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 0 毫秒
1.
Theoretical calculations of the NADPH requirement for biomass formation indicate that in yeasts this parameter is strongly dependent on the carbon and nitrogen sources used for growth. Enzyme surveys of NADPH-generating metabolic pathways and radiorespirometric studies demonstrate that in yeasts the HMP pathway is the major source of NADPH. Furthermore, radiorespirometric data suggest that in yeasts the HMP pathway activities are close to the theoretical minimum. It may be concluded that the mitochondrial NADPH oxidation, which in yeasts may yield ATP, is quantitatively not an important process.The inability of C. utilis to utilize the NADH produced in formate oxidation as an extra source of NADPH strongly suggests that transhydrogenase activity is absent. Furthermore, the absence of xylose utilization under anaerobic conditions in most facultatively fermentative yeasts indicates that also in these organisms transhydrogenase activity is absent. This conclusion is supported by the observation that anaerobic xylose utilization is observed only in those yeasts which possess a high activity of an NADH-linked xylose reductase. Hence in these organisms the redox-neutral conversion of xylose to ethanol is possible, since the second step in xylose metabolism is mediated by an NAD+-linked xylitol dehydrogenase.This paper is adapted from a treatise by the same author, entitled: The NADP(H) redox couple in yeast metabolism, that was awarded the Kluyver prize 1986 by the Netherlands Society of Microbiology  相似文献   

2.
Glutamate dehydrogenase (GDH) from vertebrates is unusual among NAD(P)H-dependent dehydrogenases in that it can use either NAD(H) or NADP(H) as cofactor. In this study, we measure the rate of cofactor utilization by bovine GDH when both cofactors are present. Methods for both reaction directions were developed, and for the first time, to our knowledge, the GDH activity has been simultaneously studied in the presence of both NAD(H) and NADP(H). Our data indicate that NADP(H) has inhibitory effects on the rate of NAD(H) utilization by GDH, a characteristic of GDH not previously recognized. The response of GDH to allosteric activators in the presence of NAD(H) and NADP(H) suggests that ADP and leucine moderate much of the inhibitory effect of NADP(H) on the utilization of NAD(H). These results illustrate that simple assumptions of cofactor preference by mammalian GDH are incomplete without an appreciation of allosteric effects when both cofactors are simultaneously present.  相似文献   

3.
The functions of NAD(H) (NAD(+) and NADH) and NADP(H) (NADP(+) and NADPH) are undoubtedly significant and distinct. Hence, regulation of the intracellular balance of NAD(H) and NADP(H) is important. The key enzymes involved in the regulation are NAD kinase and NADP phosphatase. In 2000, we first succeeded in identifying the gene for NAD kinase, thereby facilitating worldwide studies of this enzyme from various organisms, including eubacteria, archaea, yeast, plants, and humans. Molecular biological study has revealed the physiological function of this enzyme, that is to say, the significance of NADP(H), in some model organisms. Structural research has elucidated the tertiary structure of the enzyme, the details of substrate-binding sites, and the catalytic mechanism. Research on NAD kinase also led to the discovery of archaeal NADP phosphatase. In this review, we summarize the physiological functions, applications, and structure of NAD kinase, and the way we discovered archaeal NADP phosphatase.  相似文献   

4.
For better understanding of the coenzyme specificity in NAD-dependent MDH (tMDH) from Thermus flavus AT-62, we determined the crystal structures of tMDH-NADP(H) complex at maximally 1.65 A resolution. The overall structure is almost the same as that of the tMDH-NADH complex. However, NADP(H) binds to tMDH in the reverse orientation, where adenine occupies the position near the catalytic center and nicotinamide is positioned at the adenine binding site of the tMDH-NADH complex. Consistent with this, kinetic analysis of the malate-oxidizing reaction revealed that NADP(+) inhibited tMDH at high concentrations. This has provided the first evidence for the alternative binding mode of the nicotinamide coenzyme, that has pseudo-symmetry in its structure, in a single enzyme.  相似文献   

