To see the other types of publications on this topic, follow the link: NADH.

Journal articles on the topic 'NADH'

Create a spot-on reference in APA, MLA, Chicago, Harvard, and other styles

Select a source type:

Consult the top 50 journal articles for your research on the topic 'NADH.'

Next to every source in the list of references, there is an 'Add to bibliography' button. Press on it, and we will generate automatically the bibliographic reference to the chosen work in the citation style you need: APA, MLA, Harvard, Chicago, Vancouver, etc.

You can also download the full text of the academic publication as pdf and read online its abstract whenever available in the metadata.

Browse journal articles on a wide variety of disciplines and organise your bibliography correctly.

1

Tsai, Hsieh-Chin, Cheng-Hung Hsieh, Ching-Wen Hsu, Yau-Heiu Hsu, and Lee-Feng Chien. "Cloning and Organelle Expression of Bamboo Mitochondrial Complex I Subunits Nad1, Nad2, Nad4, and Nad5 in the Yeast Saccharomyces cerevisiae." International Journal of Molecular Sciences 23, no. 7 (April 6, 2022): 4054. http://dx.doi.org/10.3390/ijms23074054.

Full text
Abstract:
Mitochondrial respiratory complex I catalyzes electron transfer from NADH to ubiquinone and pumps protons from the matrix into the intermembrane space. In particular, the complex I subunits Nad1, Nad2, Nad4, and Nad5, which are encoded by the nad1, nad2, nad4, and nad5 genes, reside at the mitochondrial inner membrane and possibly function as proton (H+) and ion translocators. To understand the individual functional roles of the Nad1, Nad2, Nad4, and Nad5 subunits in bamboo, each cDNA of these four genes was cloned into the pYES2 vector and expressed in the mitochondria of the yeast Saccharomyces cerevisiae. The mitochondrial targeting peptide mt gene (encoding MT) and the egfp marker gene (encoding enhanced green fluorescent protein, EGFP) were fused at the 5′-terminal and 3′-terminal ends, respectively. The constructed plasmids were then transformed into yeast. RNA transcripts and fusion protein expression were observed in the yeast transformants. Mitochondrial localizations of the MT-Nad1-EGFP, MT-Nad2-EGFP, MT-Nad4-EGFP, and MT-Nad5-EGFP fusion proteins were confirmed by fluorescence microscopy. The ectopically expressed bamboo subunits Nad1, Nad2, Nad4, and Nad5 may function in ion translocation, which was confirmed by growth phenotype assays with the addition of different concentrations of K+, Na+, or H+.
APA, Harvard, Vancouver, ISO, and other styles
2

Scholz, T. D., M. R. Laughlin, R. S. Balaban, V. V. Kupriyanov, and F. W. Heineman. "Effect of substrate on mitochondrial NADH, cytosolic redox state, and phosphorylated compounds in isolated hearts." American Journal of Physiology-Heart and Circulatory Physiology 268, no. 1 (January 1, 1995): H82—H91. http://dx.doi.org/10.1152/ajpheart.1995.268.1.h82.

Full text
Abstract:
The effect of metabolic substrates on the relation among cytosolic redox state (NADHc/NAD+) mitochondrial NADH (NADHm), and [ATP]/([ADP] x [Pi]) was studied in isolated working rabbit hearts. Substrates were varied from 5.6 mM glucose alone to glucose in combination with 10 mM lactate and/or 10 mM pyruvate while afterload and preload were held constant. Changes in NADHm were determined from epicardial NADH fluorescence. The ratio of glycerol 3-phosphate (G-3-P) to dihydroxyacetone phosphate (DHAP), determined from tissue extracts, was used as an index of cytosolic redox. Myocardial 31P metabolites were measured using nuclear magnetic resonance spectroscopy. The addition of pyruvate to the perfusion medium caused increases in myocardial oxygen consumption (MVo2), NADHm fluorescence, phosphocreatine (PCr), and [ATP]/([ADP] x [Pi]) and a decrease in NADHc/NADc+ (decreased G-3-P/DHAP). Although the addition of lactate to the perfusion medium caused an increase in NADHm similar to pyruvate, MVo2 and PCr did not change significantly, [ATP]/([ADP] x [Pi]) increased less than with pyruvate, and there was an increase in NADHc/NADc+. The findings suggest that variations in the cytosolic redox state do not necessarily result in obligatory changes in either the mitochondrial redox state or in the [ATP]/([ADP] x [Pi]). This implies that under the conditions of this study an equilibrium is not maintained between [ATP]/([ADP] x [Pi]) and NADHc/NADc+. Furthermore, similar levels of NADHm can be associated with different values for [ATP]/([ADP] x [Pi]) and MVo2, depending on the substrates available to the heart.
APA, Harvard, Vancouver, ISO, and other styles
3

Qi, Xiangying, Kaiqi Wang, Liping Yang, Zhenshan Deng, and Zhihong Sun. "The complete mitogenome sequence of the coral lily (Lilium pumilum) and the Lanzhou lily (Lilium davidii) in China." Open Life Sciences 15, no. 1 (December 31, 2020): 1060–67. http://dx.doi.org/10.1515/biol-2020-0102.

Full text
Abstract:
AbstractBackgroundThe mitogenomes of higher plants are conserved. This study was performed to complete the mitogenome of two China Lilium species (Lilium pumilum Redouté and Lilium davidii var. unicolor (Hoog) cotton).MethodsGenomic DNA was separately extracted from the leaves of L. pumilum and L. davidii in triplicate and used for sequencing. The mitogenome of Allium cepa was used as a reference. Genome assembly, annotation and phylogenetic tree were analyzed.ResultsThe mitogenome of L. pumilum and L. davidii was 988,986 bp and 924,401 bp in length, respectively. There were 22 core protein-coding genes (including atp1, atp4, atp6, atp9, ccmB, ccmC, ccmFc, ccmFN1, ccmFN2, cob, cox3, matR, mttB, nad1, nad2, nad3, nad4, nad4L, nad5, nad6, nad7 and nad9), one open reading frame and one ribosomal protein-coding gene (rps12) in the mitogenomes. Compared with the A. cepa mitogenome, the coding sequence of the 24 genes and intergenic spacers in L. pumilum and L. davidii mitogenome contained 1,621 and 1,617 variable sites, respectively. In the phylogenetic tree, L. pumilum and L. davidii were distinct from A. cepa (NC_030100).ConclusionsL. pumilum and L. davidii mitogenomes have far distances from other plants. This study provided additional information on the species resources of China Lilium.
APA, Harvard, Vancouver, ISO, and other styles
4

TURNER, William L., Jeffrey C. WALLER, and Wayne A. SNEDDEN. "Identification, molecular cloning and functional characterization of a novel NADH kinase from Arabidopsis thaliana (thale cress)." Biochemical Journal 385, no. 1 (December 14, 2004): 217–23. http://dx.doi.org/10.1042/bj20040292.

