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 46 毫秒
1.
Doyle SA  Fung SY  Koshland DE 《Biochemistry》2000,39(46):14348-14355
Despite the structural similarities between isocitrate and isopropylmalate, isocitrate dehydrogenase (IDH) exhibits a strong preference for its natural substrate. Using a combination of rational and random mutagenesis, we have engineered IDH to use isopropylmalate as a substrate. Rationally designed mutations were based on comparison of IDH to a similar enzyme, isopropylmalate dehydrogenase (IPMDH). A chimeric enzyme that replaced an active site loop-helix motif with IPMDH sequences exhibited no activity toward isopropylmalate, and site-directed mutants that replaced IDH residues with their IPMDH equivalents only showed small improvements in k(cat). Random mutants targeted the IDH active site at positions 113 (substituted with glutamate), 115, and 116 (both randomized) and were screened for activity toward isopropylmalate. Six mutants were identified that exhibited up to an 8-fold improvement in k(cat) and increased the apparent binding affinity by as much as a factor of 80. In addition to the S113E mutation, five other mutants contained substitutions at positions 115 and/or 116. Most small hydrophobic substitutions at position 116 improved activity, possibly by generating space to accommodate the isopropyl group of isopropylmalate; however, substitution with serine yielded the most improvement in k(cat). Only two substitutions were identified at position 115, which suggests a more specific role for the wild-type asparagine residue in the utilization of isopropylmalate. Since interactions between neighboring residues in this region greatly influenced the effects of each other in unexpected ways, structural solutions were best identified in combinations, as allowed by random mutagenesis.  相似文献   

2.
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   

3.
Inactivation of Escherichia coli isocitrate dehydrogenase upon phosphorylation at S113 depends upon the direct electrostatic repulsion of the negatively charged gamma-carboxylate of isocitrate by the negatively charged phosphoserine. The effect is mimicked by replacing S113 with aspartate or glutamate, which reduce performance (kcat/K(i).isocitrat/ Km.NADP) by a factor of 10(7). Here, we demonstrate that the inactivating effects of the electrostatic repulsion are completely eliminated by a second-site mutation, and provide the structural basis for this striking example of intragenic suppression. N115 is adjacent to S113 on one face of the D-helix, interacts with isocitrate and NADP+, and has been postulated to serve in both substrate binding and in catalysis. The single N115L substitution reduces affinity for isocitrate by a factor of 50 and performance by a factor of 500. However, the N115L substitution completely suppresses the inactivating electrostatic effects of S113D or S113E: the performance of the double mutants is 10(5) higher than the S113D and S113E single mutants. These mutations have little effect on the kinetics of alternative substrates, which lack the charged gamma-carboxylate of isocitrate. Both glutamate and aspartate at site 113 remain fully ionized in the presence of leucine. In the crystal structure of the N115L mutant, the leucine adopts a different conformer from the wild-type asparagine. Repacking around the leucine forces the amino-terminus of the D-helix away from the rest of the active site. The hydrogen bond between E113 and N115 in the S113E single mutant is broken in the S113E/N115L mutant, allowing the glutamate side chain to move away from the gamma-carboxylate of isocitrate. These movements increase the distance between the carboxylates, diminish the electrostatic repulsion, and lead to the remarkably high activity of the S113E/N115L mutant.  相似文献   