5.
Using the purified NADP(H)-binding domain of proton-translocating Escherichia coli transhydrogenase (ecIII) overexpressed in (15)N- and (2)H-labeled medium, together with the purified NAD(H)-binding domain from E. coli (ecI), the interface between ecIII and ecI, the NADP(H)-binding site and the influence on the interface by NAD(P)(H) was investigated in solution by NMR chemical shift mapping. Mapping of the NADP(H)-binding site showed that the NADP(H) substrate is bound to ecIII in an extended conformation at the C-terminal end of the parallel beta-sheet. The distribution of chemical shift perturbations in the NADP(H)-binding site, and the nature of the interaction between ecI and ecIII, indicated that the nicotinamide moiety of NADP(H) is located near the loop comprising residues P346-G353, in agreement with the recently determined crystal structures of bovine [Prasad, G. S., et al. (1999) Nat. Struct. Biol. 6, 1126-1131] and human heart [White, A. W., et al. (2000) Structure 8, 1-12] transhydrogenases. Further chemical shift perturbation analysis also identified regions comprising residues G389-I406 and G430-V434 at the C-terminal end of ecIII's beta-sheet as part of the ecI-ecIII interface, which were regulated by the redox state of the NAD(P)(H) substrates. To investigate the role of these loop regions in the interaction with domain I, the single cysteine mutants T393C, R425C, G430C, and A432C were generated in ecIII and the transhydrogenase activities of the resulting mutant proteins characterized using the NAD(H)-binding domain I from Rhodospirillum rubrum (rrI). All mutants except R425C showed altered NADP(H) binding and domain interaction properties. In contrast, the R425C mutant showed almost exclusively changes in the NADP(H)-binding properties, without changing the affinity for rrI. Finally, by combining the above conclusions with information obtained by a further characterization of previously constructed mutants, the implications of the findings were considered in a mechanistic context.  相似文献   

6.
The amphibian enzyme ADH8, previously named class IV-like, is the only known vertebrate alcohol dehydrogenase (ADH) with specificity towards NADP(H). The three-dimensional structures of ADH8 and of the binary complex ADH8-NADP(+) have been now determined and refined to resolutions of 2.2A and 1.8A, respectively. The coenzyme and substrate specificity of ADH8, that has 50-65% sequence identity with vertebrate NAD(H)-dependent ADHs, suggest a role in aldehyde reduction probably as a retinal reductase. The large volume of the substrate-binding pocket can explain both the high catalytic efficiency of ADH8 with retinoids and the high K(m) value for ethanol. Preference of NADP(H) appears to be achieved by the presence in ADH8 of the triad Gly223-Thr224-His225 and the recruitment of conserved Lys228, which define a binding pocket for the terminal phosphate group of the cofactor. NADP(H) binds to ADH8 in an extended conformation that superimposes well with the NAD(H) molecules found in NAD(H)-dependent ADH complexes. No additional reshaping of the dinucleotide-binding site is observed which explains why NAD(H) can also be used as a cofactor by ADH8. The structural features support the classification of ADH8 as an independent ADH class.  相似文献   

7.
When plants experience an imbalance between the absorption of light energy and the use of that energy to drive metabolism, they are liable to suffer from oxidative stress. Such imbalances arise due to environmental conditions (e.g. heat, chilling or drought), and can result in the production of reactive oxygen species (ROS). Here, we present evidence for a novel protective process - feedback redox regulation via the redox poise of the NADP(H) pool. Photosynthetic electron transport was studied in two transgenic tobacco (Nicotiana tabacum) lines - one having reduced levels of ferredoxin NADP+-reductase (FNR), the enzyme responsible for reducing NADP+, and the other reduced levels of glyceraldehyde 3-phosphate dehydrogenase (GAPDH), the principal consumer of NADPH. Both had a similar degree of inhibition of carbon fixation and impaired electron transport. However, whilst FNR antisense plants were obviously stressed, with extensive bleaching of leaves, GAPDH antisense plants showed no visible signs of stress, beyond having a slowed growth rate. Examination of electron transport in these plants indicated that this difference is due to feedback regulation occurring in the GAPDH but not the FNR antisense plants. We propose that this reflects the occurrence of a previously undescribed regulatory pathway responding to the redox poise of the NADP(H) pool.  相似文献   

