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.
《BBA》2020,1861(3):148140
Among the thioredoxin reductase-type ferredoxin-NAD(P)+ oxidoreductase (FNR) family, FNR from photosynthetic purple non‑sulfur bacterium Rhodopseudomonas palustris (RpFNR) is distinctive because the predicted residue on the re-face of the isoalloxazine ring portion of the FAD prosthetic group is a tyrosine. Here, we report the crystal structure of wild type RpFNR and kinetic analyses of the reaction of wild type, and Y328F, Y328H and Y328S mutants with NADP+/NADPH using steady state and pre-steady state kinetic approaches.The obtained crystal structure of wild type RpFNR confirmed the presence of Tyr328 on the re-face of the isoalloxazine ring of the FAD prosthetic group through the unique hydrogen bonding of its hydroxyl group. In the steady state assays, the substitution results in the decrease of Kd for NADP+ and KM for NADPH in the diaphorase assay; however, the kcat values also decreased significantly. In the stopped-flow spectrophotometry, mixing oxidized RpFNRs with NADPH and reduced RpFNRs with NADP+ resulted in rapid charge transfer complex formation followed by hydride transfer. The observed rate constants for the hydride transfer in both directions were comparable (>400 s−1). The substitution did not drastically affect the rate of hydride transfer, but substantially slowed down the subsequent release and re-association of NADP+/NADPH in both directions. The obtained results suggest that Tyr328 stabilizes the stacking of C-terminal residues on the isoalloxazine ring portion of the FAD prosthetic group, which impedes the access of NADP+/NADPH on the isoalloxazine ring portions, in turn, enhancing the release of the NADP+/NADPH and/or reaction with electron transfer proteins.  相似文献   

2.
Rhodococcus (opacus) erythropolis HL PM-1 grows on 2,4,6-trinitrophenol or 2,4-dinitrophenol (2,4-DNP) as a sole nitrogen source. The NADPH-dependent F420 reductase (NDFR; encoded by npdG) and the hydride transferase II (HTII; encoded by npdI) of the strain were previously shown to convert both nitrophenols to their respective hydride Meisenheimer complexes. In the present study, npdG and npdI were amplified from six 2,4-DNP degrading Rhodococcus spp. The genes showed sequence similarities of 86 to 99% to the respective npd genes of strain HL PM-1. Heterologous expression of the npdG and npdI genes showed that they were involved in 2,4-DNP degradation. Sequence analyses of both the NDFRs and the HTIIs revealed conserved domains which may be involved in binding of NADPH or F420. Phylogenetic analyses of the NDFRs showed that they represent a new group in the family of F420-dependent NADPH reductases. Phylogenetic analyses of the HTIIs revealed that they form an additional group in the family of F420-dependent glucose-6-phosphate dehydrogenases and F420-dependent N5,N10-methylenetetrahydromethanopterin reductases. Thus, the NDFRs and the HTIIs may each represent a novel group of F420-dependent enzymes involved in catabolism.  相似文献   

