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 507 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.
NADH oxidase activity of mitochondrial apoptosis-inducing factor   总被引:25,自引:0,他引:25  
Apoptosis-inducing factor (AIF) is a mitochondrial flavoprotein, which translocates to the nucleus during apoptosis and causes chromatin condensation and large scale DNA fragmentation. Here we report the biochemical characterization of AIF's redox activity. Natural AIF purified from mitochondria and recombinant AIF purified from bacteria (AIFDelta1-120) exhibit NADH oxidase activity, whereas superoxide anion (O(2)(-)) is formed. AIFDelta1-120 is a monomer of 57 kDa containing 1 mol of noncovalently bound FAD/mol of protein. ApoAIFDelta1-120, which lacks FAD, has no NADH oxidase activity. However, native AIFDelta1-120, apoAIFDelta1-120, and the reconstituted (FAD-containing) holoAIFDelta1-120 protein exhibit a similar apoptosis-inducing potential when microinjected into the cytoplasm of intact cells. Inhibition of the redox function, by external addition of superoxide dismutase or covalent derivatization of FAD with diphenyleneiodonium, failed to affect the apoptogenic function of AIFDelta1-120 assessed on purified nuclei in a cell-free system. Conversely, blockade of the apoptogenic function of AIFDelta1-120 with the thiol reagent para- chloromercuriphenylsulfonic acid did not affect its NADH oxidase activity. Altogether, these data indicate that AIF has a marked oxidoreductase activity which can be dissociated from its apoptosis-inducing function.  相似文献   

3.
Mitochondrial apoptosis-inducing factor (AIF) is a central player in the caspase-independent cell death pathway whose normal physiological function remains unclear. Our study showed that naturally folded mouse AIF very slowly reacts with NAD(P)H (k cat of 0.2-0.01 s(-1)) forming tight, dimeric, and air-stable FADH2-NAD(P) charge-transfer complexes ineffective in electron transfer. FAD reduction is accompanied by a conformational change involving AIF-specific N-terminal and regulatory 509-559 peptides and the active site His 453, and it affects susceptibility of AIF to calpain and AIF-DNA interaction, the two events critical for initiating caspase-independent apoptosis. Based on our results, we propose that formation of long lived complexes with NAD(P)H and redox reorganization may be functionally important and enable AIF to act as a redox-signaling molecule linking NAD(P)H-dependent metabolic pathways to apoptosis.  相似文献   

4.
Reduction of Complex I (NADH:ubiquinone oxidoreductase I) from Escherichia coli by NADH was investigated optically by means of an ultrafast stopped-flow approach. A locally designed microfluidic stopped-flow apparatus with a low volume (0.2 μl) but a long optical path (10 mm) cuvette allowed measurements in the time range from 270 μs to seconds. The data acquisition system collected spectra in the visible range every 50 μs. Analysis of the obtained time-resolved spectral changes upon the reaction of Complex I with NADH revealed three kinetic components with characteristic times of < 270 μs, 0.45–0.9 ms and 3–6 ms, reflecting reduction of different FeS clusters and FMN. The rate of the major (τ = 0.45–0.9 ms) component was slower than predicted by electron transfer theory for the reduction of all FeS clusters in the intraprotein redox chain. This delay of the reaction was explained by retention of NAD+ in the catalytic site. The fast optical changes in the time range of 0.27–1.5 ms were not altered significantly in the presence of 10-fold excess of NAD+ over NADH. The data obtained on the NuoF E95Q variant of Complex I shows that the single amino acid replacement in the catalytic site caused a strong decrease of NADH binding and/or the hydride transfer from bound NADH to FMN.  相似文献   

5.
A comparative study on the solution structures of bovine microsomal cytochrome b5 (Tb5) and the mutant V45H has been achieved by 1D and 2D 1H-NMR spectroscopy to clarify the differences in the solution conformations between these two proteins. The results reveal that the global folding of the V45H mutant in solution is unchanged, but the subtle changes exist in the orientation of the axial ligand His39, and heme vinyl groups. The side chain of His45 in V45H mutant extends to the outer edge of the heme pocket leaving a cavity at the site originally occupied by the inner methyl group of Val45 residue. In addition, the imidazole ring of axial ligand His39 rotates counterclockwise by approximately 3 degrees around the His-Fe-His axis, and the 4-heme vinyl group turns to the space vacated by the removed side chain due to the mutation. Furthermore, the helix III of the heme pocket undergoes outward displacement, while the linkage between helix II and III is shifted leftward. These observations are not only consistent with the pattern of the pseudocontact shifts of the heme protons, but also well account for the lower stability of V45H mutant against heat and urea.  相似文献   