8.
In this report, we use synthetic, activity-variant alleles in Drosophila melanogaster to quantify interactions across the enzyme network that reduces nicotinamide adenine dinucleotide phosphate (NADP) to NADPH. We examine the effects of large-scale variation in isocitrate dehydrogenase (IDH) or glucose-6-phosphate dehydrogenase (G6PD) activity in a single genetic background and of smaller-scale variation in IDH, G6PD, and malic enzyme across 10 different genetic backgrounds. We find significant interactions among all three enzymes in adults; changes in the activity of any one source of a reduced cofactor generally result in changes in the other two, although the magnitude and directionality of change differs depending on the gene and the genetic background. Observed interactions are presumably through cellular mechanisms that maintain a homeostatic balance of NADPH/NADP, and the magnitude of change in response to modification of one source of reduced cofactor likely reflects the relative contribution of that enzyme to the cofactor pool. Our results suggest that malic enzyme makes the largest single contribution to the NADPH pool, consistent with the results from earlier experiments in larval D. melanogaster using naturally occurring alleles. The interactions between all three enzymes indicate functional interdependence and underscore the importance of examining enzymes as components of a network.IN traits determined by a network of gene products, the phenotype is a function of the alleles present and of the relative contributions of individual network member genes. Since selection is on phenotype, the total composite genotype, not just individual loci, determines the fitness of an organism. In establishing the connection between genotype and phenotype for such networks, the first challenge is to quantify the relative contribution of each member of the network to the endpoint phenotype. By addressing function on a network-wide basis, interactions and interconnections that may not be apparent in individual gene examinations can be determined (Proulx et al. 2005).In most organisms, reduction of the cofactor nicotinamide adenine dinucleotide phosphate, or NADP, to NADPH is primarily the function of four enzymes: cytosolic malic enzyme (MEN), cytosolic isocitrate dehydrogenase (IDH), and the two oxidative enzymes of the pentose shunt, glucose-6-phosphate dehydrogenase (G6PD) and 6-phosphogluconate (6PGD; recently reviewed in Ying 2008). In larval Drosophila melanogaster, MEN produces ∼30% of the available NADPH, IDH ∼20%, and G6PD and 6PGD the remaining ∼40% (Geer et al. 1979a,b). It is believed that these four enzymes interact to maintain the NADP/NADPH balance and supply of reducing power for lipogenesis and antioxidation (Geer et al. 1976, 1978, 1981; Wilton et al. 1982; Bentley et al. 1983; Geer and Laurie-Ahlberg 1984; Merritt et al. 2005; Pollak et al. 2007; Singh et al. 2007; Ying 2008). Dietary induction studies and observations of natural genetic variation have found connections between MEN activity and the activities of the pentose shunt enzymes to be generally straightforward and compensatory; reductions in one lead to increases in the other. The interactions involving IDH activity, however, have been found to be more complicated and at times counterintuitive; reductions in reducing power sometimes lead to decreases in IDH activity.In an earlier study (Merritt et al. 2005), we quantified the impact of genetic variation in Men activity on IDH and G6PD activities and triglyceride (a strong correlate with total lipid; Clark and Keith 1989) concentration. 6PGD was not independently assayed because earlier works suggest that G6PD and 6PGD activities are highly correlated, likely because of their coupled function in the pentose shunt (Wilton et al. 1982). We examined both naturally occurring Men alleles and synthetic alleles created by P-element excision and found significant associations between MEN activity and induction of the activities of both IDH and G6PD. The apparent interactions between MEN and IDH and G6PD across these 10 different third chromosome lines were quantified as mean elasticity coefficients: = −0.76 ± 0.236 and = −0.88 ± 0.208. Because MEN activity was reduced by 20%, both IDH and G6PD activity varied in a compensatory direction, increasing almost 1:1 with the decrease in MEN.The significant change in enzyme activity of two members of the NADPH network in response to our genetic reduction of the activity of a third strongly suggests that a physiological mechanism coregulates the three enzymes. Such functional interdependence would mean that individual members of the network do not act in isolation and should be examined collectively, not as isolated units. In this study, we characterize the effects of the independently varying activity levels of IDH, G6PD, and MEN on the activity of each other and triglyceride concentration in adult flies. We found significant responses to changes in all three enzymes, although the responses to genetic changes in IDH and G6PD were generally small; variation in MEN caused the greatest changes in the other enzymes.  相似文献   

9.
Simon Hald 《BBA》2008,1777(5):433-440
When plants experience an imbalance between the absorption of light energy and the use of that energy to drive metabolism, they are liable to suffer from oxidative stress. Such imbalances arise due to environmental conditions (e.g. heat, chilling or drought), and can result in the production of reactive oxygen species (ROS). Here, we present evidence for a novel protective process — feedback redox regulation via the redox poise of the NADP(H) pool. Photosynthetic electron transport was studied in two transgenic tobacco (Nicotiana tabacum) lines — one having reduced levels of ferredoxin NADP+-reductase (FNR), the enzyme responsible for reducing NADP+, and the other reduced levels of glyceraldehyde 3-phosphate dehydrogenase (GAPDH), the principal consumer of NADPH. Both had a similar degree of inhibition of carbon fixation and impaired electron transport. However, whilst FNR antisense plants were obviously stressed, with extensive bleaching of leaves, GAPDH antisense plants showed no visible signs of stress, beyond having a slowed growth rate. Examination of electron transport in these plants indicated that this difference is due to feedback regulation occurring in the GAPDH but not the FNR antisense plants. We propose that this reflects the occurrence of a previously undescribed regulatory pathway responding to the redox poise of the NADP(H) pool.  相似文献   