3.
The steady state kinetics of glyceraldehyde 3-phosphate:NADP+ oxidoreductase (GNR) (EC 1.2.1.9) have been investigated. The enzyme exhibits hyperbolic behavior over a wide range of substrate concentrations. Double-reciprocal plots are nearly parallel or distantly convergent with limiting Km values of 2 to 5 micromolar for NADP+ and 20 to 40 micromolar for D-glyceraldehyde 3-phosphate (G3P). The velocity response to NADP+ as the varied substrate is however sigmoidal if G3P concentration exceeds 10 micromolar, whereas the response to G3P may show inhibition above this concentration. This `G3P-inhibited state' is alleviated by saturating amounts of NADP+ or NADPH. Product inhibition patterns indicate NADPH as a potent competitive inhibitor to NADP+ (Ki 30 micromolar) and mixed inhibitor towards G3P, and 3-phosphoglycerate (3PGA) as mixed inhibitor to both NADP+ and G3P (Ki 10 millimolar). The data, and those obtained with dead-end inhibitors, are consistent with a nonrapid equilibrium random mechanism with two alternative kinetic pathways. Of these, a rapid kinetic sequence (probably ordered with NADP+ binding first and G3P binding as second substrate) is dominant in the range of hyperbolic responses. A reverse reaction with 3PGA and NADPH as substrates is unlikely, and was not detected. Of a number of compounds tested, erythrose 4-phosphate (Ki 7 micromolar) and Pi (Ki 2.4 millimolar) act as competitive inhibitors to G3P (uncompetitive towards NADP+) and are likely to affect the in vivo activity. Ribose 5-phosphate, phosphoenolpyruvate, ATP, and ADP are also somewhat inhibitory. Full GNR activity in the leaf seems to be allowed only under high photosynthesis conditions, when levels of several inhibitors are low and substrate is high. We suggest that a main function of leaf GNR is to supply NADPH required for photorespiration, the reaction product 3PGA being cycled back to chloroplasts.  相似文献   

4.
Glucose-6-phosphate dehydrogenase from Leuconostoc mesenteroides utilizes either NAD+ or NADP+ as coenzyme. Kinetic studies showed that NAD+ and NADP+ interact with different enzyme forms (Olive, C., Geroch, M. E., and Levy, H. R. (1971) J. Biol. Chem.246, 2047–2057). In the present study the techniques of fluorescence quenching and fluorescence enhancement were used to investigate the interaction between Leuconostoc mesenteroides glucose-6-phosphate dehydrogenase and coenzymes. In addition, kinetic studies were performed to examine interaction between the enzyme and various coenzyme analogs. The maximum quenching of protein fluorescence is 5% for NADP+ and 50% for NAD+. The dissociation constant for NADP+, determined from fluorescence quenching measurements, is 3 μm, which is similar to the previously determined Km of 5.7 μm and Ki of 5 μm. The dissociation constant for NAD+ is 2.5 mm, which is 24 times larger than the previously determined Km of 0.106 mm. Glucose 1-phosphate, a substrate-competitive inhibitor, lowers the dissociation constant and maximum fluorescence quenching for NAD+ but not for NADP+. This suggests that glucose 6-phosphate may act similarly and thus play a role in enabling the enzyme to utilize NAD+ under physiological conditions. When NADPH binds to the enzyme its fluorescence is enhanced 2.3-fold. The enzyme was titrated with NADPH in the absence and presence of NAD+; binding of these two coenzymes is competitive. The dissociation constant for NADPH from these measurements is 24 μm; the previously determined Ki is 37.6 μm. The dissociation constant for NAD′ is 2.8 mm, in satisfactory agreement with the value obtained from protein fluorescence quenching measurements. Various compounds which resemble either the adenosine or the nicotinamide portion of the coenzyme structure are coenzyme-competitive inhibitors; 2′,5′-ADP, the most inhibitory analog tested, gives NADP+-competitive and NAD+-noncompetitive inhibition, consistent with the kinetic mechanism previously proposed. By using pairs of coenzyme-competitive inhibitors it was shown in kinetic studies that the two portions of the NAD+ structure cannot be accommodated on the enzyme simultaneously unies they are covalently linked. Fluorescence studies showed that there are both “buried” and “exposed” tryptophan residues in the enzyme structure.  相似文献   