6.
Mitochondrial apoptosis-inducing factor (AIF) is a FAD-containing protein that under certain conditions translocates to the nucleus and causes a programmed cell death, apoptosis. The apoptogenic action of AIF is redox controlled as the NADH-reduced AIF dimer has lower affinity for DNA than the oxidized monomer. To gain further insights into the mechanism of AIF, we investigated its interaction with a series of quinone oxidants, including a number of anticancer quinones. Our data indicate that the NADH:quinone oxidoreduction catalyzed by AIF follows a “ping-pong” scheme, with the reductive half-reaction being rate-limiting and the FADH–NAD+ charge-transfer complex serving as an electron donor. AIF is equally reactive toward benzo- and naphthoquinones, but may discriminate structures with a higher number of aromatic rings. The reactivity of quinones is mainly defined by their one-electron reduction potential, whereas the size and nature of the substituents play a minor role. AIF is unlikely to significantly contribute to bioreductive activation of low-potential quinoidal anticancer quinones. However, high-potential quinones, e.g. a toxic natural compound naphthazarin, maintain AIF in the oxidized state when a significant excess of NADH is present. Thus, these compounds may prevent the accumulation of the reduced form of AIF in vivo, and enhance AIF-mediated apoptosis.  相似文献   

7.
Melanocarpus albomyces laccase crystals were soaked with 2,6-dimethoxyphenol, a common laccase substrate. Three complex structures from different soaking times were solved. Crystal structures revealed the binding of the original substrate and adducts formed by enzymatic oxidation of the substrate. The dimeric oxidation products were identified by mass spectrometry. In the crystals, a 2,6-dimethoxy-p-benzoquinone and a C-O dimer were observed, whereas a C-C dimer was the main product identified by mass spectrometry. Crystal structures demonstrated that the substrate and/or its oxidation products were bound in the pocket formed by residues Ala191, Pro192, Glu235, Leu363, Phe371, Trp373, Phe427, Leu429, Trp507 and His508. Substrate and adducts were hydrogen-bonded to His508, one of the ligands of type 1 copper. Therefore, this surface-exposed histidine most likely has a role in electron transfer by laccases. Based on our mutagenesis studies, the carboxylic acid residue Glu235 at the bottom of the binding site pocket is also crucial in the oxidation of phenolics. Glu235 may be responsible for the abstraction of a proton from the OH group of the substrate and His508 may extract an electron. In addition, crystal structures revealed a secondary binding site formed through weak dimerization in M. albomyces laccase molecules. This binding site most likely exists only in crystals, when the Phe427 residues are packed against each other.  相似文献   

8.
The NADH-dependent polysulfide reductase (Npsr) from Shewanella loihica PV-4 is a member of the single cysteine-containing subset of the family of disulfide reductases represented by glutathione reductase. We have determined the kinetics of the reductive half-reaction of the enzyme with NADH using stopped-flow spectroscopy and kinetic isotope effects, and these results indicate that the reductive and oxidative half-reactions are both partially rate-limiting for enzyme turnover. During reaction with NADH, the reduced nucleotide appears to bind rapidly in an unproductive conformation, followed by the formation of a productive E·NADH complex and subsequent electron transfer to FAD. F161 of Npsr fills the space in which the nicotinamide ring of NADH would be expected to bind. We have shown that while this residue is not absolutely required for catalysis, it does assist in the forward commitment to catalysis through its role in the reductive half reaction, where it appears to enhance hydride transfer in the productive E·NADH complex. While the fluorescence and absorbance spectra of the stable redox forms of the wild-type and F161A mutant enzymes are similar, intermediates formed during reduction and turnover have different characteristics and appear to indicate that the enzyme–NADH complex formed just prior to hydride transfer on the F161A enzyme has weaker FAD–NADH interactions than the wild-type enzyme, consistent with a “looser” enzyme–NADH complex. The 2.7 Å crystal structure of the F161A mutant was determined, and shows that the nicotinamide ring of NADH would have the expected freedom of motion in the more open NADH binding cavity.  相似文献   