10.
Escherichia coli cells from strain fpr, deficient in the soxRS-induced ferredoxin (flavodoxin)-NADP(H) reductase (FPR), display abnormal sensitivity to the bactericidal effects of the superoxide-generating reagent methyl viologen (MV). Neither bacteriostatic effects nor inactivation of oxidant-sensitive hydrolyases could be detected in fpr cells exposed to MV. FPR inactivation did not affect the MV-driven soxRS response, whereas FPR overexpression led to enhanced stimulation of the regulon, with concomitant oxidation of the NADPH pool. Accumulation of a site-directed FPR mutant that uses NAD(H) instead of NADP(H) had no effect on soxRS induction and failed to protect fpr cells from MV toxicity, suggesting that FPR contributes to NADP(H) homeostasis in stressed bacteria.  相似文献   

11.
Abstract We have directly selected thermosensitive alleles of cya and crp genes from the wild-type, and present preliminary characterization of these mutants. The selection procedure is based on prior growth of a mutagenized wild-type culture in a medium that counterselects, at low temperature, non-conditional relative to thermosensitive mutants, followed by routine selection of mutants at high temperature. This method should be applicable to various genetic systems.  相似文献   

12.
The purification and partial characterisation of an NADP(H) dependent artificial mediator accepting pyridine nucleotide oxidoreductase (AMAPOR) from the anaerobic Clostridium thermoaceticum is described. Depending on the redox potential of the artificial mediators the AMAPOR is able to regenerate NADP+ or NADPH rendering the enzyme useful for preparative work applying NADP(H) dependent oxidoreductases. At 37 degrees C crude extracts of C. thermoaceticum have an AMAPOR activity of 5-7 U mg(-1). This is 28 degrees under the optimal growth temperature of this microrganism. Out of apparently more than 10 AMAPOR active proteins in the crude cell extracts visible after electrophoresis and activity staining on the gel, two of these proteins were isolated. They seem to be two different oligomers. According to gel electrophoresis they show apparent molecular masses of about 200 and 400 kDa. These two forms showed after SDS gel electrophoresis two monomers with apparent molecular masses of 42 and 56 kDa which we call alpha and beta. The two oligomers may have the compositions alpha2beta2 and alpha4beta4. They contain Fe/S cluster and FAD. Various amounts of the FAD were lost during the purification procedure. This loss is partially reversible after addition of FAD. The AMAPOR reacts with rather different artificial mediators such as viologens, quinones e.g. 1,4-benzoquinone or anthraquinone-2,6-disulphonate, 2,6-dichloro-indophenol and clostridial rubredoxin. Two different ferredoxins from C. thermoaceticum, oxygen or lipoamide are no substrates indicating the here described AMAPOR is not a diaphorase in the usual sense.  相似文献   

13.
14.
A 2-Keto-L-gulonic acid (2-KLG) production process using stationary Pantoea citrea cells and a Corynebacterium 2,5-diketo-D-gluconic acid (2,5-DKG) reductase enzyme has been developed which may represent an improved method of vitamin C biosynthesis. Experimental data was collected using the F22Y/A272G 2,5-DKG reductase mutant and NADP(H) as a cofactor. An extensive kinetic analysis was performed and a kinetic rate equation model for this process was developed. A recent protein engineering effort has resulted in several 2,5-DKG reductase mutants exhibiting improved activity with NADH as a cofactor. The use of NAD(H) in the bioreactor may be preferable due to its increased stability and lower cost. The kinetic parameters in the rate equation model have been replaced in order to predict 2-KLG production with NAD(H) as a cofactor. The model was also extended to predict 2-KLG production in the presence of a range of combined cofactor concentrations. This analysis suggests that the use of the F22Y/K232G/R238H/A272G 2,5-DKG reductase mutant with NAD(H) combined with a small amount of NADP(H) could provide a significant cost benefit for in vitro enzymatic 2-KLG production.  相似文献   

