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 15 毫秒
1.
Isocitrate dehydrogenases (IDHs) catalyze oxidative decarboxylation of isocitrate (ICT) into alpha-ketoglutarate (AKG). We report here the crystal structures of Saccharomyces cerevesiae mitochondrial NADP-IDH Idp1p in binary complexes with coenzyme NADP, or substrate ICT, or product AKG, and in a quaternary complex with NADPH, AKG, and Ca(2+), which represent different enzymatic states during the catalytic reaction. Analyses of these structures identify key residues involved in the binding of these ligands. Comparisons among these structures and with the previously reported structures of other NADP-IDHs reveal that eukaryotic NADP-IDHs undergo substantial conformational changes during the catalytic reaction. Binding or release of the ligands can cause significant conformational changes of the structural elements composing the active site, leading to rotation of the large domain relative to the small and clasp domains along two hinge regions (residues 118-124 and residues 284-287) while maintaining the integrity of its secondary structural elements, and thus, formation of at least three distinct overall conformations. Specifically, the enzyme adopts an open conformation when bound to NADP, a quasi-closed conformation when bound to ICT or AKG, and a fully closed conformation when bound to NADP, ICT, and Ca(2+) in the pseudo-Michaelis complex or with NADPH, AKG, and Ca(2+) in the product state. The conformational changes of eukaryotic NADP-IDHs are quite different from those of Escherichia coli NADP-IDH, for which significant conformational changes are observed only between two forms of the apo enzyme, suggesting that the catalytic mechanism of eukaryotic NADP-IDHs is more complex than that of EcIDH, and involves more fine-tuned conformational changes.  相似文献   

2.
NADP+-dependent monomeric isocitrate dehydrogenase (IDH) from the nitrogen-fixing bacterium Azotobacter vinelandii (AvIDH) is one of members of the beta-decarboxylating dehydrogenase family and catalyzes the dehydration and decarboxylation of isocitrate to yield 2-oxoglutrate and CO2 in the Krebs cycle. We solved the crystal structure of the AvIDH in complex with cofactor NADP+ (AvIDH-NADP+ complex). The final refined model shows the closed form that has never been detected in any previously solved structures of beta-decarboxylating dehydrogenases. The structure also reveals all of the residues that interact with NADP+. The structure-based sequence alignment reveals that these residues were not conserved in any other dimeric NADP+-dependent IDHs. Therefore the NADP+ specificity of the monomeric and dimeric IDHs was independently acquired through the evolutional process. The AvIDH was known to show an exceptionally high turnover rate. The structure of the AvIDH-NADP+ complex indicates that one loop, which is not present in the Escherichia coli IDHs, reliably stabilizes the conformation of the nicotinamide mononucleotide of the bound NADP+ by forming a few hydrogen bonds, and such interactions are considered to be important for the monomeric enzyme to initiate the hydride transfer reaction immediately. Finally, the structure of the AvIDH is compared with that of other dimeric NADP-IDHs. Several structural features demonstrate that the monomeric IDHs are structurally more related to the eukaryotic dimeric IDHs than to the bacterial dimeric IDHs.  相似文献   

3.
Both monomeric and dimeric NADP+-dependent isocitrate dehydrogenase (IDH) belong to the metal-dependent beta-decarboxylating dehydrogenase family and catalyze the oxidative decarboxylation from 2R,3S-isocitrate to yield 2-oxoglutarate, CO2, and NADPH. It is important to solve the structures of IDHs from various species to correlate with its function and evolutionary significance. So far, only two crystal structures of substrate/cofactor-bound (isocitrate/NADP) NADP+-dependent monomeric IDH from Azotobacter vinelandii (AvIDH) have been solved. Herein, we report for the first time the substrate/cofactor-free structure of a monomeric NADP+-dependent IDH from Corynebacterium glutamicum (CgIDH) in the presence of Mg2+. The 1.75 A structure of CgIDH-Mg2+ showed a distinct open conformation in contrast to the closed conformation of AvIDH-isocitrate/NADP+ complexes. Fluorescence studies on CgIDH in the presence of isocitrate/or NADP+ suggest the presence of low energy barrier conformers. In CgIDH, the amino acid residues corresponding to the Escherichia coli IDH phosphorylation-loop are alpha-helical compared with the more flexible random-coil region in the E. coli protein where IDH activation is controlled by phosphorylation. This more structured region supports the idea that activation of CgIDH is not controlled by phosphorylation. Monomeric NADP+-specific IDHs have been identified from about 50 different bacterial species, such as proteobacteria, actinobacteria, and planctomycetes, whereas, dimeric NADP+-dependent IDHs are diversified in both prokaryotes and eukaryotes. We have constructed a phylogenetic tree based on amino acid sequences of all bacterial monomeric NADP+-dependent IDHs and also another one with specifically chosen species which either contains both monomeric and dimeric NADP+-dependent IDHs or have monomeric NADP+-dependent, as well as NAD+-dependent IDHs. This is done to examine evolutionary relationships.  相似文献   