9.
10.
The enzyme N-acetyl-gamma-glutamyl-phosphate reductase (AGPR) catalyzes the nicotinamide adenine dinucleotide phosphate (NADPH)-dependent reductive dephosphorylation of N-acetyl-gamma-glutamyl-phosphate to N-acetylglutamate-gamma-semialdehyde. This reaction is part of the arginine biosynthetic pathway that is essential for some microorganisms and plants, in particular, for Mycobacterium tuberculosis (Mtb). The structures of apo MtbAGPR in the space groups P2(1)2(1)2(1) and C2 and the structure of MtbAGPR bound to the cofactor NADP(+) have been solved and analyzed. Each MtbAGPR subunit consists of alpha/beta and alpha+beta domains; NADP(+) is bound in the cleft between them. The hydrogen bonds and hydrophobic contacts between the enzyme and cofactor have been examined. Comparison of the apo and the bound enzyme structures has revealed a conformational change in MtbAGPR upon NADP(+) binding. Namely, a loop (Leu88 to His92) moves more than 5 A to confine sterically the cofactor's adenine moiety in a hydrophobic pocket. To identify the catalytically important residues in MtbAGPR, a docking of the substrate to the enzyme has been performed using the present structure of the MtbAGPR/NADP(+) complex. It reveals that residues His217 and His219 could form hydrogen bonds with the docked substrate. In addition, an ion pair could form between the substrate phosphate group and the guanidinium group of Arg114. These interactions optimally place and orient the substrate for subsequent nucleophilic attack by Cys158 on the substrate gamma-carboxyl group. His219 is the most probable general base to accept a proton from Cys158 and an adjacent ion pair interaction with the side-chain carboxyl group of Glu222 could help to stabilize the resulting positive charge on His219. For this catalytic triad to function efficiently it requires a small conformational change of the order of 1 A in the loop containing His217 and His219; this could easily result from the substrate binding.  相似文献   

11.
NAD kinase catalyzes the phosphorylation of NAD+ to synthesize NADP+, whereas NADH kinase catalyzes conversion of NADH to NADPH. The mitochondrial protein Pos5 of Saccharomyces cerevisiae shows much higher NADH kinase than NAD kinase activity and is therefore referred to as NADH kinase. To clarify the structural determinant underlying the high NADH kinase activity of Pos5 and its selectivity for NADH over NAD+, we determined the tertiary structure of Pos5 complexed with NADH at a resolution of 2.0 Å. Detailed analysis, including a comparison of the tertiary structure of Pos5 with the structures of human and bacterial NAD kinases, revealed that Arg-293 of Pos5, corresponding to His-351 of human NAD kinase, confers a positive charge on the surface of NADH-binding site, whereas the corresponding His residue does not. Accordingly, conversion of the Arg-293 into a His residue reduced the ratio of NADH kinase activity to NAD kinase activity from 8.6 to 2.1. Conversely, simultaneous changes of Ala-330 and His-351 of human NAD kinase into Ser and Arg residues significantly increased the ratio of NADH kinase activity to NAD kinase activity from 0.043 to 1.39; human Ala-330 corresponds to Pos5 Ser-272, which interacts with the side chain of Arg-293. Arg-293 and Ser-272 were highly conserved in Pos5 homologs (putative NADH kinases), but not in putative NAD kinases. Thus, Arg-293 of Pos5 is a major determinant of NADH selectivity. Moreover, Ser-272 appears to assist Arg-293 in achieving the appropriate conformation.  相似文献   

12.
3α-Hydroxysteroid dehydrogenase/carbonyl reductase reversely catalyzes the oxidation of androsterone with NAD+ to form androstanedione and NADH. In this study, we investigated the function of active site residues N86, Y155, and K159 in NADH binding and catalysis in the reduction of androstanedione, using site-directed mutagenesis, steady-state kinetics, fluorescence quenching, and anisotropy measurements. The N86A, Y155F, and K159A mutant enzymes decreased the catalytic constant by 37- to 220-fold and increased the dissociation constant by 3- to 75-fold, respectively. Binding of NADH with wild-type and mutant enzymes caused different levels of fluorescence resonance energy transfer, implying a different orientation of nicotinamide ring versus W173. In addition, the enzyme-bound NADH decreased the fluorescence anisotropy value in the order WT > N86A > Y155F > K159A, indicating an increase in the mobility of the bound NADH for the mutants. Data suggest that hydrogen bonding with the hydroxyl group of nicotinamide ribose by K159 and Y155 is important to maintain the orientation of NADH and contributes greatly to the transition-state binding energy to facilitate the catalysis. N86 is important for stabilizing the position of K159. Substitution of alanine for N86 has a minor effect on NADH binding through K159, resulting in a slight increase in the mobility of the bound NADH and decreases in affinity and catalytic constant.  相似文献   