Full text
Abstract:
NADH kinase (NADHK; ATP:NADH 2′-phosphotransferase; EC 2.7.1.86), an enzyme that preferentially utilizes NADH as the diphosphonicotinamide nucleotide donor, has been identified for the first time in plants. Low activity (0.4 nmol of NADPH produced/min per mg of protein) was observed in clarified protein extracts from Arabidopsis thaliana (thale cress) cell suspension cultures. However, unlike an NADHK from yeast (Saccharomyces cerevisiae) (POS5), the enzyme from Arabidopsis did not associate with the mitochondria. NADHK was cloned (gi:30699338) from Arabidopsis and studied as a recombinant protein following affinity purification from Escherichia coli. The enzyme had a pH optimum for activity of 7.9 and a subunit molecular mass of 35 kDa. Analytical gel filtration demonstrated that the recombinant enzyme exists as a dimer. Hyperbolic saturation kinetics were observed for the binding of NADH, ATP, free Mg2+ and NAD+, with respective Km values of 0.042, 0.062, 1.16, and 2.39 mM. While NADHK could phosphorylate NADH or NAD+, the specificity constant (Vmax/Km) for NADH was 100-fold greater than for NAD+. The enzyme could utilize UTP, GTP and CTP as alternative nucleotides, although ATP was the preferred substrate. PPi or poly-Pi could not substitute as phospho donors. PPi acted as a mixed inhibitor with respect to both NADH and ATP. NADHK was inactivated by thiol-modifying reagents, with inactivation being decreased in the presence of NADH or ATP, but not NAD+. This study suggests that, in Arabidopsis, NADP+/NADPH biosynthetic capacity could, under some circumstances, become uncoupled from the redox status of the diphosphonicotinamide nucleotide pool.
APA, Harvard, Vancouver, ISO, and other styles
5

GAETANI, Gian F., Anna M. FERRARIS, Paola SANNA, and Henry N. KIRKMAN. "A novel NADPH:(bound) NADP+ reductase and NADH:(bound) NADP+ transhydrogenase function in bovine liver catalase." Biochemical Journal 385, no. 3 (January 24, 2005): 763–68. http://dx.doi.org/10.1042/bj20041495.

Full text
Abstract:
Many catalases have the shared property of containing bound NADPH and being susceptible to inactivation by their own substrate, H2O2. The presence of additional (unbound) NADPH effectively prevents bovine liver and human erythrocytic catalase from becoming compound II, the reversibly inactivated state of catalase, and NADP+ is known to be generated in the process. The function of the bound NADPH, which is tightly bound in bovine liver catalase, has been unknown. The present study with bovine liver catalase and [14C]NADPH and [14C]NADH revealed that unbound NADPH or NADH are substrates for an internal reductase and transhydrogenase reaction respectively; the unbound NADPH or NADH cause tightly bound NADP+ to become NADPH without becoming tightly bound themselves. This and other results provide insight into the function of tightly bound NADPH.
APA, Harvard, Vancouver, ISO, and other styles
6

Xia, Weiliang, Zheng Wang, Qing Wang, Jin Han, Cuiping Zhao, Yunyi Hong, Lili Zeng, Le Tang, and Weihai Ying. "Roles of NAD / NADH and NADP+ / NADPH in Cell Death." Current Pharmaceutical Design 15, no. 1 (January 1, 2009): 12–19. http://dx.doi.org/10.2174/138161209787185832.

Full text
APA, Harvard, Vancouver, ISO, and other styles
7

Nozato, Naoko, Kenji Oda, Katsuyuki Yamato, Eiji Ohta, Miho Takemura, Kinya Akashi, Hideya Fukuzawa, and Kanji Ohyama. "Cotranscriptional expression of mitochondrial genes for subunits of NADH dehydrogenase, nad5, nad4, nad2, in Marchantia polymorpha." Molecular and General Genetics MGG 237, no. 3 (March 1993): 343–50. http://dx.doi.org/10.1007/bf00279437.

Full text
APA, Harvard, Vancouver, ISO, and other styles
8

Griendling, Kathy K., and Masuko Ushio-Fukai. "NADH/NADPH Oxidase and Vascular Function." Trends in Cardiovascular Medicine 7, no. 8 (November 1997): 301–7. http://dx.doi.org/10.1016/s1050-1738(97)00088-1.

Full text
APA, Harvard, Vancouver, ISO, and other styles
9

Marbaix, Alexandre Y., Georges Chehade, Gaëtane Noël, Pierre Morsomme, Didier Vertommen, Guido T. Bommer, and Emile Van Schaftingen. "Pyridoxamine-phosphate oxidases and pyridoxamine-phosphate oxidase-related proteins catalyze the oxidation of 6-NAD(P)H to NAD(P)+." Biochemical Journal 476, no. 20 (October 28, 2019): 3033–52. http://dx.doi.org/10.1042/bcj20190602.

Full text
Abstract:
Abstract 6-NADH and 6-NADPH are strong inhibitors of several dehydrogenases that may form spontaneously from NAD(P)H. They are known to be oxidized to NAD(P)+ by mammalian renalase, an FAD-linked enzyme mainly present in heart and kidney, and by related bacterial enzymes. We partially purified an enzyme oxidizing 6-NADPH from rat liver, and, surprisingly, identified it as pyridoxamine-phosphate oxidase (PNPO). This was confirmed by the finding that recombinant mouse PNPO oxidized 6-NADH and 6-NADPH with catalytic efficiencies comparable to those observed with pyridoxine- and pyridoxamine-5′-phosphate. PNPOs from Escherichia coli, Saccharomyces cerevisiae and Arabidopsis thaliana also displayed 6-NAD(P)H oxidase activity, indicating that this ‘side-activity’ is conserved. Remarkably, ‘pyridoxamine-phosphate oxidase-related proteins’ (PNPO-RP) from Nostoc punctiforme, A. thaliana and the yeast S. cerevisiae (Ygr017w) were not detectably active on pyridox(am)ine-5′-P, but oxidized 6-NADH, 6-NADPH and 2-NADH suggesting that this may be their main catalytic function. Their specificity profiles were therefore similar to that of renalase. Inactivation of renalase and of PNPO in mammalian cells and of Ygr017w in yeasts led to the accumulation of a reduced form of 6-NADH, tentatively identified as 4,5,6-NADH3, which can also be produced in vitro by reduction of 6-NADH by glyceraldehyde-3-phosphate dehydrogenase or glucose-6-phosphate dehydrogenase. As 4,5,6-NADH3 is not a substrate for renalase, PNPO or PNPO-RP, its accumulation presumably reflects the block in the oxidation of 6-NADH. These findings indicate that two different classes of enzymes using either FAD (renalase) or FMN (PNPOs and PNPO-RPs) as a cofactor play an as yet unsuspected role in removing damaged forms of NAD(P).
APA, Harvard, Vancouver, ISO, and other styles
10

Marienfeld, J. R., and K. J. Newton. "The maize NCS2 abnormal growth mutant has a chimeric nad4-nad7 mitochondrial gene and is associated with reduced complex I function." Genetics 138, no. 3 (November 1, 1994): 855–63. http://dx.doi.org/10.1093/genetics/138.3.855.

Full text
Abstract:
Abstract The molecular basis of the maternally inherited, heteroplasmic NCS2 mutant of maize was investigated. Analysis of the NCS2 mtDNA showed that it closely resembles the progenitor cmsT mitochondrial genome, except that the mutant genome contains a fused nad4-nad7 gene and is deleted for the small fourth exon of nad4. The rearrangement has occurred at a 16-bp repeat present in the third intron of the nad4 gene and in the second intron of the nad7 gene. Transcripts containing exon 4 of the nad4 gene are greatly reduced in mtRNA preparations from heteroplasmic NCS2 plants; larger transcripts are associated with the first three nad4 exons. Identical 5' ends of the nad4 transcripts have been mapped 396 and 247 bp upstream of the start codon in mtRNAs from both NCS2 and related non-NCS plants. The putative transcription termination signal of nad4 is deleted in mutant DNA, resulting in the production of the unique longer transcripts. The complex transcript pattern associated with nad7 is also altered in the mutant. Both nad4 and nad7 encode subunits of complex I (NADH dehydrogenase) of the mitochondrial electron transfer chain. Oxygen uptake experiments show that the functioning of complex I is specifically reduced in mitochondria isolated from NCS2 mutant plants.
APA, Harvard, Vancouver, ISO, and other styles
11

Li, Yi, Jie Lin, Yong-Hui Wang, Ke Wang, Ren-Lei Wang, Xin-Chang Zhao, and Yi-Jian Yao. "Complete mitochondrial genome of Pleurocordyceps sinensis (Hypocreales, Ascomycota), a species with uncertain family-level taxonomic assignment." Quality Assurance and Safety of Crops & Foods 14, no. 4 (November 23, 2022): 212–26. http://dx.doi.org/10.15586/qas.v14i4.1134.