4.
D Garland  H G Nimmo 《FEBS letters》1984,165(2):259-264
NADP+ can protect active isocitrate dehydrogenase against attack by several proteases. Inactive phosphorylated isocitrate dehydrogenase is much less susceptible to proteolysis than the active enzyme, and it is not protected by NADP+. The results suggest that binding of NADP+ to, or phosphorylation of, active isocitrate dehydrogenase induces similar conformational states. Fluorescence titration experiments show that NADPH can bind to active but not to inactive isocitrate dehydrogenase. It is suggested that the phosphorylation of isocitrate dehydrogenase may occur close to its coenzyme binding site.  相似文献   

5.
Mitochondrial NAD(+)-specific isocitrate dehydrogenases (IDHs) are key regulators of flux through biosynthetic and oxidative pathways in response to cellular energy levels. Here we present the first structures of a eukaryotic member of this enzyme family, the allosteric, hetero-octameric, NAD(+)-specific IDH from yeast in three forms: 1) without ligands, 2) with bound analog citrate, and 3) with bound citrate + AMP. The structures reveal the molecular basis for ligand binding to homologous but distinct regulatory and catalytic sites positioned at the interfaces between IDH1 and IDH2 subunits and define pathways of communication between heterodimers and heterotetramers in the hetero-octamer. Disulfide bonds observed at the heterotetrameric interfaces in the unliganded IDH hetero-octamer are reduced in the ligand-bound forms, suggesting a redox regulatory mechanism that may be analogous to the "on-off" regulation of non-allosteric bacterial IDHs via phosphorylation. The results strongly suggest that eukaryotic IDH enzymes are exquisitely tuned to ensure that allosteric activation occurs only when concentrations of isocitrate are elevated.  相似文献   

6.
The monomeric isocitrate dehydrogenase (IDH) of Corynebacterium glutamicum is compared to the topologically distinct dimeric IDH of Escherichia coli. Both IDHs have evolved to efficiently catalyze identical reactions with similar pH optimum as well as striking specificity toward NADP and isocitrate. However, the monomeric IDH is 10-fold more active (calculated as kcat/Km.isocitrate/Km.NADP) and 7-fold more NADP-specific than the dimeric enzyme, favoring NADP over NAD by a factor of 50,000. Such an extraordinary coenzyme specificity is not rivaled by any other characterized dehydrogenases. In addition, the monomeric enzyme is 10-fold more specific for isocitrate. The spectacular substrate specificity may be predominantly attributed to the isocitrate-assisted stabilization of catalytic complex during hydride transfer. No significant overall sequence identity is found between the monomeric and dimeric enzymes. However, structure-based alignment leads to the identification of three regions in the monomeric enzyme that match closely the three motifs located in the central region of dimeric IDHs and the homologous isopropylmalate dehydrogenases. The role of Lys253 as catalytic residue has been demonstrated by site-directed mutagenesis. Our results suggest that monomeric and dimeric forms of IDHs are functionally and structurally homologous.  相似文献   