13.
Antifreeze proteins (AFPs) are essential components of many organisms adaptation to cold temperatures. Fish type III AFPs are divided into two groups, SP isoforms being much less active than QAE1 isoforms. Two type III AFPs from Zoarces viviparus, a QAE1 (ZvAFP13) and an SP (ZvAFP6) isoform, are here characterized and their crystal structures determined. We conclude that the higher activity of the QAE1 isoforms cannot be attributed to single residues, but rather a combination of structural effects. Furthermore both ZvAFP6 and ZvAFP13 crystal structures have water molecules around T18 equivalent to the tetrahedral-like waters previously identified in a neutron crystal structure. Interestingly, ZvAFP6 forms dimers in the crystal, with a significant dimer interface. The presence of ZvAFP6 dimers was confirmed in solution by native electrophoresis and gel filtration. To our knowledge this is the first report of dimerization of AFP type III proteins.  相似文献   

14.
The crystal structures of copper-containing nitrite reductase (CuNiR) from the thermophilic Gram-positive bacterium Geobacillus kaustophilus HTA426 and the amino (N)-terminal 68 residue-deleted mutant were determined at resolutions of 1.3 Å and 1.8 Å, respectively. Both structures show a striking resemblance with the overall structure of the well-known CuNiRs composed of two Greek key β-barrel domains; however, a remarkable structural difference was found in the N-terminal region. The unique region has one β-strand and one α-helix extended to the northern surface of the type-1 copper site. The superposition of the Geobacillus CuNiR model on the electron-transfer complex structure of CuNiR with the redox partner cytochrome c551 in other denitrifier system led us to infer that this region contributes to the transient binding with the partner protein during the interprotein electron transfer reaction in the Geobacillus system. Furthermore, electron-transfer kinetics experiments using N-terminal residue-deleted mutant and the redox partner, Geobacillus cytochrome c551, were carried out. These structural and kinetics studies demonstrate that the region is directly involved in the specific partner recognition.  相似文献   

15.
l-Sorbose reductase from Gluconobacter frateurii (SR) is an NADPH-dependent oxidoreductase. SR preferentially catalyzes the reversible reaction between d-sorbitol and l-sorbose with high substrate specificity. To elucidate the structural basis of the catalytic mechanism and the substrate specificity of SR, we have determined the structures of apo-SR, SR in complex with NADPH, and the inactive mutant (His116Leu) of SR in complex with NADPH and l-sorbose at 2.83 Å, 1.90 Å, and 1.80 Å resolutions, respectively. Our results show that SR belongs to the short-chain dehydrogenase/reductase (SDR) family and forms a tetrameric structure. Although His116 is not conserved among SDR family enzymes, the structures of SR have revealed that His116 is important for the stabilization of the proton relay system and for active-site conformation as a fourth catalytic residue. In the ternary complex structure, l-sorbose is recognized by 11 hydrogen bonds. Site-directed mutagenesis of residues around the l-sorbose-binding site has shown that the loss of almost full enzymatic activity was caused by not only the substitution of putative catalytic residues but also the substitution of the residue used for the recognition of the C4 hydroxyl groups of l-sorbose (Glu154) and of the residues used for the construction of the substrate-binding pocket (Cys146 and Gly188). The recognition of the C4 hydroxyl group of l-sorbose would be indispensable for the substrate specificity of SR, which recognizes only l-sorbose and d-sorbitol but not other sugars. Our results indicated that these residues were crucial for the substrate recognition and specificity of SR.  相似文献   

16.
17.
d-Arabinose isomerase (d-AI), also known as l-fucose isomerase (l-FI), catalyzes the aldose–ketose isomerization of d-arabinose to d-ribulose, and l-fucose to l-fuculose. Bacillus pallidus (B. pallidus) d-AI can catalyze isomerization of d-altrose to d-psicose, as well as d-arabinose and l-fucose. Three X-ray structures of B. pallidusd-AI in complexes with 2-methyl-2,4-pentadiol, glycerol and an inhibitor, l-fucitol, were determined at resolutions of 1.77, 1.60 and 2.60 Å, respectively. B. pallidusd-AI forms a homo-hexamer, and one subunit has three domains of almost equal size; two Rossmann fold domains and a mimic of the (β/α) barrel fold domain. A catalytic metal ion (Mn2+) was found in the active site coordinated by Glu342, Asp366 and His532, and an additional metal ion was found at the channel for the passage of a substrate coordinated by Asp453. The X-ray structures basically supported the ene-diol mechanism for the aldose–ketose isomerization by B. pallidusd-AI, as well as Escherichia coli (E. coli) l-FI, in which Glu342 and Asp366 facing each other at the catalytic metal ion transfer a proton from C2 to C1 and O1 to O2, acting as acid/base catalysts, respectively. However, considering the ionized state of Asp366, the catalytic reaction also possibly occurs through the negatively charged ene-diolate intermediate stabilized by the catalytic metal ion. A structural comparison with E. colil-FI showed that B. pallidusd-AI possibly interconverts between “open” and “closed” forms, and that the additional metal ion found in B. pallidusd-AI may help to stabilize the channel region.  相似文献   