5.
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   

6.
Ferredoxin-NAD(P)+ reductase ([EC 1.18.1.2], [EC 1.18.1.3]) from Chlorobaculum tepidum (CtFNR) is structurally homologous to the bacterial NADPH-thioredoxin reductase (TrxR), but possesses a unique C-terminal extension relative to TrxR that interacts with the isoalloxazine ring moiety of the flavin adenine dinucleotide prosthetic group. In this study, we introduce truncations to the C-terminal residues to examine their role in the reactions of CtFNR with NADP+ and NADPH by spectroscopic and kinetic analyses. The truncation of the residues from Tyr326 to Glu360 (the whole C-terminal extension region), from Phe337 to Glu360 (omitting Phe337 on the re-face of the isoalloxazine ring) and from Ser338 to Glu360 (leaving Phe337 intact) resulted in a blue-shift of the flavin absorption bands. The truncations caused a slight increase in the dissociation constant toward NADP+ and a slight decrease in the Michaelis constant toward NADPH in steady-state assays. Pre-steady-state studies of the redox reaction with NADPH demonstrated that deletions of Tyr326–Glu360 decreased the hydride transfer rate, and the amount of reduced enzyme increased at equilibrium relative to wild-type CtFNR. In contrast, the deletions of Phe337–Glu360 and Ser338–Glu360 resulted in only slight changes in the reaction kinetics and redox equilibrium. These results suggest that the C-terminal region of CtFNR is responsible for the formation and stability of charge-transfer complexes, leading to changes in redox properties and reactivity toward NADP+/NADPH.  相似文献   

7.
A sonicate of Achromobacter parvulus IFO-13182 produced NADPH from NADP+by an NADP+-linked malic enzyme [l-malate: NAD(P)+oxidoreductase, EC 1.1.1.39–40] reaction in the presence of l-malic acid and divalent metal ions. Malic enzyme of A. parvulus was stabilized by 5% l-malic acid, and activity was maintained at 60°C for 1 h. Contaminating phosphatase (orthophosphoricmonoester phosphohydrolase, EC 3.1.3.1–2) was completely inactivated by this treatment. Among the conditions tested, the optimum NADPH production was done using 36 μmol NADP+, 67 μmol l-malic acid, 63 μmol MgCl2 and 1 unit of the malic enzyme in 3 ml of 55 mm phosphate buffer (pH 7.8). Conversion ratio of NADPH from NADP+ reached 100% after 4 h incubation at 30°C and the amount of NADPH accumulated was ~12 μmol ml?1of the reaction mixture. No dephosphorylation of NADP+to NAD+or of NADPH to NADH was found by high performance liquid chromatography. The NADPH produced by such enzymatic reduction was purified by ethanol precipitation and dried in vacuo in powdered form with 97% purity, judged from the ratio of the absorbances at 340 and 260 nm. The purity of the NADPH produced was determined to be 95% from its coenzyme activity with NAD(P)+-linked glutathione reductase [NAD(P)H: oxidized-glutathione oxidoreductase, EC 1.6.4.2].  相似文献   

8.
In photosynthetic organisms, ferredoxin:NADP+ oxidoreductase (FNR) is known to provide NADPH for CO2 assimilation, but it also utilizes NADPH to provide reduced ferredoxin. The cyanobacterium Synechocystis sp. strain PCC6803 produces two FNR isoforms, a small one (FNRS) similar to the one found in plant plastids and a large one (FNRL) that is associated with the phycobilisome, a light-harvesting complex. Here we show that a mutant lacking FNRL exhibits a higher NADP+/NADPH ratio. We also purified to homogeneity a phycobilisome subcomplex comprising FNRL, named FNRL-PC. The enzymatic activities of FNRL-PC were compared with those of FNRS. During NADPH oxidation, FNRL-PC exhibits a 30% decrease in the Michaelis constant Km(NADPH), and a 70% increase in Km(ferredoxin), which is in agreement with its predicted lower activity of ferredoxin reduction. During NADP+ reduction, the FNRL-PC shows a 29/43% decrease in the rate of single electron transfer from reduced ferredoxin in the presence/absence of NADP+. The increase in Km(ferredoxin) and the rate decrease of single reduction are attributed to steric hindrance by the phycocyanin moiety of FNRL-PC. Both isoforms are capable of catalyzing the NADP+ reduction under multiple turnover conditions. Furthermore, we obtained evidence that, under high ionic strength conditions, electron transfer from reduced ferredoxin is rate limiting during this process. The differences that we observe might not fully explain the in vivo properties of the Synechocystis mutants expressing only one of the isoforms. Therefore, we advocate that FNR localization and/or substrates availability are essential in vivo.  相似文献   