Full text
Abstract:
The complete mitochondrial (mt) genome of the ex-type strain of Pleurocordyceps sinensis, a fungus originally isolated from Ophiocordyceps sinensis, was sequenced, and assembled as a single circular DNA of 31,841 bp. The mt genome encoded 15 conserved proteins (rps3, cox1, cox2, cox3, cob, atp6, atp8, atp9, nad1, nad2, nad3, nad4, nad4L, nad5, and nad6), 2 rRNA (rnl and rns), and 25 tRNA, as well as 10 additional non-conserved open reading frames (ncORFs). Comparative analyses showed that mt genomes within the order Hypocreales encoded the same number and synteny of conserved protein coding genes despite an obvious size variation among this group of fungi. Phylogenetic analyses using 14 conserved protein sequences revealed that this fungus may not belong to the current designated family Ophiocordycipitaceae but is more closely related to the species of Clavicipitaceae. The mt genome presented herein would give valuable information on reconstructing the evolutionary history of clavicipitaceous fungi and also aid in resolving the family-level taxonomic assignment of Polycephalomyces s. l. species.
APA, Harvard, Vancouver, ISO, and other styles
12

Shidhi, Pattayampadam Ramakrishnan, Vadakkemukadiyil Chellappan Biju, Sasi Anu, Chandrasekharan Laila Vipin, Kumar Raveendran Deelip, and Sukumaran Nair Achuthsankar. "Genome Characterization, Comparison and Phylogenetic Analysis of Complete Mitochondrial Genome of Evolvulus alsinoides Reveals Highly Rearranged Gene Order in Solanales." Life 11, no. 8 (July 30, 2021): 769. http://dx.doi.org/10.3390/life11080769.

Full text
Abstract:
Mitogenome sequencing provides an understanding of the evolutionary mechanism of mitogenome formation, mechanisms driving plant gene order, genome structure, and migration sequences. Data on the mitochondrial genome for family Convolvulaceae members is lacking. E. alsinoides, also known as shankhpushpi, is an important medicinal plant under the family Convolvulaceae, widely used in the Ayurvedic system of medicine. We identified the mitogenome of E. alsinoides using the Illumina mate-pair sequencing platform, and annotated using bioinformatics approaches in the present study. The mitogenome of E. alsinoides was 344184 bp in length and comprised 46 unique coding genes, including 31 protein-coding genes (PCGs), 12 tRNA genes, and 3 rRNA genes. The secondary structure of tRNAs shows that all the tRNAs can be folded into canonical clover-leaf secondary structures, except three trnW, trnG, and trnC. Measurement of the skewness of the nucleotide composition showed that the AT and GC skew is positive, indicating higher A’s and G’s in the mitogenome of E. alsinoides. The Ka/Ks ratios of 11 protein-coding genes (atp1, ccmC, cob, cox1, rps19, rps12, nad3, nad9, atp9, rpl5, nad4L) were <1, indicating that these genes were under purifying selection. Synteny and gene order analysis were performed to identify homologous genes among the related species. Synteny blocks representing nine genes (nad9, nad2, ccmFc, nad1, nad4, nad5, matR, cox1, nad7) were observed in all the species of Solanales. Gene order comparison showed that a high level of gene rearrangement has occurred among all the species of Solanales. The mitogenome data obtained in the present study could be used as the Convolvulaceae family representative for future studies, as there is no complex taxonomic history associated with this plant.
APA, Harvard, Vancouver, ISO, and other styles
13

O'DONNELL, Valerie B., and Hartmut KÜHN. "Co-oxidation of NADH and NADPH by a mammalian 15-lipoxygenase: inhibition of lipoxygenase activity at near-physiological NADH concentrations." Biochemical Journal 327, no. 1 (October 1, 1997): 203–8. http://dx.doi.org/10.1042/bj3270203.

Full text
Abstract:
The purified 15-lipoxygenase from rabbit reticulocytes is capable of oxidizing NADH in the presence of linoleic acid and oxygen. This co-oxidation proceeds at a rate that amounts to approx. 7% of linoleic acid oxygenation rates. Although NADH inhibits the lipoxygenase reaction with linoleic acid as substrate (46% inhibition at 0.2 mM NADH), the reaction specificity of the enzyme was not altered since (13S)-hydroperoxy-(9Z,11E)-octadecadienoic acid was identified as the major reaction product. NADH oxidation was inhibited by NAD+ (uncompetitive with respect to linoleate and mixed/competitive with respect to NADH), and NADPH or NMNH could substitute for NADH with slightly different apparent Km values. NADH oxidation was enhanced at lower oxygen tension, but was completely prevented under anaerobic conditions. Computer-assisted modelling of 15-lipoxygenase/NADH interaction and sequence alignments of mammalian lipoxygenases with NADH-dependent enzymes suggested that there is no specific binding of the coenzyme at the putative fatty acid-binding site of lipoxygenases. These results suggest that NAD(P)H might be oxidized by a radical intermediate formed during the dioxygenase cycle of the lipoxygenase reaction but that NADH oxidation might not proceed at the active site of the enzyme. The mechanism and possible biological consequences of 15-lipoxygenase-catalysed NAD(P)H oxidation are discussed.
APA, Harvard, Vancouver, ISO, and other styles
14

Studničková, Marie, Hana Paulová-Klukanová, Jaroslav Turánek, and Jan Kovář. "Reduction of NAD+ and NADP+ and reductive cleavage of NADH and NADPH yielding NAD. and NADP." Journal of Electroanalytical Chemistry and Interfacial Electrochemistry 252, no. 2 (October 1988): 383–94. http://dx.doi.org/10.1016/0022-0728(88)80224-9.

Full text
APA, Harvard, Vancouver, ISO, and other styles
15

Thiagalingam, Sambasivamoorthy, and Tsanyen Yang. "NADH dehydrogenase and NADH oxidation inBacillus megaterium." Current Microbiology 14, no. 4 (July 1986): 217–20. http://dx.doi.org/10.1007/bf01568521.

Full text
APA, Harvard, Vancouver, ISO, and other styles
16

Šoba, Barbara, Špela Gašperšič, Darja Keše, and Tadeja Kotar. "Molecular Characterization of Echinococcus granulosus sensu lato from Humans in Slovenia." Pathogens 9, no. 7 (July 12, 2020): 562. http://dx.doi.org/10.3390/pathogens9070562.

Full text
Abstract:
The larval form of tapeworms of the Echinococcus granulosus sensu lato species cluster cause an important zoonotic infection, cystic echinococcosis (CE). Molecular characterization of the cluster’s isolates from different hosts greatly contributes to a better understanding of its transmission dynamics. To date, no genetic information is available on CE in Slovenia. In this work, we characterized isolates from human CE cases. Parasite samples from 18 patients were collected, together with the patients’ demographic and clinical data. Genomic DNA was analyzed by conventional PCR and sequencing at four mitochondrial loci (cytochrome c oxidase subunit 1, cox1; NADH dehydrogenase subunit 1, nad1; NADH dehydrogenase subunit 5, nad5; and small ribosomal RNA, rrnS). Thirteen isolates were successfully amplified and sequenced. Seven (58.8%) patients were infected with E. granulosus sensu stricto (s.s.) G1, five (38.5%) with E. canadensis G7 and one (7.7%) with E. granulosus s.s. G3. Echinococcus canadensis G7, the pig genotype, was identified exclusively in autochthonous Slovenes, while the patients originating from the Western Balkans were all infected with E. granulosus s.s. Our findings suggest that pigs are important intermediate hosts for human CE in Slovenia.
APA, Harvard, Vancouver, ISO, and other styles
17

Hedeskov, C. J., K. Capito, and P. Thams. "Cytosolic ratios of free [NADPH]/[NADP+] and [NADH]/[NAD+] in mouse pancreatic islets, and nutrient-induced insulin secretion." Biochemical Journal 241, no. 1 (January 1, 1987): 161–67. http://dx.doi.org/10.1042/bj2410161.

