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 671 毫秒
1.
José Ramón Peregrina 《BBA》2010,1797(9):1638-1264
Two transient charge-transfer complexes (CTC) form prior and upon hydride transfer (HT) in the reversible reaction of the FAD-dependent ferredoxin-NADP+ reductase (FNR) with NADP+/H, FNRox-NADPH (CTC-1), and FNRrd-NADP+ (CTC-2). Spectral properties of both CTCs, as well as the corresponding interconversion HT rates, are here reported for several Anabaena FNR site-directed mutants. The need for an adequate initial interaction between the 2′P-AMP portion of NADP+/H and FNR that provides subsequent conformational changes leading to CTC formation is further confirmed. Stronger interactions between the isoalloxazine and nicotinamide rings might relate with faster HT processes, but exceptions are found upon distortion of the active centre. Thus, within the analyzed FNR variants, there is no strict correlation between the stability of the transient CTCs formation and the rate of the subsequent HT. Kinetic isotope effects suggest that, while in the WT, vibrational enhanced modulation of the active site contributes to the tunnel probability of HT; complexes of some of the active site mutants with the coenzyme hardly allow the relative movement of isoalloxazine and nicotinamide rings along the HT reaction. The architecture of the WT FNR active site precisely contributes to reduce the stacking probability between the isoalloxazine and nicotinamide rings in the catalytically competent complex, modulating the angle and distance between the N5 of the FAD isoalloxazine and the C4 of the coenzyme nicotinamide to values that ensure efficient HT processes.  相似文献   

2.
The role of the highly conserved C266 and L268 of pea ferredoxin–NADP+ reductase (FNR) in formation of the catalytically competent complex of the enzyme with NADP(H) was investigated. Previous studies suggest that the volume of these side-chains, situated facing the side of the C-terminal Y308 catalytic residue not stacking the flavin isoalloxazine ring, may be directly involved in the fine-tuning of the catalytic efficiency of the enzyme. Wild-type pea FNR as well as single and double mutants of C266 and L268 residues were analysed by fast transient-kinetic techniques and their midpoint reduction potentials were determined. For the C266A, C266M and C266A/L268A mutants a significant reduction in the overall hydride transfer (HT) rates was observed along with the absence of charge-transfer complex formation. The HT rate constants for NADPH oxidation were lower than those for NADP+ reduction, reaching a 30-fold decrease in the double mutant. In agreement, these variants exhibited more negative midpoint potentials with respect to the wild-type enzyme. The three-dimensional structures of C266M and L268V variants were solved. The C266M mutant shows a displacement of E306 away from the relevant residue S90 to accommodate the bulky methionine introduced. The overall findings indicate that in FNR the volume of the residue at position 266 is essential to attain the catalytic architecture between the nicotinamide and isoalloxazine rings at the active site and, therefore, for an efficient HT process. In addition, flexibility of the 268–270 loop appears to be critical for FNR to achieve catalytically competent complexes with NADP(H).  相似文献   

3.
《BBA》2014,1837(2):251-263
Ferredoxin-nicotinamide–adenine dinucleotide phosphate (NADP+) reductase (FNR) catalyses the production of reduced nicotinamide–adenine dinucleotide phosphate (NADPH) in photosynthetic organisms, where its flavin adenine dinucleotide (FAD) cofactor takes two electrons from two reduced ferredoxin (Fd) molecules in two sequential steps, and transfers them to NADP+ in a single hydride transfer (HT) step. Despite the good knowledge of this catalytic machinery, additional roles can still be envisaged for already reported key residues, and new features are added to residues not previously identified as having a particular role in the mechanism. Here, we analyse for the first time the role of Ser59 in Anabaena FNR, a residue suggested by recent theoretical simulations as putatively involved in competent binding of the coenzyme in the active site by cooperating with Ser80. We show that Ser59 indirectly modulates the geometry of the active site, the interaction with substrates and the electronic properties of the isoalloxazine ring, and in consequence the electron transfer (ET) and HT processes. Additionally, we revise the role of Tyr79 and Ser80, previously investigated in homologous enzymes from plants. Our results probe that the active site of FNR is tuned by a H-bond network that involves the side-chains of these residues and that results to critical optimal substrate binding, exchange of electrons and, particularly, competent disposition of the C4n (hydride acceptor/donor) of the nicotinamide moiety of the coenzyme during the reversible HT event.  相似文献   