9.
The effect of NADP+ on light-induced steady-state redox changes of membrane-bound cytochromes was investigated in membrane fragments prepared from the blue-green algae Nostoc muscorum (Strain 7119) that had high rates of electron transport from water to NADP+ and from an artificial electron donor, reduced dichlorophenolindophenol (DCIPH2) to NADP+. The membrane fragments contained very little phycocyanin and had excellent optical properties for spectrophotometric assays. With DCIPH2 as the electron donor, NADP+ had no effect on the light-induced redox changes of cytochromes: with or without NADP+, 715- or 664-nm illumination resulted mainly in the oxidation of cytochrome f and of other component(s) which may include a c-type cytochrome with an α peak at 549 nm. With 664 nm illumination and water as the electron donor, NADP+ had a pronounced effect on the redox state of cytochromes, causing a shift toward oxidation of a component with a peak at 549 nm (possibly a c-type cytochrome), cytochrome f, and particularly cytochrome b559. Cytochrome b559 appeared to be a component of the main noncyclic electron transport chain and was photooxidized at physiological temperatures by Photosystem II. This photooxidation was apparent only in the presence of a terminal acceptor (NADP+) for the electron flow from water.  相似文献   

10.
SidA (siderophore A) is a flavin-dependent N-hydroxylating monooxygenase that is essential for virulence in Aspergillus fumigatus. SidA catalyzes the NADPH- and oxygen-dependent formation of N5-hydroxyornithine. In this reaction, NADPH reduces the flavin, and the resulting NADP+ is the last product to be released. The presence of NADP+ is essential for activity, as it is required for stabilization of the C4a-hydroperoxyflavin, which is the hydroxylating species. As part of our efforts to determine the molecular details of the role of NADP(H) in catalysis, we targeted Ser-257 for site-directed mutagenesis and performed extensive characterization of the S257A enzyme. Using a combination of steady-state and stopped-flow kinetic experiments, substrate analogs, and primary kinetic isotope effects, we show that the interaction between Ser-257 and NADP(H) is essential for stabilization of the C4a-hydroperoxyflavin. Molecular dynamics simulation results suggest that Ser-257 functions as a pivot point, allowing the nicotinamide of NADP+ to slide into position for stabilization of the C4a-hydroperoxyflavin.  相似文献   

11.
《BBA》2019,1860(10):148058
Brucella ovis encodes a bacterial subclass 1 ferredoxin-NADP(H) reductase (BoFPR) that, by similarity with other FPRs, is expected either to deliver electrons from NADPH to the redox-based metabolism and/or to oxidize NADPH to regulate the soxRS regulon that protects bacteria against oxidative damage. Such potential roles for the pathogen survival under infection conditions make of interest to understand and to act on the BoFPR mechanism. Here, we investigate the NADP+/H interaction and NADPH oxidation by hydride transfer (HT) to BoFPR. Crystal structures of BoFPR in free and in complex with NADP+ hardly differ. The latter shows binding of the NADP+ adenosine moiety, while its redox-reactive nicotinamide protrudes towards the solvent. Nonetheless, pre-steady-state kinetics show formation of a charge-transfer complex (CTC-1) prior to the hydride transfer, as well as conversion of CTC-1 into a second charge-transfer complex (CTC-2) concomitantly with the HT event. Thus, during catalysis nicotinamide and flavin reacting rings stack. Kinetic data also identify the HT itself as the rate limiting step in the reduction of BoFPR by NADPH, as well as product release limiting the overall reaction. Using all-atom molecular dynamics simulations with a thermal effect approach we are able to visualise a potential transient catalytically competent interaction of the reacting rings. Simulations indicate that the architecture of the FAD folded conformation in BoFPR might be key in catalysis, pointing to its adenine as an element to orient the reactive atoms in conformations competent for HT.  相似文献   