15.
16.
17.
Proton-translocating transhydrogenase (TH) couples direct and stereospecific hydride transfer between NAD(H) and NADP(H), bound to soluble domains dI and dIII, respectively, to proton translocation across a membrane bound domain, dII. The reaction occurs with proton-gradient coupled conformational changes, which affect the energetics of substrate binding and interdomain interactions. The crystal structure of TH dIII from Rhodospirillum rubrum has been determined in the presence of NADPH (2.4 A) and NADP (2.1 A) (space group P6(1)22). Each structure has two molecules in the asymmetric unit, differing in the conformation of the NADP(H) binding loop D. In one molecule, loop D has an open conformation, with the B face of (dihydro)nicotinamide exposed to solvent. In the other molecule, loop D adopts a hitherto unobserved closed conformation, resulting in close interactions between NADP(H) and side chains of the highly conserved residues, betaSer405, betaPro406, and betaIle407. The conformational change shields the B face of (dihydro)nicotinamide from solvent, which would block hydride transfer in the intact enzyme. It also alters the environments of invariant residues betaHis346 and betaAsp393. However, there is little difference in either the open or the closed conformation upon change in oxidation state of nicotinamide, i.e., for NADP vs. NADPH. Consequently, the occurrence of two loop D conformations for both substrate oxidation states gives rise to four states: NADP-open, NADP-closed, NADPH-open, and NADPH-closed. Because these states are distinguished by protein conformation and by net charge they may be important in the proton translocating mechanism of intact TH.  相似文献   

18.
即时浸酸显著提高滞育性家蚕卵辅酶Ⅰ和Ⅱ含量   总被引:1,自引:0,他引:1  
即时浸酸在阻止家蚕Bombyx mori卵滞育发动的同时, 提高了其呼吸耗氧量, 抑制了山梨醇积累。本研究利用HPLC法测定了家蚕滞育卵和5 min即时浸酸滞育性卵中辅酶Ⅰ和Ⅱ含量。结果表明: 产下后24-72 h, 家蚕滞育卵中NAD, NADH, NADP和NADPH含量分别下降了30%, 37%, 50%和4%; 而即时浸酸滞育性卵中分别增加了77%, 46%, 142%和241%。不过, 即时浸酸并未显著改变滞育性家蚕卵中NADH/NAD和NADPH/NADP比值。据此推测, 即时浸酸提高滞育性家蚕卵辅酶Ⅰ含量与其呼吸耗氧量增加有关; 即时浸酸显著提高辅酶Ⅱ含量与山梨醇积累抑制无关, 而主要与生物合成加强有关。  相似文献   

19.
Summary A sulfonated polysulfone membrane reactor was used forin situ regeneration and retention of coenzymes NADP (H) using the xylose reductase ofCandida pelliculosa coupled with oxidoreductase system ofMethanobacterium sp. in the reduction of xylose to xylitol with hydrogen gas. The membrane could almost completely reject the permeation of NADP (H) (92 and 97%), F420 (97%) and the required enzymes (100%), but not reject for the permeation of xylitol (product). After 4-h reaction for the production of xylitol from xylose (93% yield), although 25% NADP (H) initially added was lost its activity due to unavoidable degradation, the membrane could reject the permeation of the remaining NADP (H) and F420 at the level of 90 and 95%, respectively.  相似文献   

20.
Ferredoxin-NADP(H) reductases catalyse the reversible hydride/electron exchange between NADP(H) and ferredoxin/flavodoxin, comprising a structurally defined family of flavoenzymes with two distinct subclasses. Those present in Gram-negative bacteria (FPRs) display turnover numbers of 1-5 s(-1) while the homologues of cyanobacteria and plants (FNRs) developed a 100-fold activity increase. We investigated nucleotide interactions and hydride transfer in Rhodobacter capsulatus FPR comparing them to those reported for FNRs. NADP(H) binding proceeds as in FNRs with stacking of the nicotinamide on the flavin, which resulted in formation of charge-transfer complexes prior to hydride exchange. The affinity of FPR for both NADP(H) and 2'-P-AMP was 100-fold lower than that of FNRs. The crystal structure of FPR in complex with 2'-P-AMP and NADP(+) allowed modelling of the adenosine ring system bound to the protein, whereas the nicotinamide portion was either not visible or protruding toward solvent in different obtained crystals. Stabilising contacts with the active site residues are different in the two reductase classes. We conclude that evolution to higher activities in FNRs was partially favoured by modification of NADP(H) binding in the initial complexes through changes in the active site residues involved in stabilisation of the adenosine portion of the nucleotide and in the mobile C-terminus of FPR.  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