4.
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   

5.
Inositol monophosphatase (IMPase) family of proteins are Mg(2+) activated Li(+) inhibited class of ubiquitous enzymes with promiscuous substrate specificity. Herein, the molecular basis of IMPase substrate specificity is delineated by comparative crystal structural analysis of a Staphylococcal dual specific IMPase/NADP(H) phosphatase (SaIMPase - I) with other IMPases of different substrate compatibility, empowered by in silico docking and Escherichia coli SuhB mutagenesis analysis. Unlike its eubacterial and eukaryotic NADP(H) non-hydrolyzing counterparts, the composite structure of SaIMPase - I active site pocket exhibits high structural resemblance with archaeal NADP(H) hydrolyzing dual specific IMPase/FBPase. The large and shallow SaIMPase - I active site cleft efficiently accommodate large incoming substrates like NADP(H), and therefore, justifies the eminent NADP(H) phosphatase activity of SaIMPase - I. Compared to other NADP(H) non-hydrolyzing IMPases, the profound difference in active site topology as well as the unique NADP(H) recognition capability of SaIMPase - I stems from the differential length and orientation of a distant helix α4 (in human and bovine α5) and its preceding loop. We identified the length of α4 and its preceding loop as the most crucial factor that regulates IMPase substrate specificity by employing a size exclusion mechanism. Hence, in SaIMPase - I, the substrate promiscuity is a gain of function by trimming the length of α4 and its preceding loop, compared to other NADP(H) non-hydrolyzing IMPases. This study thus provides a biochemical - structural framework revealing the length and orientation of α4 and its preceding loop as the predisposing factor for the determination of IMPase substrate specificity.  相似文献   

6.
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   

7.
The kinetic mechanism of the oxidative decarboxylation of 2R,3S-isopropylmalate by the NAD-dependent isopropylmalate dehydrogenase of Thermus thermophilus was investigated. Initial rate results typical of random or steady-state ordered sequential mechanisms are obtained for both the wild-type and two mutant enzymes (E87G and E87Q) regardless of whether natural or alternative substrates (2R-malate, 2R,3S-tartrate and/or NADP) are utilized. Initial rate data fail to converge on a rapid equilibrium-ordered pattern despite marked reductions in specificity (kcat/Km) caused by the mutations and alternative substrates. Although the inhibition studies alone might suggest an ordered kinetic mechanism with cofactor binding first, a detailed analysis reveals that the expected noncompetitive patterns appear uncompetitive because the dissociation constants from the ternary complexes are far smaller than those from the binary complexes. Equilibrium fluorescence studies both confirm the random binding of substrates and the kinetic estimates of the dissociation constants of the substrates from the binary complexes. The latter are not distributed markedly by the mutations at site 87. Mutations at site 87 do not affect the dissociation constants from the binary complexes, but do greatly increase the Michaelis constants, indicating that E87 helps stabilize the Michaelis complex of the wild-type enzyme. The available structural data, the patterns of the kinetics results, and the structure of a pseudo-Michaelis complex of the homologous isocitrate dehydrogenase of Escherichia coli suggest that E87 interacts with the nicotinamide ring.  相似文献   

8.
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   

9.
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   

10.
The substrate specificity of the NADP-dependent isocitrate dehydrogenase of Escherichia coli was investigated by combining site-directed mutagenesis and utilization of alternative substrates. A comparison of the kinetics of the wild-type enzyme with 2R-malate reveals that the gamma-carboxylate of 2R,3S-isocitrate contributes a factor of 12,000,000 to enzyme performance. Analysis of kinetic data compiled for 10 enzymes and nine different substrates reveals that a factor of 1,650 can be ascribed to the hydrogen bond formed between S113 and the gamma-carboxylate of bound isocitrate, a factor of 150 to the negative charge of the gamma-carboxylate, and a factor of 50 for the gamma-methyl. These results are entirely consistent with X-ray structures of Michaelis complexes that show a hydrogen bond positions the gamma-carboxylate of isocitrate so that a salt bridge can form to the nicotinamide ring of NADP.  相似文献   