12.
Gupta VK  Singh R 《Plant physiology》1988,87(3):741-744
NADP+-isocitrate dehydrogenase (threo-DS-isocitrate: NADP+ oxidoreductase [decarboxylating]; EC 1.1.1.42) (IDH) from pod walls of chickpea (Cicer arietinum L.) was purified 192-fold using ammonium sulfate fractionation, ion exchange chromatography on DEAE-Sephadex A-50, and gel filtration through Sephadex G-200. The purified enzyme, having a molecular weight of about 126,000, exhibited a broad pH optima from 8.0 to 8.6. It was quite stable at 4°C and had an absolute requirement for a divalent cation, either Mg2+ or Mn2+, for its activity. Typical hyperbolic kinetics was obtained with increasing concentrations of NADP+, dl-isocitrate, Mn2+, and Mg2+. Their Km values were 15, 110, 15, and 192 micromolar, respectively. The enzyme activity was inhibited by sulfhydryl reagents. Various amino acids, amides, organic acids, nucleotides, each at a concentration of 5 millimolar, had no effect on the activity of the enzyme. The activity was not influenced by adenylate energy charge but decreased linearly with increasing ratio of NADPH to NADP+. Initial velocity studies indicated kinetic mechanism to be sequential. NADPH inhibited the forward reaction competitively with respect to NADP+ at fixed saturating concentration of isocitrate, whereas 2-oxoglutarate inhibited the enzyme noncompetitively at saturating concentrations of both NADP+ and isocitrate, indicating the reaction mechanism to be random sequential. Results suggest that the activity of NADP+-IDH in situ is likely to be controlled by intracellular NADPH to NADP+ ratio as well as by the concentration of various substrates and products.  相似文献   

13.
Product inhibition studies with Rhodopseudomonas spheriodes NADP+ specific isocitrate dehydrogenase indicate that the enzyme mechanism involves the ordered addition of the substrates NADP+ and threo-ds-isocitrate and the ordered release of products CO2 (HCOs?), 2-ketoglutarate, and NADPH. In addition, the presence of a ternary complex consisting of enzyme, NADP+, and 2-ketoglutarate is indicated. Binding studies with radioactive substrates support the kinetically derived mechanism. The Rhodopseudomonas enzyme is dimeric and contains but a single active site. Different combinations of substrate were ineffective in causing gross changes in molecular structure as monitored by gel filtration techniques. A comparison of the amino acid composition of this enzyme with the bacterial enzyme from Azotobacter vinelandii indicate very significant differences in the amino acid compositions.  相似文献   

14.
Sugar beets (Beta vulgaris L. cv. F58-554H1) were cultured hydroponically in growth chambers at 25°C, with a photon flux density of 500 mol m-2s-1. Measurements were made of net CO2 exchange, leaf adenylates (ATP, ADP and AMP), and leaf nicotinamide nucleotides (NAD+, NADP+, NADH, NADPH), over the diurnal period (16h light/8 h dark) and during photosynthetic induction. All the measurements were carried out on recently expanded leaves from 5-week-old plants. When the lights were switched on in the growth chamber, the rate of photosynthetic CO2 uptake, and the levels of leaf ATP and NADPH increased to a maximum in 30 min and remained there throughout the light period. The increase in ATP over the first few minutes of illumination was associated with the phosphorylation of ADP to ATP and the increase in NADPH with the reduction of NADP+; subsequently, the increase in ATP was associated with an increase in total adenylates while the increase in NADPH was associated with an accumulation of NADP+ and NADPH due to the light-driven phosphorylation of NAD+ to NADP+. On return to darkness, ATP and NADPH values decreased much more slowly, requiring 2 to 4 hours to reach minimum values. From these results we suggest that (i) the total adenylate and NADPH and NADP+ (but not NAD+ and NADH) pools increase following exposure to light; (ii) the increase in pool size is not accompanied by any large change in the energy or redox states of the system; and (iii) the measured ratios of ATP/ADP and NADPH/NADP+ for intact leaves are low and constant during steady-state illumination.Abbreviations AEC adenylate energy charge - DHAP dihydroxyacetone phosphate - MTT 3-(4,5-dimethylthiazolyl-2)-2,5-diphenyltetrazolium bromide - PES phenazine ethosulfate - PEP phosphoenolpyruvate - PGA 3-phosphoglycerate - PFD photon flux density - Ru5P ribulose-5-phosphate - Rubisco ribulose 1,5-bisphosphate carboxylase/oxygenase  相似文献   

