bioluminescent system, which included luciferase and low levels of a NADH-specific oxidoreductase, was used to produce a constant light intensity directly proportional to the amount of NADH in the tissue extract sample. When the NADH had been measured, the NAD+ present in the extract was enzymatically converted to NADH by the addition of alcohol dehydrogenase, after which the second increase in light level was recorded. The sensitivity of the bioluminescent assay presented here is 5 × 10?14 mol NADH or NAD+ per assay.  相似文献   

8.
Escherichia coli BA002, the ldhA and pflB deletion strain, cannot utilize glucose anaerobically due to the inability to regenerate NAD+. To regulate NAD(H) pool size and NADH/NAD+ ratio, overexpression of the enzymes in the NAD(H) biosynthetic pathways in BA002 was investigated. The results clearly demonstrate that the increased NAD(H) pool size and the decreased NADH/NAD+ ratio improved the glucose consumption and cell growth, which improved succinic acid production. When the pncB and the nadD genes were co-overexpressed in CA102, the ratio of NADH/NAD+ was decreased from 0.60 to 0.12, and the concentration of NAD(H) was the highest among that of all the strains. Moreover, the dry cell weight (DCW), glucose consumption, and the concentration of succinic acid in CA102 were also the highest. Based on the sufficient NAD+ supply after gene modification in the NAD(H) biosynthetic pathways, reductive carbon sources with different amounts of NADH can further change the distribution of metabolites. When sorbitol was used as a carbon source in CA102, the byproducts were lower than those of glucose fermentation, and the yield of succinic acid was increased.  相似文献   

9.
The catalytic activity of the LDH-isoenzymes depends on their tetrameric structure. Low pH or other denaturants leads to dissociation into monomers and to the loss of the specific activity. After removal of the denaturing conditions reassociation and reactivation occur spontaneously. Neither NADH nor NAD+ shows a significant effect on the reactivation. We have isolated two different peptides which isoenzyme specifically inhibit the reactivation of dissociated LDH. Inhibition was abolished by treating with proteases. Additionally, NAD+ and NADH were found to be antagonists of the inhibitors. The heart-type enzyme-inhibitor system is especially susceptible for NADH whereas NAD+ affects the inhibition only slightly. The muscle-type system shows the opposite behavior, e.g., the completely inhibited system can be fully reactivated by NAD+ but not by NADH. These findings together with first kinetic studies suggest a possible specific regulatory function of these peptides.  相似文献   

10.
The kinetic mechanism of betaine aldehyde dehydrogenase from leaves of the plant Amaranthus hypochondriacus is ordered with NAD+ adding first. NADH is a noncompetitive inhibitor against NAD+, which was interpreted before as evidence of an iso mechanism, in which NAD+ and NADH binds to different forms of free enzyme. With the aim of testing the proposed kinetic mechanism, we have now investigated the ability of NADH to form different complexes with the enzyme. By initial velocity and equilibrium binding studies, we found that the steady-state levels of E.glycine betaine are negligible, ruling out binding of NADH to this complex. However, NADH readily bind to E.betaine aldehyde, whose levels most likely are kinetically significant given its low dissociation constant. Also, NADH combined with E.NADH and E.NAD+. Finally, NADH was not able to revert the hydride transfer step, what suggest that there is no acyl-enzyme intermediate, i.e. the release of the reduced dinucleotide takes place after the deacylation step. Although formation of the complex E.NAD+.NADH would produce an uncompetitive effect in the inhibition of NADH against NAD+, the iso mechanism cannot be conclusively discarded.  相似文献   

11.

Background

The ratio of NAD+/NADH is a key indicator that reflects the overall redox state of the cells. Until recently, there were no methods for real time NAD+/NADH monitoring in living cells. Genetically encoded fluorescent probes for NAD+/NADH are fundamentally new approach for studying the NAD+/NADH dynamics.

Methods

We developed a genetically encoded probe for the nicotinamide adenine dinucleotide, NAD(H), redox state changes by inserting circularly permuted YFP into redox sensor T-REX from Thermus aquaticus. We characterized the sensor in vitro using spectrofluorometry and in cultured mammalian cells using confocal fluorescent microscopy.

Results

The sensor, named RexYFP, reports changes in the NAD+/NADH ratio in different compartments of living cells. Using RexYFP, we were able to track changes in NAD+/NADH in cytoplasm and mitochondrial matrix of cells under a variety of conditions. The affinity of the probe enables comparison of NAD+/NADH in compartments with low (cytoplasm) and high (mitochondria) NADH concentration. We developed a method of eliminating pH-driven artifacts by normalizing the signal to the signal of the pH sensor with the same chromophore.