11.
The enzyme isocitrate dehydrogenase (IDH, EC 1.1.1.42) can exhibit activation by one of its products, NADPH. This activation is competitively inhibited by the substrate NADP+, whereas NADPH competes with NADP+ for the catalytic site. Experimental observations briefly presented here have shown that if IDH is coupled to another enzyme, diaphorase (EC 1.8.1.4), which transforms NADPH into NADP+, the system can attain either one of two stable states, corresponding to a low and a high NADPH concentration. The evolution toward either one of these stable states depends on the time of addition of diaphorase to the medium containing IDH and its substrate NADP+. We present a theoretical and numerical analysis of a model for the IDH-diaphorase bienzymatic system, based on the regulatory properties of IDH. The results confirm the occurrence of bistability for parameter values derived from the experiments. Depending on the total concentration of NADP+ plus NADPH and the concentration of IDH, the system can either admit a single steady state or display bistability. We obtain an expression for the critical time t*, before which diaphorase addition leads to the lower steady state and after which addition of the enzyme leads to the upper steady state of NADPH. The analysis is extended to the case where the second substrate of IDH, isocitrate, is consumed in the course of the reaction without being regenerated. Bistability occurs only as a transient phenomenon in these conditions.  相似文献   

12.
Heterozygously expressed single-point mutations in isocitrate dehydrogenase 1 and 2 (IDH1 and IDH2, respectively) render these dimeric enzymes capable of producing the novel metabolite α-hydroxyglutarate (αHG). Accumulation of αHG is used as a biomarker for a number of cancer types, helping to identify tumors with similar IDH mutations. With IDH1, it has been shown that one role of the mutation is to increase the rate of conversion from αKG to αHG. To improve our understanding of the function of this mutation, we have detailed the kinetics of the normal (isocitrate to αKG) and neomorphic (αKG to αHG) reactions, as well as the coupled conversion of isocitrate to αHG. We find that the mutant IDH1 is very efficient in this coupled reaction, with the ability to form αHG from isocitrate and NADP(+). The wild type/wild type IDH1 is also able to catalyze this conversion, though it is much more sensitive to concentrations of isocitrate. This difference in behavior can be attributed to the competitive binding between isocitrate and αKG, which is made more favorable for αKG by the neomorphic mutation at arginine 132. Thus, each partial reaction in the heterodimer is functionally isolated from the other. To test whether there is a cooperative effect resulting from the two subunits being in a dimer, we selectively inactivated each subunit with a secondary mutation in the NADP/H binding site. We observed that the remaining, active subunit was unaffected in its associated activity, reinforcing the notion of each subunit being functionally independent. This was further demonstrated using a monomeric form of IDH from Azotobacter vinelandii, which can be shown to gain the same neomorphic reaction when a homologous mutation is introduced into that protein.  相似文献   

13.
Isocitrate dehydrogenase (IDH)(1) of Escherichia coli is regulated by a bifunctional protein, IDH kinase/phosphatase. In this paper, we demonstrate that the effectors controlling these activities belong to two distinct classes that differ in mechanism and in the locations of their binding sites. NADPH and isocitrate are representative members of one of these effector classes. NADPH inhibits both IDH kinase and IDH phosphatase, whereas isocitrate inhibits only IDH kinase. Isocitrate can "activate" IDH phosphatase by reversing product inhibition by dephospho-IDH. Mutations in icd, which encodes IDH, had parallel effects on the binding of these ligands to the IDH active site and on their effects on IDH kinase and phosphatase, indicating that these ligands regulate IDH kinase/phosphatase through the IDH active site. Kinetic analyses suggested that isocitrate and NADPH prevent formation of the complex between IDH kinase/phosphatase and its protein substrate. AMP, 3-phosphoglycerate, and pyruvate represent a class of regulatory ligands that is distinct from that which includes isocitrate and NADPH. These ligands bind directly to IDH kinase/phosphatase, a conclusion which is supported by the observation that they inhibit the IDH-independent ATPase activity of this enzyme. These effector classes can also be distinguished by the observation that mutant derivatives of IDH kinase/phosphatase expressed from aceK3 and aceK4 exhibited dramatic changes in their responses to AMP, 3-phosphoglycerate, and pyruvate but not to NADPH and isocitrate.  相似文献   