Full text
Abstract:
When the extracellular concentration of glucose was raised from 3 mM to 7 mM (the concentration interval in which beta-cell depolarization and the major decrease in K+ permeability occur), the cytosolic free [NADPH]/[NADP+] ratio in mouse pancreatic islets increased by 29.5%. When glucose was increased to 20 mM, a 117% increase was observed. Glucose had no effect on the cytosolic free [NADH]/[NAD+] ratio. Neither the cytosolic free [NADPH]/[NADP+] ratio nor the corresponding [NADH]/[NAD+] ratio was affected when the islets were incubated with 20 mM-fructose or with 3 mM-glucose + 20 mM-fructose, although the last-mentioned condition stimulated insulin release. The insulin secretagogue leucine (10 mM) stimulated insulin secretion, but lowered the cytosolic free [NADPH]/[NADP+] ratio; 10 mM-leucine + 10 mM-glutamine stimulated insulin release and significantly enhanced both the [NADPH]/[NADP+] ratio and the [NADH]/[NAD+] ratio. It is concluded that the cytosolic free [NADPH]/[NADP+] ratio may be involved in coupling beta-cell glucose metabolism to beta-cell depolarization and ensuing insulin secretion, but it may not be the sole or major coupling factor in nutrient-induced stimulation of insulin secretion.
APA, Harvard, Vancouver, ISO, and other styles
18

Matsushita, Kazunobu, Asuka Otofuji, Midori Iwahashi, Hirohide Toyama, and Osao Adachi. "NADH dehydrogenase of Corynebacterium glutamicum. Purification of an NADH dehydrogenase II homolog able to oxidize NADPH." FEMS Microbiology Letters 204, no. 2 (November 2001): 271–76. http://dx.doi.org/10.1111/j.1574-6968.2001.tb10896.x.

Full text
APA, Harvard, Vancouver, ISO, and other styles
19

Thiagalingam, Sam, and Tsanyen Yang. "Purification and characterization of NADH dehydrogenase from Bacillus megaterium." Canadian Journal of Microbiology 39, no. 9 (September 1, 1993): 826–33. http://dx.doi.org/10.1139/m93-123.

Full text
Abstract:
NADH dehydrogenase of Bacillus megaterium was isolated from the sonicate soluble fraction. The enzyme was purified approximately 61-fold by a combination of ammonium sulfate fractionation and column chromatography on DEAE-Sephadex and hydroxyapatite. The purified enzyme has an apparent molecular weight of 42 000 as determined by SDS-polyacrylamide gel electrophoresis and activity staining for NADH-MTT (3-(4,5-dimethylthiazol-2-yl)-2,5-diphenyl-tetrazolium bromide) oxidoreductase. The enzyme is specific for NADH and has a pH optimum of 7.5–7.8. The apparent Km values for NADH are 15.7, 34.8, and 69.2 μM as determined for NADH-DCIP (dichlorophenol–indophenol), NADH-ferricyanide, and NADH-MTT oxidoreductases. FAD is the prosthetic group of the enzyme. NAD+ acts as a competitive inhibitor. The inhibition studies suggest that NADH dehydrogenase is the primary electron donor of the NADH oxidase system. Localization studies and inhibition studies together indicate that the NADH oxidase is a complex of membrane-bound enzymes and coenzymes.Key words: NADH dehydrogenase, NADH oxidase, Bacillus megaterium, purification, characterization.
APA, Harvard, Vancouver, ISO, and other styles
20

Didion, Sean P., and Frank M. Faraci. "Effects of NADH and NADPH on superoxide levels and cerebral vascular tone." American Journal of Physiology-Heart and Circulatory Physiology 282, no. 2 (February 1, 2002): H688—H695. http://dx.doi.org/10.1152/ajpheart.00576.2001.

Full text
Abstract:
Reactive oxygen species are important modulators of cerebral vascular tone. Recent evidence, mainly from the aorta, suggests that NAD(P)H oxidase is a major source of vascular superoxide. The goal of the present study was to examine the effects of NADH and NADPH that are commonly used to stimulate NAD(P)H oxidase activity, on superoxide levels and cerebral vascular tone. Basilar arteries and cerebral arterioles from normal rabbits were studied in vitro using isolated tissue baths and in vivo using a cranial window, respectively. In the basilar artery, NADH produced a biphasic response; low concentrations (0.1–10 μM NADH) produced marked relaxation, whereas higher concentrations (30–100 μM NADH) produced contraction. Responses to NADH were significantly ( P < 0.05) inhibited in the presence of 4,5-dihydroxy-1,3-benzene-disulfonic acid (Tiron; a scavenger of superoxide, 10 mM). In contrast, NADPH (10–100 μM) produced moderate contraction of the basilar artery, which was inhibited in the presence of Tiron. In vivo, NADH produced Tiron-sensitive dilatation of cerebral arterioles. NADH and NADPH dose dependently increased superoxide levels in the basilar artery, as detected by lucigenin (5 μM)-enhanced chemiluminescence, but increases in superoxide were significantly greater for NADPH than NADH. These increases in superoxide were markedly reduced in the presence of polyethylene glycol-superoxide dismutase (300 U/ml) or diphenylene iodonium [0.1 mM, an inhibitor of flavin-containing enzymes, including NAD(P)H oxidase] but were not affected by indomethacin, N G-nitro-l-arginine, or allopurinol. These data suggest that NADH- and NADPH-induced changes in cerebral vascular tone are mediated by superoxide, produced by a flavin-containing enzyme, most likely NAD(P)H oxidase, but not xanthine oxidase or nitric oxide synthase.
APA, Harvard, Vancouver, ISO, and other styles
21

Cracan, Valentin, Denis Titov, Zenon Grabarek, and Vamsi Mootha. "Genetically Encoded Tools for Compartment-Specific Manipulation of NAD+/NADH or NADP+/NADPH Ratios." Free Radical Biology and Medicine 100 (November 2016): S182. http://dx.doi.org/10.1016/j.freeradbiomed.2016.10.479.

Full text
APA, Harvard, Vancouver, ISO, and other styles
22

Sonet, Gontran, Yannick De Smet, Min Tang, Massimiliano Virgilio, Andrew Donovan Young, Jeffrey H. Skevington, Ximo Mengual, et al. "First mitochondrial genomes of five hoverfly species of the genus Eristalinus (Diptera: Syrphidae)." Genome 62, no. 10 (October 2019): 677–87. http://dx.doi.org/10.1139/gen-2019-0009.

Full text
Abstract:
The hoverfly genus Eristalinus (Diptera, Syrphidae) contains many widespread pollinators. The majority of the species of Eristalinus occur in the Afrotropics and their molecular systematics still needs to be investigated. This study presents the first complete and annotated mitochondrial genomes for five species of Eristalinus. They were obtained by high-throughput sequencing of total genomic DNA. The total length of the mitogenomes varied between 15 757 and 16 245 base pairs. Gene composition, positions, and orientation were shared across species, and were identical to those observed for other Diptera. Phylogenetic analyses (maximum likelihood and Bayesian inference) based on the 13 protein coding and both rRNA genes suggested that the subgenus Eristalinus was paraphyletic with respect to the subgenus Eristalodes. An analysis of the phylogenetic informativeness of all protein coding and rRNA genes suggested that NADH dehydrogenase subunit 5 (nad5), cytochrome c oxidase subunit 1, nad4, nad2, cytochrome b, and 16S rRNA genes are the most promising mitochondrial molecular markers to result in supported phylogenetic hypotheses of the genus. In addition to the five complete mitogenomes currently available for hoverflies, the five mitogenomes published here will be useful for broader molecular phylogenetic analyses among hoverflies.
APA, Harvard, Vancouver, ISO, and other styles
23

Umrikhina, A. V., A. N. Luganskaya, and A. A. Krasnovsky. "ESR signals of NADH and NADPH under illumination." FEBS Letters 260, no. 2 (January 29, 1990): 294–96. http://dx.doi.org/10.1016/0014-5793(90)80127-5.