Conclusion

RexYFP is suitable for detecting the NAD(H) redox state in different cellular compartments.

General significance

RexYFP has several advantages over existing NAD+/NADH sensors such as smallest size and optimal affinity for different compartments. Our results show that normalizing the signal of the sensor to the pH changes is a good strategy for overcoming pH-induced artifacts in imaging.  相似文献   

12.
Mitochondria from the parasitic helminth, Hymenolepis diminuta, catalyzed both NADPH:NAD+ and NADH:NADP+ transhydrogenase reactions which were demonstrable employing the appropriate acetylpyridine nucleotide derivative as the hydride ion acceptor. Thionicotinamide NAD+ would not serve as the oxidant in the former reaction. Under the assay conditions employed, neither reaction was energy linked, and the NADPH:NAD+ system was approximately five times more active than the NADH:NADP+ system. The NADH:NADP+ reaction was inhibited by phosphate and imidazole buffers, EDTA, and adenyl nucleotides, while the NADPH:NAD+ reaction was inhibited only slightly by imidazole and unaffected by EDTA and adenyl nucleotides. Enzyme coupling techniques revealed that both transhydrogenase systems functioned when the appropriate physiological pyridine nucleotide was the hydride ion acceptor. An NADH:NAD+ transhydrogenase system, which was unaffected by EDTA, or adenyl nucleotides, also was demonstrable in the mitochondria of H. diminuta. Saturation kinetics indicated that the NADH:NAD+ reaction was the product of an independent enzyme system. Mitochondria derived from another parasitic helminth, Ascaris lumbricoides, catalyzed only a single transhydrogenase reaction, i.e., the NADH:NAD+ activity. Transhydrogenase systems from both parasites were essentially membrane bound and localized on the inner mitochondrial membrane. Physiologically, the NADPH:NAD+ transhydrogenase of H. diminuta may serve to couple the intramitochondrial metabolism of malate (via an NADP linked “malic” enzyme) to the anaerobic NADH-dependent ATP-generating fumarate reductase system. In A. lumbricoides, where the intramitochondrial metabolism of malate depends on an NAD-linked “malic” enzyme which is localized primarily in the intermembrane space, the NADH:NAD+ transhydrogenase activity may serve physiologically in the translocation of hydride ions across the inner membrane to the anaerobic energy-generating fumarate reductase system.  相似文献   

13.
The effects of Tinopals (cationic benzoxazoles) AMS-GX and 5BM-GX on NADH-oxidase, NADH:ferricyanide reductase, and NADH APAD+ transhydrogenase reactions and energy-linked NAD+ reduction by succinate, catalyzed by NADH:ubiquinone oxidoreductase (Complex I) in submitochondrial particles (SMP), were investigated. AMS-GX competes with NADH in NADH-oxidase and NADH:ferricyanide reductase reactions (K i = 1 M). 5BM-GX inhibits those reactions with mixed type with respect to NADH (K i = 5 M) mechanism. Neither compound affects reverse electron transfer from succinate to NAD+. The type of the Tinopals' effect on the NADH APAD+ transhydrogenase reaction, occurring with formation of a ternary complex, suggests the ordered binding of nucleotides by the enzyme during the reaction: AMS-GX and 5BM-GX inhibit this reaction uncompetitively just with respect to one of the substrates (APAD+ and NADH, correspondingly). The competition between 5BM-GX and APAD+ confirms that NADH is the first substrate bound by the enzyme. Direct and reverse electron transfer reactions demonstrate different specificity for NADH and NAD+ analogs: the nicotinamide part of the molecule is significant for reduced nucleotide binding. The data confirm the model suggesting that during NADH APAD+ reaction, occurring with ternary complex formation, reduced nucleotide interacts with the center participating in NADH oxidation, whereas oxidized nucleotide reacts with the center binding NAD+ in the reverse electron transfer reaction.  相似文献   

14.
The effect of NAD(P) and analogs of this nucleotide on nitrogenase activity in Rhodospirillum rubrum has been studied. Addition of NAD+ to nitrogen fixing Rsp. rubrum leads to inhibition of nitrogenase. NADP+ has the same effect but NADH or analogs modified in the nicotinamide portion do not cause inhibition. In contrast to ammonium ions, addition of NAD+ leads to inhibition of nitrogenase in cells that have been N-starved under argon. The inhibitory effect of NAD+ is more pronounced at lower light intensities. Addition of NAD+ also leads to inhibition of glutamine synthetase, a phenomenon also occurring when “switchoff” is produced by the addition of effectors such as ammonium ions or glutamine. It is also shown that NAD+ is taken up by Rsp. rubrum cells.  相似文献   