14.
Pyridoxal 5'-phosphate is an essential cofactor for many enzymes responsible for the metabolic conversions of amino acids. Two pathways for its de novo synthesis are known. The pathway utilized by Escherichia coli consists of six enzymatic steps catalyzed by six different enzymes. The fourth step is catalyzed by 4-hydroxythreonine-4-phosphate dehydrogenase (PdxA, E.C. 1.1.1.262), which converts 4-hydroxy-l-threonine phosphate (HTP) to 3-amino-2-oxopropyl phosphate. This divalent metal ion-dependent enzyme has a strict requirement for the phosphate ester form of the substrate HTP, but can utilize either NADP+ or NAD+ as redox cofactor. We report the crystal structure of E. coli PdxA and its complex with HTP and Zn2+. The protein forms tightly bound dimers. Each monomer has an alpha/beta/alpha-fold and can be divided into two subdomains. The active site is located at the dimer interface, within a cleft between the two subdomains and involves residues from both monomers. A Zn2+ ion is bound within each active site, coordinated by three conserved histidine residues from both monomers. In addition two conserved amino acids, Asp247 and Asp267, play a role in maintaining integrity of the active site. The substrate is anchored to the enzyme by the interactions of its phospho group and by coordination of the amino and hydroxyl groups by the Zn2+ ion. PdxA is structurally similar to, but limited in sequence similarity with isocitrate dehydrogenase and isopropylmalate dehydrogenase. These structural similarities and the comparison with a NADP-bound isocitrate dehydrogenase suggest that the cofactor binding mode of PdxA is very similar to that of the other two enzymes and that PdxA catalyzes a stepwise oxidative decarboxylation of the substrate HTP.  相似文献   

15.
To describe published experimental data on the functioning of E. coli isocitrate dehydrogenase (IDH), a Rapid Equilibrium Random Bi Ter mechanism involving the formation of two dead-end enzyme complexes is proposed and a kinetic model of enzyme functioning is constructed. The enzyme is shown to be regulated through reversible phosphorylation by IDH kinase/phosphatase; the latter, in its turn, is controlled by IDH substrates and also by a number of central metabolites—pyruvate, 3-phosphoglycerate, and AMP—reflecting the energy demand of the cell. Using the model, it is shown that an increase in the concentration of the above effectors raises the fraction of active IDH and thus enhances the Krebs cycle flux. The ratio between the free and the phosphorylated forms of IDH is more sensitive to AMP, NADP, and isocitrate than to pyruvate and 3-phosphoglycerate. The model also predicts changes in the ratio between phosphorylated and active forms of IDH in the Krebs cycle that occur with a change in the energy and biosynthetic loads on E. coli cells.  相似文献   

16.
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   

17.
Lin AP  McAlister-Henn L 《Biochemistry》2011,50(38):8241-8250
Yeast NAD(+)-specific isocitrate dehydrogenase is an allosterically regulated octameric enzyme composed of four heterodimers of a catalytic IDH2 subunit and a regulatory IDH1 subunit. Despite structural predictions that the enzyme would contain eight isocitrate binding sites, four NAD(+) binding sites, and four AMP binding sites, only half of the sites for each ligand can be measured in binding assays. On the basis of a potential interaction between side chains of Cys-150 residues in IDH2 subunits in each tetramer of the enzyme, ligand binding assays of wild-type (IDH1/IDH2) and IDH1/IDH2(C150S) octameric enzymes were conducted in the presence of dithiothreitol. These assays demonstrated the presence of eight isocitrate and four AMP binding sites for the wild-type enzyme in the presence of dithiothreitol and for the IDH1/IDH2(C150S) enzyme in the absence or presence of this reagent, suggesting that interactions between sulfhydryl side chains of IDH2 Cys-150 residues limit access to these sites. However, only two NAD(+) sites could be measured for either enzyme. A tetrameric form of IDH (an IDH1(G15D)/IDH2 mutant enzyme) demonstrated half-site binding for isocitrate (two sites) in the absence of dithiothreitol and full-site binding (four sites) in the presence of dithiothreitol. Only one NAD(+) site could be measured for the tetramer under both conditions. In the context of the structure of the enzyme, these results suggest that an observed asymmetry between heterotetramers in the holoenzyme contributes to interactions between IDH2 Cys-150 residues and to half-site binding of isocitrate, but that a form of negative cooperativity may limit access to apparently equivalent NAD(+) binding sites.  相似文献   