Full text
APA, Harvard, Vancouver, ISO, and other styles
24

Goodman, Russell P., Sarah E. Calvo, and Vamsi K. Mootha. "Spatiotemporal compartmentalization of hepatic NADH and NADPH metabolism." Journal of Biological Chemistry 293, no. 20 (March 7, 2018): 7508–16. http://dx.doi.org/10.1074/jbc.tm117.000258.

Full text
APA, Harvard, Vancouver, ISO, and other styles
25

Wang, Zheyuan, Weiwei Chen, Song Zhang, Jiawen Lu, Rongrong Chen, Junjie Fu, Riliang Gu, Guoying Wang, Jianhua Wang, and Yu Cui. "Dek504 Encodes a Mitochondrion-Targeted E+-Type Pentatricopeptide Repeat Protein Essential for RNA Editing and Seed Development in Maize." International Journal of Molecular Sciences 23, no. 5 (February 24, 2022): 2513. http://dx.doi.org/10.3390/ijms23052513.

Full text
Abstract:
In flowering plants, RNA editing is a post-transcriptional process that selectively deaminates cytidines (C) to uridines (U) in organellar transcripts. Pentatricopeptide repeat (PPR) proteins have been identified as site-specific recognition factors for RNA editing. Here, we report the map-based cloning and molecular characterization of the defective kernel mutant dek504 in maize. Loss of Dek504 function leads to delayed embryogenesis and endosperm development, which produce small and collapsed kernels. Dek504 encodes an E+-type PPR protein targeted to the mitochondria, which is required for RNA editing of mitochondrial NADH dehydrogenase 3 at the nad3-317 and nad3-44 sites. Biochemical analysis of mitochondrial protein complexes revealed a significant reduction in the mitochondrial NADH dehydrogenase complex I activity, indicating that the alteration of the amino acid sequence at nad3-44 and nad3-317 through RNA editing is essential for NAD3 function. Moreover, the amino acids are highly conserved in monocots and eudicots, whereas the events of C-to-U editing are not conserved in flowering plants. Thus, our results indicate that Dek504 is essential for RNA editing of nad3, which is critical for NAD3 function, mitochondrial complex I stability, and seed development in maize.
APA, Harvard, Vancouver, ISO, and other styles
26

Van Niel, Johannes C. G., and Upendra K. Pandit. "Nadh models." Tetrahedron 41, no. 24 (January 1985): 6005–11. http://dx.doi.org/10.1016/s0040-4020(01)91441-3.

Full text
APA, Harvard, Vancouver, ISO, and other styles
27

Alegria, Antonio E., Sheila Sanchez, Pedro Sanchez-Cruz, Ileana Nieves, Nadya G. Cruz, Marina Gordaliza, and Maria Luz Martín-Martín. "Terpenylnaphthoquinones are reductively activated by NADH/NADH dehydrogenase." Toxicological & Environmental Chemistry 87, no. 2 (April 2005): 237–45. http://dx.doi.org/10.1080/02772240400026872.

Full text
APA, Harvard, Vancouver, ISO, and other styles
28

STEHLE, Thilo, Al CLAIBORNE, and Georg E. SCHULZ. "NADH binding site and catalysis of NADH peroxidase." European Journal of Biochemistry 211, no. 1-2 (January 1993): 221–26. http://dx.doi.org/10.1111/j.1432-1033.1993.tb19889.x.

Full text
APA, Harvard, Vancouver, ISO, and other styles
29

Bibi, Shabana, Dong Wang, Yuanbing Wang, Ghazala Mustafa, and Hong Yu. "Mitogenomic and Phylogenetic Analysis of the Entomopathogenic Fungus Ophiocordyceps lanpingensis and Comparative Analysis with Other Ophiocordyceps Species." Genes 14, no. 3 (March 14, 2023): 710. http://dx.doi.org/10.3390/genes14030710.

Full text
Abstract:
Ophiocordyceps lanpingensis (O. lanpingensis) belongs to the genus Ophiocordyceps, which is often found in Yunnan Province, China. This species is pharmacologically important for the treatment of renal disorders induced by oxidative stress and an inadequate immune response. In the present study, the mitogenome of O. lanpingensis was determined to be a circular molecule 117,560 bp in length, and to have 31% G + C content and 69% A + T content. This mitogenome comprised 82% of the whole genome that codes for significant genes. The protein-coding regions of the O. lanpingensis mitogenome, containing 24 protein-coding genes, were associated with respiratory chain complexes, such as 3 ATP-synthase complex F0 subunits (atp6, atp8, and atp9), 2 complex IV subunits/cytochrome c oxidases (cox2 and cox3), 1 complex III subunit (cob), 4 electron transport complex I subunits/NADH dehydrogenase complex subunits (nad1, nad4, nad5, and nad6), 2 ribosomal RNAs (rns, rnl), and 11 hypothetical/predicted proteins, i.e., orf609, orf495, orf815, orf47, orf150, orf147, orf292, orf127, orf349, orf452, and orf100. It was noted that all genes were positioned on the same strand. Further, 13 mitochondrial genes with respiratory chain complexes, which presented maximum similarity with other fungal species of Ophiocordyceps, were investigated. O. lanpingensis was compared with previously sequenced species within Ophiocordycepitaceae. Comparative analysis indicated that O. lanpingensis was more closely related to O. sinensis, which is one of the most remarkable and expensive herbs due to its limited availability and the fact that it is difficult to culture. Therefore, O. lanpingensis is an important medicinal resource that can be effectively used for medicinal purposes. More extensive metabolomics research is recommended for O. lanpingensis.
APA, Harvard, Vancouver, ISO, and other styles
30

Pfeifer, Tom A., Dwayne D. Hegedus, and George G. Khachatourians. "The mitochondrial genome of the entomopathogenic fungus Beauveria bassiana: analysis of the ribosomal RNA region." Canadian Journal of Microbiology 39, no. 1 (January 1, 1993): 25–31. http://dx.doi.org/10.1139/m93-004.

Full text
Abstract:
The 28.5-kbp mitochondrial (mt) genome from the entomopathogenic fungus Beauveria bassiana was studied using restriction enzyme analysis, gene probe hybridization, and DNA sequence comparisons. A detailed restriction enzyme map allowed cloning of the entire genome into a number of segments. Hybridization of heterologous gene probes to the mtDNA resulted in the identification of the large ribosomal RNA (lrRNA) and small ribosomal RNA (srRNA) genes. Gene probes derived from several yeasts and fungi failed to identify any additional genes. However, partial DNA sequence analysis revealed the lrRNA and srRNA genes as well as four protein-encoding genes: the NADH dehydrogenase subunit 1 (NAD1), NADH dehydrogenase subunit 6 (NAD6), cytochrome oxidase subunit 3 (C03), and ATPase subunit 6 (ATP6) genes. The ATPase subunit 9 (ATP9) gene was not identified by hybridization to mtDNA, but could be detected by hybridization to total cellular DNA. The portions of the genes sequenced were homologous to the equivalent genes from yeast and other filamentous fungi, most notably Aspergillus nidulans. No introns were identified in these regions. The organization of the sequenced region of the B. bassiana mt genome more closely resembled that of A. nidulans than that of Podospora anserina or Neurospora crassa.Key words: Beauveria bassiana, mtDNA, gene mapping, mitochondrial rRNA, mitochondrial organization.
APA, Harvard, Vancouver, ISO, and other styles
31

Bao, Teng, Xian Zhang, Xiaojing Zhao, Zhiming Rao, Taowei Yang, and Shangtian Yang. "Regulation of the NADH pool and NADH/NADPH ratio redistributes acetoin and 2,3-butanediol proportion inBacillus subtilis." Biotechnology Journal 10, no. 8 (August 2015): 1298–306. http://dx.doi.org/10.1002/biot.201400577.