15.
In Klebsiella pneumoniae, 3-hydroxypropaldehyde is converted to 3-hydroxypropionic acid (3-HP) by aldehyde dehydrogenase (ALDH) with NAD+ as a cofactor. Although ALDH overexpression stimulates the formation of 3-HP, it ceases to accumulate when NAD+ is exhausted. Here we show that NAD+ regeneration, together with ALDH overexpression, facilitates 3-HP production and benefits cell growth. Three distinct NAD+-regenerating enzymes: NADH oxidase and NADH dehydrogenase from K. pneumoniae, and glycerol-3-phosphate dehydrogenase (GPD1) from Saccharomyces cerevisiae, were individually expressed in K. pneumoniae. In vitro assay showed their higher activities than that of the control, indicating their capacities to regenerate NAD+. When they were respectively co-expressed with ALD4, an ALDH from S. cerevisiae, the activities of ALD4 were significantly elevated compared with that expressing ALD4 alone, suggesting that the regenerated NAD+ enhanced the activity of ALD4. More interestingly, the growth rates of all NAD+-regenerating strains were prolonged in comparison with the control, indicating that NAD+ regeneration stimulated cell proliferation. This study not only reveals the reliance of ALD4 activity on NAD+ availability but also provides a method for regulating the dha regulon.  相似文献   

16.
Perturbed metabolism of ammonia, an endogenous cytotoxin, causes mitochondrial dysfunction, reduced NAD+/NADH (redox) ratio, and postmitotic senescence. Sirtuins are NAD+-dependent deacetylases that delay senescence. In multiomics analyses, NAD metabolism and sirtuin pathways are enriched during hyperammonemia. Consistently, NAD+-dependent Sirtuin3 (Sirt3) expression and deacetylase activity were decreased, and protein acetylation was increased in human and murine skeletal muscle/myotubes. Global acetylomics and subcellular fractions from myotubes showed hyperammonemia-induced hyperacetylation of cellular signaling and mitochondrial proteins. We dissected the mechanisms and consequences of hyperammonemia-induced NAD metabolism by complementary genetic and chemical approaches. Hyperammonemia inhibited electron transport chain components, specifically complex I that oxidizes NADH to NAD+, that resulted in lower redox ratio. Ammonia also caused mitochondrial oxidative dysfunction, lower mitochondrial NAD+-sensor Sirt3, protein hyperacetylation, and postmitotic senescence. Mitochondrial-targeted Lactobacillus brevis NADH oxidase (MitoLbNOX), but not NAD+ precursor nicotinamide riboside, reversed ammonia-induced oxidative dysfunction, electron transport chain supercomplex disassembly, lower ATP and NAD+ content, protein hyperacetylation, Sirt3 dysfunction and postmitotic senescence in myotubes. Even though Sirt3 overexpression reversed ammonia-induced hyperacetylation, lower redox status or mitochondrial oxidative dysfunction were not reversed. These data show that acetylation is a consequence of, but is not the mechanism of, lower redox status or oxidative dysfunction during hyperammonemia. Targeting NADH oxidation is a potential approach to reverse and potentially prevent ammonia-induced postmitotic senescence in skeletal muscle. Since dysregulated ammonia metabolism occurs with aging, and NAD+ biosynthesis is reduced in sarcopenia, our studies provide a biochemical basis for cellular senescence and have relevance in multiple tissues.  相似文献   

17.
Cofactors cannot be retained within polyamide membrane microcapsules unless the cofactors have been covalently linked to macromolecules. In this paper, a new approach using lipid-polyamide membrane microcapsules has resulted in the retention of unmodified cofactors. Lipid-polyamide microcapsules can be made to contain urease (urea amidohydrolase, EC 3.5.1.5), glutamate dehydrogenase (NAD(P)+) [l-glutamate: NAD(P)+ oxidoreductase (deaminating), EC 1.4.1.3], alcohol dehydrogenase (alcohol:NAD+ oxidoreductase, EC 1.1.1.1), NAD+, NADH and α-ketoglutarate. Lipophilic substrates like ammonia can equilibrate rapidly into the microcapsules. The rate of conversion of ammonia into glutamate was studied. NAD+ retained in the microcapsules was effectively recycled into NADH and 0.25 μmol NAD+ converted 10 μmol ammonia into glutamate. Without cofactor recycling, 10 μmol NADH had to be microencapsulated to convert the same amount of ammonia into glutamate. By adjusting the ratio of cholesterol and lecithin in the lipid component of the membrane, it was also possible to achieve a good urea-permeable membrane without any leakage of cofactor or α-ketoglutarate. This way urea permeated through the lipid-polyamide membrane microcapsules was sequentially converted into ammonia and then glutamate.  相似文献   