18.
ExoS and ExoT are bifunctional type III cytotoxins of Pseudomonas aeruginosa that contain an N-terminal RhoGAP domain and a C-terminal ADP-ribosylation domain. Although they share 76% amino acid identity, ExoS and ExoT ADP-ribosylate different substrates. Using protein modeling and site-directed mutagenesis, the regions of ExoS and ExoT that define substrate specificity were determined. Regions B (active site loop), C (ARTT motif) and E (PN loop) on ExoS are necessary and sufficient to recognize ExoS targets, whereas regions B, C and E on ExoT are necessary but not sufficient to recognize ExoT targets, such as the Crk proteins. A specific Crk recognition motif on ExoT was defined as region A (helix alpha1). The electrostatic properties of regions A, B, C and E define the substrate specificity of ExoS and ExoT and these interactions can explain how other bacterial ADP-ribosylating toxins recognize their unique substrates.  相似文献   

19.
Mesecar AD  Koshland DE 《IUBMB life》2000,49(5):457-466
The stereospecificity of the enzyme isocitrate dehydrogenase was examined by steady-state kinetics and x-ray crystallography. The enzyme has the intriguing property that the apoenzyme in the absence of divalent metal showed a selectivity for the inactive l-enantiomer of the substrate isocitrate, whereas the enzyme containing magnesium showed selectivity for the physiologically active d-enantiomer. The hydrogen atom on the C2 carbon that is transferred during the reaction was, in both the d- and l-isocitrate complexes, in an orientation very close to that expected for delivery of a hydride ion to the cosubstrate NADP+. The beta-carboxylate that is eliminated as a CO2 molecule during the reaction occupied the same site on the protein in both the d- and l-isocitrate complexes. In addition, the C3 carbon was in the same protein site in both the d- and l-enantiomers. Only the fourth group, the OH atom, was in a very different position in the apo enzyme and in the metal-containing complexes. A four-location model is necessary to explain the enantiomeric specificity of IDH in contrast to the conventional three-point attachment model. The thermodynamic and kinetic ramifications of this model are explored.  相似文献   

20.
Yeast NAD(+)-specific isocitrate dehydrogenase (IDH) is an allosterically regulated octameric enzyme composed of two types of homologous subunits designated IDH1 and IDH2. Based on sequence comparisons and structural models, both subunits are predicted to have adenine nucleotide binding sites. This was tested by alanine replacement of residues in putative sites in each subunit. Targets included adjacent aspartate/isoleucine residues implicated as important for determining cofactor specificity in related dehydrogenases and a residue in each IDH subunit in a position occupied by histidine in other cofactor binding sites. The primary kinetic effects of D286A/I287A and of H281A replacements in IDH2 were found to be a dramatic reduction in apparent affinity of the holoenzyme for NAD(+) and a concomitant reduction in V(max). Ligand binding assays also showed that the H281A mutant enzyme fails to bind NAD(+) under conditions that are saturating for the wild-type enzyme. In contrast, the primary effect of corresponding D279A/D280A and of R274A replacements in IDH1 is a reduction in holoenzyme binding of AMP, with concomitant alterations in kinetic and isocitrate binding properties normally associated with activation by this allosteric effector. These results suggest that the nucleotide cofactor binding site is primarily contributed by the IDH2 subunit, whereas the homologous nucleotide binding site in IDH1 has evolved for regulatory binding of AMP. These results are consistent with previous studies demonstrating that the catalytic isocitrate binding sites are comprised of residues primarily contributed by IDH2, whereas sites for regulatory binding of isocitrate are contributed by analogous residues of IDH1. In this study, we also demonstrate that a prerequisite for holoenzyme binding of NAD(+) is binding of isocitrate/Mg(2+) at the IDH2 catalytic site. This is comparable to the dependence of AMP binding upon binding of isocitrate at the IDH1 regulatory site.  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