Full text
APA, Harvard, Vancouver, ISO, and other styles
32

Wos, M. L., and P. C. Pollard. "Cellular nicotinamide adenine dinucleotide (NADH) as an indicator of bacterial metabolic activity dynamics in activated sludge." Water Science and Technology 60, no. 3 (July 1, 2009): 783–91. http://dx.doi.org/10.2166/wst.2009.393.

Full text
Abstract:
In this study, native fluorescent nicotinamide adenine dinucleotide (NADH) was used as a direct indicator of bacterial metabolic activity in activated sludge. Specific NADH concentration was dynamic and varied between 106–108 molecules per bacterial cell. Low concentrations (106–107 NADH molecules cell−1) indicate efficient bacterial metabolic activity while high concentrations (107–108 NADH molecules cell−1) indicate inefficient bacterial metabolic activity. Specific [NADH] did not correlate to changes in dissolved organic carbon, but increases correlated to decreases in oxygen uptake rates. Perhaps a lack of oxygen as the terminal electron acceptor prevented efficient reoxidization of NADH to NAD+, which resulted in an accumulation of NADH within the cells. Also, significant amounts of NADH were released and accumulated into the extracellular medium of metabolically active E. coli cells in log phase. Such overflow metabolism may be the product of favourable conditions. Thus, the flux of both specific intracellular and extracellular [NADH] indicates the dynamics of bacterial metabolic activity in activated sludge.
APA, Harvard, Vancouver, ISO, and other styles
33

Rasmussen, U. F., and H. N. Rasmussen. "The NADH oxidase system (external) of muscle mitochondria and its role in the oxidation of cytoplasmic NADH." Biochemical Journal 229, no. 3 (August 1, 1985): 631–41. http://dx.doi.org/10.1042/bj2290631.

Full text
Abstract:
An exo-NADH oxidase system [NADH oxidase system (external)], effecting intact-mitochondrial oxidation of added NADH, was studied in pigeon heart mitochondria. Breast muscle mitochondria showed an equal specific activity of the system. The exo-NADH oxidase activity (200 micron mol of NADH/min per g of protein) equalled two-thirds of the State-3 respiratory activity with malate + pyruvate or one-seventh of the total NADH oxidase activity of heart mitochondria. The activity was not caused by use of proteinase in the preparation procedure and all measured parameters were very reproducible from preparation to preparation. The activity is therefore most likely not due to preparation artefacts. The exo-NADH oxidase system is present in all mitochondria in the preparation and is not confined to a subpopulation. The system reduced all cytochrome anaerobically and direct interaction with all cytochrome oxidase was demonstrated by interdependent cyanide inhibition. The exo-NADH oxidase system seems to be located at the outer surface of the mitochondrial inner membrane because, for instance, only this system was rapidly inhibited by rotenone, and ferricyanide could act as acceptor in the rotenone-inhibited system (reductase activity = 20 times oxidase activity). In the presence of antimycin, added NADH reduced only a part of the b-cytochromes. Freezing and thawing the mitochondria, one of the methods used for making them permeable to NADH, destroyed this functional compartmentation. The characteristics of the exo-NADH oxidase system and the malate-aspartate shuttle are compared and the evidence for the shuttle's function in heart in vivo is re-evaluated. It is proposed that oxidation of cytoplasmic NADH in red muscles primarily is effected by the exo-NADH oxidase system.
APA, Harvard, Vancouver, ISO, and other styles
34

Ying, Weihai. "NAD+/NADH and NADP+/NADPH in Cellular Functions and Cell Death: Regulation and Biological Consequences." Antioxidants & Redox Signaling 10, no. 2 (February 2008): 179–206. http://dx.doi.org/10.1089/ars.2007.1672.

Full text
APA, Harvard, Vancouver, ISO, and other styles
35

Riebe, Oliver, Ralf-Jörg Fischer, David A. Wampler, Donald M. Kurtz, and Hubert Bahl. "Pathway for H2O2 and O2 detoxification in Clostridium acetobutylicum." Microbiology 155, no. 1 (January 1, 2009): 16–24. http://dx.doi.org/10.1099/mic.0.022756-0.

Full text
Abstract:
An unusual non-haem diiron protein, reverse rubrerythrin (revRbr), is known to be massively upregulated in response to oxidative stress in the strictly anaerobic bacterium Clostridium acetobutylicum. In the present study both in vivo and in vitro results demonstrate an H2O2 and O2 detoxification pathway in C. acetobutylicum involving revRbr, rubredoxin (Rd) and NADH : rubredoxin oxidoreductase (NROR). RevRbr exhibited both NADH peroxidase (NADH : H2O2 oxidoreductase) and NADH oxidase (NADH : O2 oxidoreductase) activities in in vitro assays using NROR as the electron-transfer intermediary from NADH to revRbr. Rd increased the NADH consumption rate by serving as an intermediary electron-transfer shuttle between NROR and revRbr. While H2O2 was found to be the preferred substrate for revRbr, its relative oxidase activity was found to be significantly higher than that reported for other Rbrs. A revRbr-overexpressing strain of C. acetobutylicum showed significantly increased tolerance to H2O2 and O2 exposure. RevRbr thus appears to protect C. acetobutylicum against oxidative stress by functioning as the terminal component of an NADH peroxidase and NADH oxidase.
APA, Harvard, Vancouver, ISO, and other styles
36

Vinogradov, Andrei D. "NADH/NAD+ interaction with NADH: Ubiquinone oxidoreductase (complex I)." Biochimica et Biophysica Acta (BBA) - Bioenergetics 1777, no. 7-8 (July 2008): 729–34. http://dx.doi.org/10.1016/j.bbabio.2008.04.014.

Full text
APA, Harvard, Vancouver, ISO, and other styles
37

Kucera, I., L. Lampardová, and V. Dadák. "Control of respiration rate in non-growing cells of Paracoccus denitrificans." Biochemical Journal 246, no. 3 (September 15, 1987): 779–82. http://dx.doi.org/10.1042/bj2460779.

Full text
Abstract:
By means of fluorimetric measurement and by direct determination of intracellular NAD+ and NADH contents, it was proved that the respiration rate of Paracoccus denitrificans cells utilizing glucose is limited by processes preceding NADH oxidation in the respiratory chain, so that the membrane NADH dehydrogenase is not saturated by its substrate. In the separated membrane fraction on saturation with exogenous NADH the main limiting factor is represented by NADH: ubiquinone oxidoreductase.
APA, Harvard, Vancouver, ISO, and other styles
38

Kovář, Jan, Jaromír Stejskal, Hana Paulová, and Jiří Slavík. "Reduction of quaternary benzophenanthridine alkaloids by NADH and NADPH." Collection of Czechoslovak Chemical Communications 51, no. 11 (1986): 2626–34. http://dx.doi.org/10.1135/cccc19862626.

Full text
Abstract:
Spectroscopic techniques have shown that quaternary benzophenanthridine alkaloids (sanguinarine (Ia), chelerythrine (Ie), sanguilutine (If), sanguirubine (Ic), chelirubine (Ib), and chelilutine (Id)) are reduced by NADH or NADPH to their dihydro forms. The formation of the oxidized form of a coenzyme was demonstrated by HPLC. The kinetics of the reaction of sanguinarine with NADH has been studied rather in detail; the reaction was found to be reversible, its stoichiometry was 1 : 1. Reactivity of the tested alkaloids toward the coenzymes was given by reactivity of the imine bond.This was graded like pK values for the formation of the corresponding pseudobases and correlated with the energy of the lowest unoccupied molecular orbital of these compounds. The possible biological and pharmacological importance of the studied reactions is discussed.
APA, Harvard, Vancouver, ISO, and other styles
39

Overkamp, Karin M., Barbara M. Bakker, Peter Kötter, Arjen van Tuijl, Simon de Vries, Johannes P. van Dijken, and Jack T. Pronk. "In Vivo Analysis of the Mechanisms for Oxidation of Cytosolic NADH by Saccharomyces cerevisiaeMitochondria." Journal of Bacteriology 182, no. 10 (May 15, 2000): 2823–30. http://dx.doi.org/10.1128/jb.182.10.2823-2830.2000.