18.
The introduction of an NADH/NAD+ regeneration system can regulate the distribution between acetoin and 2,3‐butanediol. NADH regeneration can also enhance butanol production in coculture fermentation. In this work, a novel artificial consortium of Paenibacillus polymyxa CJX518 and recombinant Escherichia coli LS02T that produces riboflavin (VB2) was used to regulate the NADH/NAD+ ratio and, consequently, the distribution of acetoin and 2,3‐butanediol by P. polymyxa. Compared with a pure culture of P. polymyxa, the level of acetoin was increased 76.7% in the P. polymyxa and recombinant E. coli coculture. Meanwhile, the maximum production and yield of acetoin in an artificial consortium with fed‐batch fermentation were 57.2 g/L and 0.4 g/g glucose, respectively. Additionally, the VB2 production of recombinant E. coli could maintain a relatively low NADH/NAD+ ratio by changing NADH dehydrogenase activity. It was also found that 2,3‐butanediol dehydrogenase activity was enhanced and improved acetoin production by the addition of exogenous VB2 or by being in the artificial consortium that produces VB2. These results illustrate that the coculture of P. polymyxa and recombinant E. coli has enormous potential to improve acetoin production. It was also a novel strategy to regulate the NADH/NAD+ ratio to improve the acetoin production of P. polymyxa.  相似文献   

19.
Anaerobic fermentation was relatively difficult to optimize due to lack of monitoring parameters. In this paper, a new method was reported using extracellular oxidoreduction potential (ORP) to monitor 1,3-propanediol (1,3-PD) biosynthesis process by Klebsiella pneumoniae. In batch fermentation, cell growth, 1,3-propanediol production and by-products distribution were studied at four different ORP levels: 10, −140, −190 and −240 mV. From the results, the ORP level of −190 mV was preferable, which resulted in fast cell growth and high 1,3-propanediol concentration. The NAD+/NADH ratio was determined at different ORP levels, and a critical NAD+/NADH ratio of 4 was defined to divide fermentation environments into two categories: relatively oxidative environment (NAD+/NADH>4) and relatively reductive environment (NAD+/NADH<4). The former was correlative with high 1,3-propanediol productivity and high specific growth rate. The mechanism of ORP regulation was discussed. It is suggested that ORP regulation of fermentation might be due to its influence on the ratio of NAD+/NADH, which determined metabolic flux. Furthermore, a batch fermentation of modulating ORP following a profile in different levels corresponding to different fermentation stage was tested. The 1,3-PD concentration was 22.3% higher than that of constant ORP fermentation at −190 mV. Therefore, ORP is a valuable parameter to monitor and control anaerobic fermentation production.  相似文献   

20.
In this study, the hydrogenase-3 gene cluster (hycDEFGH) was isolated and identified from Enterobacter aerogenes CCTCC AB91102. All gene products were highly homologous to the reported bacterial hydrogenase-3 (Hyd-3) proteins. The genes hycE, hycF, hycG encoding the subunits of hydrogenase-3 were targeted for genetic knockout to inhibit the FHL hydrogen production pathway via the Red recombination system, generating three mutant strains AB91102-E (ΔhycE), AB91102-F (ΔhycF) and AB91102-G (ΔhycG). Deletion of the three genes affected the integrity of hydrogenase-3. The hydrogen production experiments with the mutant strains showed that no hydrogen was detected compared with the wild type (0.886 mol/mol glucose), demonstrating that knocking out any of the three genes could inhibit NADH hydrogen production pathway. Meanwhile, the metabolites of the mutant strains were significantly changed in comparison with the wild type, indicating corresponding changes in metabolic flux by mutation. Additionally, the activity of NADH-mediated hydrogenase was found to be nil in the mutant strains. The chemostat experiments showed that the NADH/NAD+ ratio of the mutant strains increased nearly 1.4-fold compared with the wild type. The NADH-mediated hydrogenase activity and NADH/NAD+ ratio analysis both suggested that NADH pathway required the involvement of the electron transport chain of hydrogenase-3.  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