4.
Ferredoxin-NADP+ reductase (FNR) catalyses the electron transfer from ferredoxin to NADP+ via its flavin FAD cofactor. A molecular dynamics theoretical approach is applied here to visualise the transient catalytically competent interaction of Anabaena FNR with its coenzyme, NADP+. The particular role of some of the residues identified as key in binding and accommodating the 2′P-AMP moiety of the coenzyme is confirmed in molecular terms. Simulations also indicate that the architecture of the active site precisely contributes to the orientation of the N5 of the FAD isoalloxazine ring and the C4 of the coenzyme nicotinamide ring in the conformation of the catalytically competent hydride transfer complex and, therefore, contributes to the efficiency of the process. In particular, the side chain of the C-terminal Y303 in Anabaena FNR appears key to providing the optimum geometry by reducing the stacking probability between the isoalloxazine and nicotinamide rings, thus providing the required co-linearity and distance among the N5 of the flavin cofactor, the C4 of the coenzyme nicotinamide and the hydride that has to be transferred between them. All these factors are highly related to the reaction efficiency, mechanism and reversibility of the process.  相似文献   

5.
《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   

6.
Biosynthesis of UDP-N-acetylmuramic acid in bacteria is a committed step towards peptidoglycan production. In an NADPH- and FAD-dependent reaction, the UDP-N-acetylglucosamine-enolpyruvate reductase (MurB) reduces UDP-N-acetylglucosamine-enolpyruvate to UDP-N-acetylmuramic acid. We determined the three-dimensional structures of the ternary complex of Pseudomonas aeruginosa MurB with FAD and NADP+ in two crystal forms to resolutions of 2.2 and 2.1 Å, respectively, to investigate the structural basis of the first half-reaction, hydride transfer from NADPH to FAD. The nicotinamide ring of NADP+ stacks against the si face of the isoalloxazine ring of FAD, suggesting an unusual mode of hydride transfer to flavin. Comparison with the structure of the Escherichia coli MurB complex with UDP-N-acetylglucosamine-enolpyruvate shows that both substrates share the binding site located between two lobes of the substrate-binding domain III, consistent with a ping pong mechanism with sequential substrate binding. The nicotinamide and the enolpyruvyl moieties are strikingly well-aligned upon superimposition, both positioned for hydride transfer to and from FAD. However, flexibility of the substrate channel allows the non-reactive parts of the two substrates to bind in different conformations. A potassium ion in the active site may assist in substrate orientation and binding. These structural models should help in structure-aided drug design against MurB, which is essential for cell wall biogenesis and hence bacterial survival.  相似文献   

7.
Experiments designed to elucidate the nature of 17β-hydroxysteroid dehydrogenase from human red blood cells have shown that NADP+ activates and protects the enzyme, while also serving as substrate for the reaction. Enzyme activity was measured by the conversion of 17β-estradiol to estrone and by the production of NADPH with 17β-estradiol-3-sulfate as substrate. It appears that the reaction sequence is first, binding with NADP+ and second, binding with the steroid. The binding with NADP+ is essentially irreversible: the activated enzyme is completely protected against loss of activity by dilution. On dilution of the unactivated enzyme, much of the activity is lost. The bireactant rate equation of the sequential type has been restated for the case of activation by one of the reactants. Since it has been found that activation of enzyme is linear with NADP+ concentration, it follows that the Michaelis constant for the steroid substrate is independent of the concentration of NADP+ activating the enzyme. This is substantiated by the determination of the Michaelis constant for 17β-estradiol-3-sulfate from data on double-reciprocal plots of activated and unactivated enzyme with limiting amounts of steroid. The activating effect increases linearly up to a concentration of 1.2 × 10?5m of NADP+ and then levels off. The activation is highly specific for NADP+; neither NAD+, ATP, NADPH, nicotinic acid, ncr nicotinamide prevent the loss of activity after storing the enzyme for 1 hr at 37 °C. The steroid substrate appears to interfere with the activation of NADP+.  相似文献   