7.
In most living organisms, isocitrate dehydrogenases (IDHs) convert isocitrate into ɑ-ketoglutarate (ɑ-KG). Phylogenetic analyses divide the IDH protein family into two subgroups: types I and II. Based on cofactor usage, IDHs are either NAD+-specific (NAD-IDH) or NADP+-specific (NADP-IDH); NADP-IDH evolved from NAD-IDH. Type I IDHs include NAD-IDHs and NADP-IDHs; however, no type II NAD-IDHs have been reported to date. This study reports a novel type II NAD-IDH from the marine bacterium Congregibacter litoralis KT71 (ClIDH, GenBank accession no. EAQ96042). His-tagged recombinant ClIDH was produced in Escherichia coli and purified; the recombinant enzyme was NAD+-specific and showed no detectable activity with NADP+. The K m values of the enzyme for NAD+ were 262.6±7.4 μM or 309.1±11.2 μM with Mg2+ or Mn2+ as the divalent cation, respectively. The coenzyme specificity of a ClIDH Asp487Arg/Leu488His mutant was altered, and the preference of the mutant for NADP+ was approximately 24-fold higher than that for NAD+, suggesting that ClIDH is an NAD+-specific ancestral enzyme in the type II IDH subgroup. Gel filtration and analytical ultracentrifugation analyses revealed the homohexameric structure of ClIDH, which is the first IDH hexamer discovered thus far. A 163-amino acid segment of CIIDH is essential to maintain its polymerization structure and activity, as a truncated version lacking this region forms a non-functional monomer. ClIDH was dependent on divalent cations, the most effective being Mn2+. The maximal activity of purified recombinant ClIDH was achieved at 35°C and pH 7.5, and a heat inactivation experiment showed that a 20-min incubation at 33°C caused a 50% loss of ClIDH activity. The discovery of a NAD+-specific, type II IDH fills a gap in the current classification of IDHs, and sheds light on the evolution of type II IDHs.  相似文献   

8.
Equilibrium binding studies demonstrate that purified Escherichia coli isocitrate dehydrogenase binds isocitrate, alpha-ketoglutarate, NADP, and NADPH at 1:1 ratios of substrate to enzyme monomer. The phosphorylated enzyme, which is completely inactive, is unable to bind isocitrate but retains the ability to bind NADP and NADPH. Replacement of serine 113, which is the site of phosphorylation, by aspartate results in an inactive enzyme that is unable to bind isocitrate. Replacement of the same serine with other amino acids (lysine, threonine, cysteine, tyrosine, and alanine) produces active enzymes that bind both substrates. Hence, the negative charge of an aspartate or a phosphorylated serine at site 113 inactivates the enzyme by preventing the binding of isocitrate.  相似文献   

9.
The crystal structure of porcine heart mitochondrial NADP+-dependent isocitrate dehydrogenase (IDH) complexed with Mn2+ and isocitrate was solved to a resolution of 1.85 A. The enzyme was expressed in Escherichia coli, purified as a fusion protein with maltose binding protein, and cleaved with thrombin to yield homogeneous enzyme. The structure was determined by multiwavelength anomalous diffraction phasing using selenium substitution in the form of selenomethionine as the anomalous scatterer. The porcine NADP+-IDH enzyme is structurally compared with the previously solved structures of IDH from E. coli and Bacillus subtilis that share 16 and 17% identity, respectively, with the mammalian enzyme. The porcine enzyme has a protein fold similar to the bacterial IDH structures with each monomer folding into two domains. However, considerable differences exist between the bacterial and mammalian forms of IDH in regions connecting core secondary structure. Based on the alignment of sequence and structure among the porcine, E. coli, and B. subtilis IDH, a putative phosphorylation site has been identified for the mammalian enzyme. The active site, including the bound Mn2+-isocitrate complex, is highly ordered and, therefore, mechanistically informative. The consensus IDH mechanism predicts that the Mn2+-bound hydroxyl of isocitrate is deprotonated prior to its NADP+-dependent oxidation. The present crystal structure has an active site water that is well positioned to accept the proton and ultimately transfer the proton to solvent through an additional bound water.  相似文献   