Full text
Abstract:
ABSTRACT During respiratory glucose dissimilation, eukaryotes produce cytosolic NADH via glycolysis. This NADH has to be reoxidized outside the mitochondria, because the mitochondrial inner membrane is impermeable to NADH. In Saccharomyces cerevisiae, this may involve external NADH dehydrogenases (Nde1p or Nde2p) and/or a glycerol-3-phosphate shuttle consisting of soluble (Gpd1p or Gpd2p) and membrane-bound (Gut2p) glycerol-3-phosphate dehydrogenases. This study addresses the physiological relevance of these mechanisms and the possible involvement of alternative routes for mitochondrial oxidation of cytosolic NADH. Aerobic, glucose-limited chemostat cultures of agut2Δ mutant exhibited fully respiratory growth at low specific growth rates. Alcoholic fermentation set in at the same specific growth rate as in wild-type cultures (0.3 h−1). Apparently, the glycerol-3-phosphate shuttle is not essential for respiratory glucose dissimilation. An nde1Δ nde2Δmutant already produced glycerol at specific growth rates of 0.10 h−1 and above, indicating a requirement for external NADH dehydrogenase to sustain fully respiratory growth. An nde1Δ nde2Δ gut2Δ mutant produced even larger amounts of glycerol at specific growth rates ranging from 0.05 to 0.15 h−1. Apparently, even at a low glycolytic flux, alternative mechanisms could not fully replace the external NADH dehydrogenases and glycerol-3-phosphate shuttle. However, at low dilution rates, thende1Δ nde2Δ gut2Δ mutant did not produce ethanol. Since glycerol production could not account for all glycolytic NADH, another NADH-oxidizing system has to be present. Two alternative mechanisms for reoxidizing cytosolic NADH are discussed: (i) cytosolic production of ethanol followed by its intramitochondrial oxidation and (ii) a redox shuttle linking cytosolic NADH oxidation to the internal NADH dehydrogenase.
APA, Harvard, Vancouver, ISO, and other styles
40

Hòa, Lê Thanh, Nguyễn Thị Khuê, Nguyễn Thị Bích Nga, Đỗ Thị Roan, Đỗ Trung Dũng, Lê Thị Kim Xuyến, and Đoàn Thị Thanh Hương. "Genetic characterization of mitochondrial genome of the small intestinal fluke, Haplorchis taichui (Trematoda: Heterophyidae), Vietnamese sample." Vietnam Journal of Biotechnology 14, no. 2 (June 30, 2016): 215–24. http://dx.doi.org/10.15625/1811-4989/14/2/9333.

Full text
Abstract:
The small intestinal fluke, Haplorchis taichui Nishigori, 1924, belonging to genus Haplorchis (family Heterophyidae, class Trematoda, phylum Platyhelminthes), is a zoonotic pathogen causing disease in humans and animals. Complete mitochondrial genome (mtDNA) of H. taichui (strain HTAQT, collected from Quang Tri) was obtained and characterized for structural genomics providing valuable data for studies on epidemiology, species identification, diagnosis, classification, molecular phylogenetic relationships and prevention of the disease. The entire nucleotide mtDNA sequence of H. taichui (HTAQT) is 15.119 bp in length, containing 36 genes, including 12 protein-coding genes (cox1, cox2, cox3, nad1, nad2, nad3, nad4L, nad4, nad5, nad6, atp6 and cob); 2 ribosomal RNA genes, rrnL (16S) and rrnS (12S); 22 transfer RNA genes (tRNA or trn), and a non-coding region (NR), divided into two sub-regions of short non-coding (short, SNR) and long non-coding (long, LNR). LNR region, 1.692 bp in length, located between the position of trnG (transfer RNA-Glycine) and trnE (Glutamic acid), contains 6 tandem repeats (TR), arranged as TR1A, TR2A, TR1B, TR2B, TR3A, TR3B, respectively. Each protein coding gene (overall, 12 genes), ribosomal rRNA (2 genes) and tRNA (22 genes) were analyzed, in particular, protein-coding genes were defined in length, start and stop codons, and rRNA and tRNA genes for secondary structure.
APA, Harvard, Vancouver, ISO, and other styles
41

Wiederkehr, Andreas, and Nicolas Demaurex. "Illuminating redox biology using NADH- and NADPH-specific sensors." Nature Methods 14, no. 7 (July 2017): 671–72. http://dx.doi.org/10.1038/nmeth.4336.

Full text
APA, Harvard, Vancouver, ISO, and other styles
42

Seno, Manabu, Meiling Lin, and Kazutoshi Iwamoto. "Chromatographic behaviour of cyclodextrin complexes of NADH and NADP." Journal of Chromatography A 508 (January 1990): 127–32. http://dx.doi.org/10.1016/s0021-9673(00)91245-7.

Full text
APA, Harvard, Vancouver, ISO, and other styles
43

Blacker, Thomas S., and Michael R. Duchen. "Investigating mitochondrial redox state using NADH and NADPH autofluorescence." Free Radical Biology and Medicine 100 (November 2016): 53–65. http://dx.doi.org/10.1016/j.freeradbiomed.2016.08.010.

Full text
APA, Harvard, Vancouver, ISO, and other styles
44

Li, Fei-Long, Qiang Zhou, Wei Wei, Jian Gao, and Ye-Wang Zhang. "Switching the substrate specificity from NADH to NADPH by a single mutation of NADH oxidase from Lactobacillus rhamnosus." International Journal of Biological Macromolecules 135 (August 2019): 328–36. http://dx.doi.org/10.1016/j.ijbiomac.2019.05.146.

Full text
APA, Harvard, Vancouver, ISO, and other styles
45

Lopez de Felipe, Felix, Michiel Kleerebezem, Willem M. de Vos, and Jeroen Hugenholtz. "Cofactor Engineering: a Novel Approach to Metabolic Engineering in Lactococcus lactis by Controlled Expression of NADH Oxidase." Journal of Bacteriology 180, no. 15 (August 1, 1998): 3804–8. http://dx.doi.org/10.1128/jb.180.15.3804-3808.1998.

Full text
Abstract:
ABSTRACT NADH oxidase-overproducing Lactococcus lactis strains were constructed by cloning the Streptococcus mutans nox-2gene, which encodes the H2O-forming NADH oxidase, on the plasmid vector pNZ8020 under the control of the L. lactis nisA promoter. This engineered system allowed a nisin-controlled 150-fold overproduction of NADH oxidase at pH 7.0, resulting in decreased NADH/NAD ratios under aerobic conditions. Deliberate variations on NADH oxidase activity provoked a shift from homolactic to mixed-acid fermentation during aerobic glucose catabolism. The magnitude of this shift was directly dependent on the level of NADH oxidase overproduced. At an initial growth pH of 6.0, smaller amounts of nisin were required to optimize NADH oxidase overproduction, but maximum NADH oxidase activity was twofold lower than that found at pH 7.0. Nonetheless at the highest induction levels, levels of pyruvate flux redistribution were almost identical at both initial pH values. Pyruvate was mostly converted to acetoin or diacetyl via α-acetolactate synthase instead of lactate and was not converted to acetate due to flux limitation through pyruvate dehydrogenase. The activity of the overproduced NADH oxidase could be increased with exogenously added flavin adenine dinucleotide. Under these conditions, lactate production was completely absent. Lactate dehydrogenase remained active under all conditions, indicating that the observed metabolic effects were only due to removal of the reduced cofactor. These results indicate that the observed shift from homolactic to mixed-acid fermentation under aerobic conditions is mainly modulated by the level of NADH oxidation resulting in low NADH/NAD+ratios in the cells.
APA, Harvard, Vancouver, ISO, and other styles
46

Liu, W., G. H. Liu, F. Li, D. S. He, T. Wang, X. F. Sheng, D. L. Zeng, F. F. Yang, and Y. Liu. "Sequence variability in three mitochondrial DNA regions of Spirometra erinaceieuropaei spargana of human and animal health significance." Journal of Helminthology 86, no. 3 (July 20, 2011): 271–75. http://dx.doi.org/10.1017/s0022149x1100037x.