8.
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   

9.
We measured both pyridine nucleotide levels and ribonucleotide reductase-specific activity in Yoshida ascites hepatoma cells as a function of growth in vivo and during recruitment from non-cycling to cycling state in vitro. Oxidized nicotinamide adenine dinucleotide (NAD+) and reduced nicotinamide adenine dinucleotide (NADP) levels remained unchanged during tumour growth, while NADP+ and reduced nicotinamide adenine dinucleotide phosphate (NADPH) levels were very high in exponentially growing cells and markedly decreased in the resting phase. Ribonucleotide reductase activity paralleled NADP(H) (NADP+ plus NADPH) intracellular content. The concomitant increase in both NADP(H) levels and ribonucleotide reductase activity was also observed during G1-S transition in vitro. Cells treated with hydroxyurea showed a comparable correlation between the pool size of NADP(H) and ribonucleotide reductase activity. On the basis of these findings, we suggest that fluctuations in NADP(H) levels and ribonucleotide reductase activity might play a critical role in cell cycle regulation.  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.
The role of the highly conserved C266 and L268 of pea ferredoxin-NADP(+) reductase (FNR) in formation of the catalytically competent complex of the enzyme with NADP(H) was investigated. Previous studies suggest that the volume of these side-chains, situated facing the side of the C-terminal Y308 catalytic residue not stacking the flavin isoalloxazine ring, may be directly involved in the fine-tuning of the catalytic efficiency of the enzyme. Wild-type pea FNR as well as single and double mutants of C266 and L268 residues were analysed by fast transient-kinetic techniques and their midpoint reduction potentials were determined. For the C266A, C266M and C266A/L268A mutants a significant reduction in the overall hydride transfer (HT) rates was observed along with the absence of charge-transfer complex formation. The HT rate constants for NADPH oxidation were lower than those for NADP(+) reduction, reaching a 30-fold decrease in the double mutant. In agreement, these variants exhibited more negative midpoint potentials with respect to the wild-type enzyme. The three-dimensional structures of C266M and L268V variants were solved. The C266M mutant shows a displacement of E306 away from the relevant residue S90 to accommodate the bulky methionine introduced. The overall findings indicate that in FNR the volume of the residue at position 266 is essential to attain the catalytic architecture between the nicotinamide and isoalloxazine rings at the active site and, therefore, for an efficient HT process. In addition, flexibility of the 268-270 loop appears to be critical for FNR to achieve catalytically competent complexes with NADP(H).  相似文献   

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

13.
F420H2:NADP+ Oxidoreductase (Fno) catalyzes the reversible reduction of NADP+ to NADPH by transferring a hydride from the reduced F420 cofactor. Here, we have employed binding studies, steady-state and pre steady-state kinetic methods upon wtFno and isoleucine 135 (I135) Fno variants in order to study the effects of side chain length on the donor-acceptor distance between NADP+ and the F420 precursor, FO. The conserved I135 residue of Fno was converted to a valine, alanine and glycine, thereby shortening the side chain length. The steady-state kinetic analysis of wtFno and the variants showed classic Michaelis-Menten kinetics with varying FO concentrations. The data revealed a decreased kcat as side chain length decreased, with varying FO concentrations. The steady-state plots revealed non-Michaelis-Menten kinetic behavior when NADPH was varied. The double reciprocal plot of the varying NADPH concentrations displays a downward concave shape, while the NADPH binding curves gave Hill coefficients of less than 1. These data suggest that negative cooperativity occurs between the two identical monomers. The pre steady-state Abs420 versus time trace revealed biphasic kinetics, with a fast phase (hydride transfer) and a slow phase. The fast phase displayed an increased rate constant as side chain length decreased. The rate constant for the second phase, remained ~2 s?1 for each variant. Our data suggest that I135 plays a key role in sustaining the donor-acceptor distance between the two cofactors, thereby regulating the rate at which the hydride is transferred from FOH2 to NADP+. Therefore, Fno is a dynamic enzyme that regulates NADPH production.  相似文献   

14.
A study on product inhibition of 17ß-hydroxysteroid dehydrogenase from porcine testes was carried out by measuring the initial velocities of NADPH formation using testosterone as the substrate steroid. Type of inhibition by NADPH against NADP+ was competitive in both saturated and unsaturated concentrations of testosterone. In the saturated concentration of NADP+, activity of the enzyme was not inhibited by NADPH against testosterone. In the unsaturated concentrations of NADP+, however, NADPH brought mixed type inhibition against testosterone. The similar modes of inhibition by the product steroid, androstenedione were observed in the saturated and unsaturated concentrations of NADP+ and testosterone.The fluorescence of NADPH was increased in the presence of the enzyme, and fluorometric titration indicated that 1 mol of NADPH was bound to 1 mol of the 17ß-hydroxysteroid dehydrogenase. Addition of testosterone to enzyme-NADPH complex reduced the intensity of fluorescence of NADPH, suggesting formation of testosterone-enzyme-NADPH complex as a ternary dead end complex.From the analyses of product inhibition and spectral changes of NADPH, the kinetic mechanism of the enzyme was revealed as rapid equilibrium random system with two dead end complexes which consisted of the two reduced reactants bound to the enzyme and the two oxidized ones bound to it.  相似文献   