10.
Isocitrate dehydrogenase from Aeropyrum pernix (ApIDH) is a homodimeric enzyme that belongs to the beta-decarboxylating dehydrogenase family and is the most thermostable IDH identified. It catalyzes the NADP+ and metal-dependent oxidative decarboxylation of isocitrate to alpha-ketoglutarate. We have solved the crystal structures of a native ApIDH at 2.2 A, a pseudo-native ApIDH at 2.1 A, and of ApIDH in complex with NADP+, Ca2+ and d-isocitrate at 2.3 A. The pseudo-native ApIDH is in complex with etheno-NADP+ which was located at the surface instead of in the active site revealing a novel adenine-nucleotide binding site in ApIDH. The native and the pseudo-native ApIDHs were found in an open conformation, whereas one of the subunits of the ternary complex was closed upon substrate binding. The closed subunit showed a domain rotation of 19 degrees compared to the open subunit. The binding of isocitrate in the closed subunit was identical with that of the binary complex of porcine mitochondrial IDH, whereas the binding of NADP+ was similar to that of the ternary complex of IDH from Escherichiacoli. The reaction mechanism is likely to be conserved in the different IDHs. A proton relay chain involving at least five solvent molecules, the 5'-phosphate group of the nicotinamide-ribose and a coupled lysine-tyrosine pair in the active site, is postulated as essential in both the initial and the final steps of the catalytic reaction of IDH. ApIDH was found to be highly homologous to the mesophilic IDHs and was subjected to a comparative analysis in order to find differences that could explain the large difference in thermostability. Mutational studies revealed that a disulfide bond at the N terminus and a seven-membered inter-domain ionic network at the surface are major determinants for the higher thermostability of ApIDH compared to EcIDH. Furthermore, the total number of ion pairs was dramatically higher in ApIDH compared to the mesophilic IDHs if a cutoff of 4.2 A was used. A calculated net charge of only +1 compared to -19 and -25 in EcIDH and BsIDH, respectively, suggested a high degree of electrostatic optimization, which is known to be an important determinant for increased thermostability.  相似文献   

11.
The structure of crystalline porcine mitochondrial NADP-dependent isocitrate dehydrogenase (IDH) has been determined in complex with Mn2+-isocitrate. Based on structural alignment between this porcine enzyme and seven determined crystal structures of complexes of NADP with bacterial IDHs, Arg83, Thr311, and Asn328 were chosen as targets for site-directed mutagenesis of porcine IDH. The circular dichroism spectra of purified wild-type and mutant enzymes are similar. The mutant enzymes exhibit little change in Km for isocitrate or Mn2+, showing that these residues are not involved in substrate binding. In contrast, the Arg83 mutants, Asn328 mutants, and T311A exhibit 3-20-fold increase in the Km(NADP). We propose that Arg83 enhances NADP affinity by hydrogen bonding with the 3'-OH of the nicotinamide ribose, whereas Asn328 hydrogen bonds with N1 of adenine. The pH dependence of Vmax for Arg83 and Asn328 mutants is similar to that of wild-type enzyme, but for all the Thr311 mutants, pK(es) is increased from 5.2 in the wild type to approximately 6.0. We have previously attributed the pH dependence of Vmax to the deprotonation of the metal-bound hydroxyl of isocitrate in the enzyme-substrate complex, prior to the transfer of a hydride from isocitrate to NADP's nicotinamide moiety. Thr311 interacts with the nicotinamide ribose and is the closest of the target amino acids to the nicotinamide ring. Distortion of the nicotinamide by Thr311 mutation will likely be transmitted to Mn2+-isocitrate resulting in an altered pK(es). Because porcine and human mitochondrial NADP-IDH have 95% sequence identity, these results should be applicable to the human enzyme.  相似文献   

12.
J S McKee  R Hlodan  H G Nimmo 《Biochimie》1989,71(9-10):1059-1064
Escherichia coli isocitrate dehydrogenase is completely inactivated by phosphorylation of a single serine residue per subunit. We have examined the conformations of the active and phosphorylated forms of the enzyme using circular dichroism spectroscopy. The results support the view that phosphorylation prevents the binding of NADP, probably by direct blocking of the coenzyme-binding site. Labelling studies suggest that an arginine residue at the coenzyme-binding site may be close to the phosphorylatable serine residue. The phosphorylation of isocitrate dehydrogenase is thus unusual in that it occurs at the active site of the enzyme. We therefore investigated the recognition of isocitrate dehydrogenase by isocitrate dehydrogenase kinase/phosphatase. The kinase activity of this enzyme can phosphorylate intact isocitrate dehydrogenase but not proteolytic fragments derived from it, nor a synthetic peptide corresponding to the sequence round the phosphorylation site.  相似文献   