18.
Apoptosis-inducing factor (AIF) and AMID (AIF-homologous mitochondrion-associated inducer of death) are flavoproteins. Although AIF was originally discovered as a caspase-independent cell death effector, bioenergetic roles of AIF, particularly relating to complex I functions, have since emerged. However, the role of AIF in mitochondrial respiration and redox metabolism has remained unknown. Here, we investigated the redox properties of human AIF and AMID by comparing them with yeast Ndi1, a type 2 NADH:ubiquinone oxidoreductase (NDH-2) regarded as alternative complex I. Isolated AIF and AMID containing naturally incorporated FAD displayed no NADH oxidase activities. However, after reconstituting isolated AIF or AMID into bacterial or mitochondrial membranes, N-terminally tagged AIF and AMID displayed substantial NADH:O2 activities and supported NADH-linked proton pumping activities in the host membranes almost as efficiently as Ndi1. NADH:ubiquinone-1 activities in the reconstituted membranes were highly sensitive to 2-n-heptyl-4-hydroxyquinoline-N-oxide (IC50 = ∼1 μm), a quinone-binding inhibitor. Overexpressing N-terminally tagged AIF and AMID enhanced the growth of a double knock-out Escherichia coli strain lacking complex I and NDH-2. In contrast, C-terminally tagged AIF and NADH-binding site mutants of N-terminally tagged AIF and AMID failed to show both NADH:O2 activity and the growth-enhancing effect. The disease mutant AIFΔR201 showed decreased NADH:O2 activity and growth-enhancing effect. Furthermore, we surprisingly found that the redox activities of N-terminally tagged AIF and AMID were sensitive to rotenone, a well known complex I inhibitor. We propose that AIF and AMID are previously unidentified mammalian NDH-2 enzymes, whose bioenergetic function could be supplemental NADH oxidation in cells.  相似文献   

19.
Malaria parasites adapt to the oxidative stress during their erythrocytic stages with the help of vital thioredoxin redox system and glutathione redox system. Glutathione reductase and thioredoxin reductase are important enzymes of these redox systems that help parasites to maintain an adequate intracellular redox environment. In the present study, activities of glutathione reductase and thioredoxin reductase were investigated in normal and Plasmodium berghei-infected mice red blood cells and their fractions. Activities of glutathione reductase and thioredoxin reductase in P. berghei-infected host erythrocytes were found to be higher than those in normal host cells. These enzymes were mainly confined to the cytosolic part of cell-free P. berghei. Full characterization and understanding of these enzymes may promise advances in chemotherapy of malaria.  相似文献   

20.
Cytoplasmic desulfothioredoxin (Dtrx) from the anaerobe Desulfovibrio vulgaris Hildenborough has been identified as a new member of the thiol disulfide oxidoreductase family. The active site of Dtrx contains a particular consensus sequence, CPHC, never seen in the cytoplasmic thioredoxins and generally found in periplasmic oxidases. Unlike canonical thioredoxins (Trx), Dtrx does not present any disulfide reductase activity, but it presents instead an unusual disulfide isomerase activity. We have used NMR spectroscopy to gain insights into the structure and the catalytic mechanism of this unusual Dtrx. The redox potential of Dtrx (-181 mV) is significantly less reducing than that of canonical Trx. A pH dependence study allowed the determination of the pK(a) of all protonable residues, including the cysteine and histidine residues. Thus, the pK(a) values for the thiol group of Cys(31) and Cys(34) are 4.8 and 11.3, respectively. The His(33) pK(a) value, experimentally determined for the first time, differs notably as a function of the redox states, 7.2 for the reduced state and 4.6 for the oxidized state. These data suggest an important role for His(33) in the molecular mechanism of Dtrx catalysis that is confirmed by the properties of mutant DtrxH33G protein. The NMR structure of Dtrx shows a different charge repartition compared with canonical Trx. The results presented are likely indicative of the involvement of this protein in the catalysis of substrates specific of the anaerobe cytoplasm of DvH. The study of Dtrx is an important step toward revealing the molecular details of the thiol-disulfide oxidoreductase catalytic mechanism.  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