15.
Bacillus subtilis yumC encodes a novel type of ferredoxin‐NADP+ oxidoreductase (FNR) with a primary sequence and oligomeric conformation distinct from those of previously known FNRs. In this study, the crystal structure of B. subtilis FNR (BsFNR) complexed with NADP+ has been determined. BsFNR features two distinct binding domains for FAD and NADPH in accordance with its structural similarity to Escherichia coli NADPH‐thioredoxin reductase (TdR) and TdR‐like protein from Thermus thermophilus HB8 (PDB code: 2ZBW). The deduced mode of NADP+ binding to the BsFNR molecule is nonproductive in that the nicotinamide and isoalloxazine rings are over 15 Å apart. A unique C‐terminal extension, not found in E. coli TdR but in TdR‐like protein from T. thermophilus HB8, covers the re‐face of the isoalloxazine moiety of FAD. In particular, Tyr50 in the FAD‐binding region and His324 in the C‐terminal extension stack on the si‐ and re‐faces of the isoalloxazine ring of FAD, respectively. Aromatic residues corresponding to Tyr50 and His324 are also found in the plastid‐type FNR superfamily of enzymes, and the residue corresponding to His324 has been reported to be responsible for nucleotide specificity. In contrast to the plastid‐type FNRs, replacement of His324 with Phe or Ser had little effect on the specificity or reactivity of BsFNR with NAD(P)H, whereas replacement of Arg190, which interacts with the 2′‐phosphate of NADP+, drastically decreased its affinity toward NADPH. This implies that BsFNR adopts the same nucleotide binding mode as the TdR enzyme family and that aromatic residue on the re‐face of FAD is hardly relevant to the nucleotide selectivity.  相似文献   

16.
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   

17.
Type I Baeyer–Villiger monooxygenases (BVMOs) strongly prefer NADPH over NADH as an electron donor. In order to elucidate the molecular basis for this coenzyme specificity, we have performed a site-directed mutagenesis study on phenylacetone monooxygenase (PAMO) from Thermobifida fusca. Using sequence alignments of type I BVMOs and crystal structures of PAMO and cyclohexanone monooxygenase in complex with NADP+, we identified four residues that could interact with the 2′-phosphate moiety of NADPH in PAMO. The mutagenesis study revealed that the conserved R217 is essential for binding the adenine moiety of the nicotinamide coenzyme while it also contributes to the recognition of the 2′-phosphate moiety of NADPH. The substitution of T218 did not have a strong effect on the coenzyme specificity. The H220N and H220Q mutants exhibited a ~3-fold improvement in the catalytic efficiency with NADH while the catalytic efficiency with NADPH was hardly affected. Mutating K336 did not increase the activity of PAMO with NADH, but it had a significant and beneficial effect on the enantioselectivity of Baeyer–Villiger oxidations and sulfoxidations. In conclusion, our results indicate that the function of NADPH in catalysis cannot be easily replaced by NADH. This finding is in line with the complex catalytic mechanism and the vital role of the coenzyme in BVMOs.  相似文献   

18.
To study the role of the mobile C-terminal extension present in bacterial class of plant type NADP(H):ferredoxin reductases during catalysis, we generated a series of mutants of the Rhodobacter capsulatus enzyme (RcFPR). Deletion of the six C-terminal amino acids beyond alanine 266 was combined with the replacement A266Y, emulating the structure present in plastidic versions of this flavoenzyme. Analysis of absorbance and fluorescence spectra suggests that deletion does not modify the general geometry of FAD itself, but increases exposure of the flavin to the solvent, prevents a productive geometry of FAD:NADP(H) complex and decreases the protein thermal stability. Although the replacement A266Y partially coats the isoalloxazine from solvent and slightly restores protein stability, this single change does not allow formation of active charge-transfer complexes commonly present in the wild-type FPR, probably due to restraints of C-terminus pliability. A proton exchange process is deduced from ITC measurements during coenzyme binding. All studied RcFPR variants display higher affinity for NADP+ than wild-type, evidencing the contribution of the C-terminus in tempering a non-productive strong (rigid) interaction with the coenzyme. The decreased catalytic rate parameters confirm that the hydride transfer from NADPH to the flavin ring is considerably hampered in the mutants. Although the involvement of the C-terminal extension from bacterial FPRs in stabilizing overall folding and bent-FAD geometry has been stated, the most relevant contributions to catalysis are modulation of coenzyme entrance and affinity, promotion of the optimal geometry of an active complex and supply of a proton acceptor acting during coenzyme binding.  相似文献   

19.
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   

20.
Butanedione in the presence of borate buffer reversibly inhibits Rhodospirillum rubrum chromatophore transhydrogenase complex and the separated membrane-bound and soluble factor components of the complex. NADP+ completely protected against inactivation of the membrane-bound component, whereas NAD+ was without effect. Soluble factor was maximally protected only partially by either NAD+ or NADP+, but a mixture of the substrates afforded complete protection. NADP+-dependent association of soluble factor with factor-depleted membranes was markedly decreased after incubation of membranes with butanedione in the absence, but not in the presence, of NADP+. Soluble factor was bound to agarose-NAD and was eluted by NAD+, but not by NADP+. These results demonstrate the presence of at least three nicotinamide adenine dinucleotide binding sites on R. rubrum transhydrogenase complex, including separate NADP and NAD binding sites on soluble factor and a NADP binding site on the membrane-bound component.  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