13.
Isocitrate dehydrogenase (IDH) has been studied extensively due to its central role in the Krebs cycle, catalyzing the oxidative NAD(P)(+)-dependent decarboxylation of isocitrate to alpha-ketoglutarate and CO(2). Here, we present the first crystal structure of IDH from a psychrophilic bacterium, Desulfotalea psychrophila (DpIDH). The structural information is combined with a detailed biochemical characterization and a comparative study with IDHs from the mesophilic bacterium Desulfitobacterium hafniense (DhIDH), porcine (PcIDH), human cytosolic (HcIDH) and the hyperthermophilic Thermotoga maritima (TmIDH). DpIDH was found to have a higher melting temperature (T(m)=66.9 degrees C) than its mesophilic homologues and a suboptimal catalytic efficiency at low temperatures. The thermodynamic activation parameters indicated a disordered active site, as seen also for the drastic increase in K(m) for isocitrate at elevated temperatures. A methionine cluster situated at the dimeric interface between the two active sites and a cluster of destabilizing charged amino acids in a region close to the active site might explain the poor isocitrate affinity. On the other hand, DpIDH was optimized for interacting with NADP(+) and the crystal structure revealed unique interactions with the cofactor. The highly acidic surface, destabilizing charged residues, fewer ion pairs and reduced size of ionic networks in DpIDH suggest a flexible global structure. However, strategic placement of ionic interactions stabilizing the N and C termini, and additional ionic interactions in the clasp domain as well as two enlarged aromatic clusters might counteract the destabilizing interactions and promote the increased thermal stability. The structure analysis of DpIDH illustrates how psychrophilic enzymes can adjust their flexibility in dynamic regions during their catalytic cycle without compromising the global stability of the protein.  相似文献   

14.
With the aim of gaining insight into the molecular and phylogenetic relationships of isocitrate dehydrogenase (IDH) from hyperthermophiles, we carried out a comparative study of putative IDHs identified in the genomes of the eubacterium Thermotoga maritima and the archaea Aeropyrum pernix and Pyrococcus furiosus. An optimum for activity at 90 degrees C or above was found for each IDH. PfIDH and ApIDH were the most thermostable with a melting temperature of 103.7 and 109.9 degrees C, respectively, compared with 98.3 and 98.5 degrees C for TmIDH and AfIDH, respectively. Analytical ultracentrifugation revealed a tetrameric oligomeric state for TmIDH and a homodimeric state for ApIDH and PfIDH. TmIDH and ApIDH were NADP-dependent (K(m)((NADP)) of 55.2 and 44.4 microm, respectively) whereas PfIDH was NAD-dependent (K(m)((NAD)) of 68.3 microm). These data document that TmIDH represents a novel tetrameric NADP-dependent form of IDH and that PfIDH is a homodimeric NAD-dependent IDH not previously found among the archaea. The homodimeric NADP-IDH present in A. pernix is the most common form of IDH known so far. The evolutionary relationships of ApIDH, PfIDH, and TmIDH with all of the available amino acid sequences of di- and multimeric IDHs are described and discussed.  相似文献   

15.
NADP(+)-dependent isocitrate dehydrogenase is a member of the beta-decarboxylating dehydrogenase family and catalyzes the oxidative decarboxylation reaction from 2R,3S-isocitrate to yield 2-oxoglutarate and CO(2) in the Krebs cycle. Although most prokaryotic NADP(+)-dependent isocitrate dehydrogenases (IDHs) are homodimeric enzymes, the monomeric IDH with a molecular weight of 80-100 kDa has been found in a few species of bacteria. The 1.95 A crystal structure of the monomeric IDH revealed that it consists of two distinct domains, and its folding topology is related to the dimeric IDH. The structure of the large domain repeats a motif observed in the dimeric IDH. Such a fusional structure by domain duplication enables a single polypeptide chain to form a structure at the catalytic site that is homologous to the dimeric IDH, the catalytic site of which is located at the interface of two identical subunits.  相似文献   