15.
Malic enzyme (EC 1.1.1.40) converts l-malate to pyruvate and CO2 providing NADPH for metabolism especially for lipid biosynthesis in oleaginous microorganisms. However, its role in the oleaginous yeast, Yarrowia lipolytica, is unclear. We have cloned the malic enzyme gene (YALI0E18634g) from Y. lipolytica into pET28a, expressed it in Escherichia coli and purified the recombinant protein (YlME). YlME used NAD+ as the primary cofactor. Km values for NAD+ and NADP+ were 0.63 and 3.9 mM, respectively. Citrate, isocitrate and α-ketoglutaric acid (>5 mM) were inhibitory while succinate (5–15 mM) increased NADP+- but not NAD+-dependent activity. To determine if fatty acid biosynthesis could be increased in Y. lipolytica by providing additional NADPH from an NADP+-dependent malic enzyme, the malic enzyme gene (mce2) from an oleaginous fungus, Mortierella alpina, was expressed in Y. lipolytica. No significant changes occurred in lipid content or fatty acid profiles suggesting that malic enzyme is not the main source of NADPH for lipid accumulation in Y. lipolytica.  相似文献   

16.
Two isoenzymes of glucose 6-phosphate dehydrogenase (EC 1.1.1.49) have been separated from the plant fraction of soybean (Glycine max L. Merr. cv Williams) nodules by a procedure involving (NH4)2SO4 gradient fractionation, gel chromatography, chromatofocusing, and affinity chromatography. The isoenzymes, which have been termed glucose 6-phosphate dehydrogenases I and II, were specific for NADP+ and glucose 6-phosphate and had optimum activity at pH 8.5 and pH 8.1, respectively. Both isoenzymes were labile in the absence of NADP+. The apparent molecular weight of glucose 6-phosphate dehydrogenases I and II at pH 8.3 was estimated by gel chromatography to be approximately 110,000 in the absence of NADP+ and double this size in the presence of NADP+. The apparent molecular weight did not increase when glucose 6-phosphate was added with NADP+ at pH 8.3. Both isoenzymes had very similar kinetic properties, displaying positive cooperativity in their interaction with NADP+ and negative cooperativity with glucose 6-phosphate. The isoenzymes had half-maximal activity at approximately 10 micromolar NADP+ and 70 to 100 micromolar glucose 6-phosphate. NADPH was a potent inhibitor of both of the soybean nodule glucose 6-phosphate dehydrogenases.  相似文献   