Full text
Abstract:
AbstractSequence variability in three mitochondrial DNA (mtDNA) regions, namely cytochrome c oxidase subunit 3 (cox3), NADH dehydrogenase subunits 1 and 4 (nad1 and nad4) in Spirometra erinaceieuropaei spargana from different geographical regions in China was examined. A portion of each of the cox3 (pcox3), nad1 (pnad1) and nad4 genes (pnad4) were amplified separately from individual S. erinaceieuropaei spargana by polymerase chain reaction (PCR). Representative amplicons were subjected to sequencing in order to estimate sequence variability. The sequences of pcox3, pnad1 and pnad4 were 541, 607 and 847 bp in length, respectively. The A+T contents of the sequences were 68.39–68.76% (pcox3), 63.76–64.91% (pnad1) and 67.18–67.77% (pnad4), respectively, while the intra-specific sequence variations within each of the S. erinaceieuropaei spargana were 0–1.5% for pcox3, 0–2.8% for pnad1 and 0–2.7% for pnad4. Phylogenetic analysis using neighbour joining (NJ), maximum likelihood (ML) and maximum parsimony (MP) methods, indicated that all the spargana isolates in Hunan Province represented S. erinaceieuropaei. These findings demonstrated clearly the usefulness of the three mtDNA sequences for population genetics studies of S. erinaceieuropaei spargana of human and animal health significance.
APA, Harvard, Vancouver, ISO, and other styles
47

Liu, G. H., B. Li, J. Y. Li, H. Q. Song, R. Q. Lin, X. Q. Cai, F. C. Zou, et al. "Genetic variation among Clonorchis sinensis isolates from different geographic regions in China revealed by sequence analyses of four mitochondrial genes." Journal of Helminthology 86, no. 4 (December 12, 2011): 479–84. http://dx.doi.org/10.1017/s0022149x11000757.

Full text
Abstract:
AbstractThe present study examined sequence variation in four mitochondrial (mt) genes, namely cytochrome c oxidase subunits 1 (cox1) and 2 (cox2), and NADH dehydrogenase subunits 1 and 2 (nad1 and nad2) among Clonorchis sinensis isolates from different endemic regions in China, and their phylogenetic relationships with other zoonotic trematodes were reconstructed. A portion of the cox1 and cox2 genes (pcox1 and pcox2), and nad1 and nad2 genes (pnad1 and pnad2) were amplified separately from individual liver flukes by polymerase chain reaction (PCR) and the amplicons were subjected to sequencing from both directions. The intra-specific sequence variations within C. sinensis were 0–1.6% for pcox1, 0–1.4% for pcox2, 0–0.9% for pnad1 and 0–1.0% for pnad2. Phylogenetic analyses based on the combined sequences of pcox1, pcox2, pnad1 and pnad2 revealed that all the C. sinensis isolates grouped together and were closely related to Opisthorchis felineus. These findings revealed the existence of intra-specific variation in mitochondrial DNA (mtDNA) sequences among C. sinensis isolates from different geographic regions, and demonstrated that mtDNA sequences provide reliable genetic markers for phylogenetic studies of zoonotic trematodes.
APA, Harvard, Vancouver, ISO, and other styles
48

Shu, Xiaohan, Zekai Li, Ruizhong Yuan, Pu Tang, and Xuexin Chen. "Novel Gene Rearrangements in the Mitochondrial Genomes of Cynipoid Wasps (Hymenoptera: Cynipoidea)." Genes 13, no. 5 (May 20, 2022): 914. http://dx.doi.org/10.3390/genes13050914.

Full text
Abstract:
Cynipoidea is a medium-sized superfamily of Hymenoptera with diverse lifestyles. In this study, 16 mitochondrial genomes were newly sequenced, 11 of which were the first obtained mitochondrial genomes in the family Liopteridae and four subfamilies (Anacharitinae, Aspicerinae, Figitinae, and Parnipinae) of Figitidae. All of the newly sequenced mitogenomes have unique rearrangement types within Cynipoidea, whereas some gene patterns are conserved in several groups. nad5-nad4-nad4L-nad6-cytb was remotely inverted and two rRNA genes were translocated to nad3 downstream in Ibaliidae and three subfamilies (Anacharitinae, Eucoilinae, and Parnipinae within Figitidae); two rRNA genes in Aspicerinae, Figitinae, and Liopteridae were remotely inverted to the cytb-nad1 junction; rrnL-rrnS was translocated to the cytb-nad1 junction in Cynipidae. Phylogenetic inference suggested that Figitidae was a polyphyletic group, while the Ibaliidae nested deep within Cynipoidea and was a sister-group to the Figitidae. These results will improve our understanding of the gene rearrangement of the mitogenomes and the phylogenetic relationships in the Cynipoidea.
APA, Harvard, Vancouver, ISO, and other styles
49

Saba, Tony, Joseph W. H. Burnett, Jianwei Li, Panagiotis N. Kechagiopoulos, and Xiaodong Wang. "A facile analytical method for reliable selectivity examination in cofactor NADH regeneration." Chemical Communications 56, no. 8 (2020): 1231–34. http://dx.doi.org/10.1039/c9cc07805c.

Full text
APA, Harvard, Vancouver, ISO, and other styles
50

Small, W. Curtis, and Lee McAlister-Henn. "Identification of a Cytosolically Directed NADH Dehydrogenase in Mitochondria of Saccharomyces cerevisiae." Journal of Bacteriology 180, no. 16 (August 15, 1998): 4051–55. http://dx.doi.org/10.1128/jb.180.16.4051-4055.1998.

Full text
Abstract:
ABSTRACT The reoxidation of NADH generated in reactions within the mitochondrial matrix of Saccharomyces cerevisiae is catalyzed by an NADH dehydrogenase designated Ndi1p (C. A. M. Marres, S. de Vries, and L. A. Grivell, Eur. J. Biochem. 195:857–862, 1991). Gene disruption analysis was used to examine possible metabolic functions of two proteins encoded by open reading frames having significant primary sequence similarity to Ndi1p. Disruption of the gene designated NDH1 results in a threefold reduction in total mitochondrial NADH dehydrogenase activity in cells cultivated with glucose and in a fourfold reduction in the respiration of isolated mitochondria with NADH as the substrate. Thus, Ndh1p appears to be a mitochondrial dehydrogenase capable of using exogenous NADH. Disruption of a closely related gene designated NDH2 has no effect on these properties. Growth phenotype analyses suggest that the external NADH dehydrogenase activity of Ndh1p is important for optimum cellular growth with a number of nonfermentable carbon sources, including ethanol. Codisruption of NDH1 and genes encoding malate dehydrogenases essentially eliminates growth on nonfermentable carbon sources, suggesting that the external mitochondrial NADH dehydrogenase and the malate-aspartate shuttle may both contribute to reoxidation of cytosolic NADH under these growth conditions.
APA, Harvard, Vancouver, ISO, and other styles
We offer discounts on all premium plans for authors whose works are included in thematic literature selections. Contact us to get a unique promo code!

To the bibliography