16.
Isocitrate dehydrogenase (IDH) from the hyperthermophile Thermotoga maritima (TmIDH) catalyses NADP+- and metal-dependent oxidative decarboxylation of isocitrate to alpha-ketoglutarate. It belongs to the beta-decarboxylating dehydrogenase family and is the only hyperthermostable IDH identified within subfamily II. Furthermore, it is the only IDH that has been characterized as both dimeric and tetrameric in solution. We solved the crystal structure of the dimeric apo form of TmIDH at 2.2 A. The R-factor of the refined model was 18.5% (R(free) 22.4%). The conformation of the TmIDH structure was open and showed a domain rotation of 25-30 degrees compared with closed IDHs. The separate domains were found to be homologous to those of the mesophilic mammalian IDHs of subfamily II and were subjected to a comparative analysis in order to find differences that could explain the large difference in thermostability. Mutational studies revealed that stabilization of the N- and C-termini via long-range electrostatic interactions were important for the higher thermostability of TmIDH. Moreover, the number of intra- and intersubunit ion pairs was higher and the ionic networks were larger compared with the mesophilic IDHs. Other factors likely to confer higher stability in TmIDH were a less hydrophobic and more charged accessible surface, a more hydrophobic subunit interface, more hydrogen bonds per residue and a few loop deletions. The residues responsible for the binding of isocitrate and NADP+ were found to be highly conserved between TmIDH and the mammalian IDHs and it is likely that the reaction mechanism is the same.  相似文献   

17.
L C Costello  R Franklin  R Stacey 《Enzyme》1976,21(6):495-506
Mitochondrial preparations isolated from rat ventral prostate were capable of oxidizing isocitrate by way of NADP isocitrate dehydrogenase (NADP-IDH) and NAD-IDH. NAD-IDH activity required ADP for activation. The pH responses for NAD-IDH and NADP-IDH were quite different. The results indicated that two different enzymes were involved in the NAD- and NADP-IDH activities. Indirect evidence indicated that NADPH-NAD transhydrogenase activity might also be involved in the mitochondrial pathway for isocitrate oxidation. NADP-IDH activity was significantly greater than NAD-IDH activity. The oxidation of isocitrate through IDH activity was coupled to the cytochrome system by NADPH- and NADH-cytochrome c reductase activities. Citrate, via isocitrate, oxidation proceeded at a much slower rate suggesting that aconitase activity could be limiting in the oxidation of citrate. In comparison to other tissues, the prostate oxidative enzyme activities are considerably lower. The results suggest that the accumulation of high prostate citrate levels is not due to a limitation imposed by a lack of IDH activity in prostate mitochondria.  相似文献   

18.
NADP(+)-dependent isocitrate dehydrogenase [IDH: EC 1.1.1.42] was purified to electrophoretic homogeneity from Vibrio parahaemolyticus Y-4, and shown to be a monomeric protein of molecular weight 80,000 with a pI of 5.0. The amino acid composition and partial sequence at the N-terminus resembled those reported for other bacterial monomeric IDHs. Immunotitration with antisera to the monomeric and dimeric enzymes (antisera to IDH-II and -I of Vibrio ABE-1) showed an immunochemical distinction between the monomeric and dimeric IDHs, but there is similarity within the IDHs of each group. The circular dichroism spectra of the native and heat-denatured enzyme are also similar to those of monomeric IDH (IDH-II of Vibrio ABE-1). These monomeric IDHs are proteins comprising 17-22% helix and 25-35% beta-pleated sheet in the native state.  相似文献   

19.
20.
The role of Asp-328 and Ile-329 as a cofactor discrimination site of the NAD-dependent isocitrate dehydrognase (NAD-IDH) from Pyrococcus furiosus has been verified by replacing these residues with Lys and Tyr, respectively, which are the corresponding residues in NADP-IDH from Escherichia coli. The Asp-328-Lys mutant showed dual coenzyme specificity, whereas introduction of the double mutation, Asp-328-Lys/Ile-329-Tyr shifted the cofactor preference from NAD to NADP. NADP-dependent P. furiosus IDH retained thermostability and thermoactivity compared with NAD-IDH.  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