17.
These studies indicate that the interconversions of Δ1-pyrroline-5-carboxylate and proline can function as a shuttle that generates extra-mitochondrial NADP+ and transfers hydride ions into mitochondria in a cell-free rat liver system. A phosphate-free buffer with high concentrations of triethanolamine and 2-mercaptoethanol prevented the cold inactivation of pyrroline-5-carboxylate reductase (ED 1.5.1.2) in liver extracts. This enzyme had an apparent KmNADPH that was 2% of the apparent KmNADH. VmaxNADPH was approx. 50% of VmaxNADH. Unlabeled proline was converted to [5-3H]proline in incubations containing liver soluble fraction, mitochondria and a [4S-3H]NADPH generating system. This demonstrated one turn of the proposed shuttle in a homologous liver system. [5-3H]Proline production increased linearly over 60 min and decreased by 87% or more when specific components were eliminated. Rotenone was required for maximal activity, suggesting that inhibition of Δ1-pyrroline-5-carboxylate efflux would be required for significant shuttle activity in vivo. Both the relative concentrations of NADPH and NADH in liver cytosol and the kinetic characteristics of liver pyrroline-5-carboxylate reductase predict that the described shuttle should be overwhelmingly linked to NADPH rather than NADH. A NADPH-linked Δ1-pyrroline-5-carboxylate-proline shuttle may occur in hepatocytes and function at specific times to regulate pathways limited by cytosolic [NADP+].  相似文献   

18.
The stoichiometry of hydroxylation reactions catalyzed by cytochrome P-450 was studied in a reconstituted enzyme system containing the highly purified cytochrome from phenobarbital-induced rabbit liver microsomes. Hydrogen peroxide was shown to be formed in the reconstituted system in the presence of NADPH and oxygen; the amount of peroxide produced varied with the substrated added. NADPH oxidation, oxygen consumption, and total product formation (sum of hydroxylated compound and hydrogen peroxide) were shown to be equimolar when cyclohexane, benzphetamine, or dimethylaniline served as the substrate. The stoichiometry observed represents the sum of two activities associated with cytochrome P-450. These are (1) hydroxylase activity: NADPH + H+ + O2 + RH → NADP+ + H2O + ROH; and (2) oxidase activity: NADPH + H+ + O2 → NADP+ + H2O2. Benzylamphetamine (desmethylbenzphetamine) acts as a pseudosubstrate in that it stimulates peroxide formation to the same extent as the parent compound (benzphetamine), but does not undergo hydroxylation. Accordingly, when benzylamphetamine alone is added in control experiments to correct for the NADPH and O2 consumption not associated with benzphetamine hydroxylation, the expected 1:1:1 stoichiometry for NADPH oxidation, O2 consumption, and formaldehyde formation in the hydroxylation reaction is observed.  相似文献   

19.
The kinetic mechanism of the reaction catalyzed by glucose-6-phosphate dehydrogenase (EC 1.1.1.49) from Dicentrarchus labrax liver was examined using initial velocity studies,NADPH and glucosamine 6-phosphate inhibition and alternate coenzyme experiments. The results are consistent with a steady-state ordered sequential mechanism in which NADP+ binds first to the enzyme and NADPH is released last. Replots of NADPH inhibition show an uncommon parabolic pattern for this enzyme that has not been previously described. A kinetic model is proposed in agreement with our kinetic results and with previously published structural studies (Bautista et al. (1988) Biochem. Soc. Trans. 16, 903–904). The kinetic mechanism presented provides a possible explanation for the regulation of the enzyme by the [NADPH]/[NADP+] ratio.  相似文献   

20.
Two pathways serve for assimilation of ammonia inParacoccus denitrificans. Glutamate dehydrogenase (NADP+) catalyzes the assimilation at a high NH4 + concentration. If nitrate serves as the nitrogen source, glutamate is synthesized by glutamate-ammonia ligase and glutamate synthase (NADPH). At a very low NH4 + concentration, all three enzymes are synthesized simultaneously. No direct relationship exists between glutamate dehydrogenase (NADP+) and glutamate-ammonia ligase inP. denitrificans, while the glutamate synthase (NADPH) activity changes in parallel with that of the latter enzyme. Ammonia does not influence the induction or repression of glutamate dehydrogenase (NADP+). The inner concentration of metabolites indicates a possible repression of glutamate dehydrogenase (NADP+) by the high concentration of glutamine or its metabolic products as in the case when NH4 + is formed by assimilative nitrate reduction. No direct effect of the intermediates of nitrate assimilation on the synthesis of glutamate dehydrogenase (NADP+) was observed.  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