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.
Yeast mitochondrial NAD(+)-specific isocitrate dehydrogenase is an octamer composed of four each of two nonidentical but related subunits designated IDH1 and IDH2. IDH2 was previously shown to contain the catalytic site, whereas IDH1 contributes regulatory properties including cooperativity with respect to isocitrate and allosteric activation by AMP. In this study, interactions between IDH1 and IDH2 were detected using the yeast two-hybrid system, but interactions between identical subunit polypeptides were not detected with this or other methods. A model for heterodimeric interactions between the subunits is therefore proposed for this enzyme. A corollary of this model, based on the three-dimensional structure of the homologous enzyme from Escherichia coli, is that some interactions between subunits occur at isocitrate binding sites. Based on this model, two residues (Lys-183 and Asp-217) in the regulatory IDH1 subunit were predicted to be important in the catalytic site of IDH2. We found that individually replacing these residues with alanine results in mutant enzymes that exhibit a drastic reduction in catalysis both in vitro and in vivo. Also based on this model, the two analogous residues (Lys-189 and Asp-222) of the catalytic IDH2 subunit were predicted to contribute to the regulatory site of IDH1. A K189A substitution in IDH2 was found to produce a decrease in activation of the enzyme by AMP and a loss of cooperativity with respect to isocitrate. A D222A substitution in IDH2 produces similar regulatory defects and a substantial reduction in V(max) in the absence of AMP. Collectively, these results suggest that the basic structural/functional unit of yeast isocitrate dehydrogenase is a heterodimer of IDH1 and IDH2 subunits and that each subunit contributes to the isocitrate binding site of the other.  相似文献   

2.
3.
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   

4.
5.
The expression of the NAD+-specific glutamate dehydrogenase (NAD-GDH) gene of Neurospora crassa is subject to catabolite repression. To identify the minimal sequence necessary for promoter function, the 5'-flanking region of the NAD-GDH gene was screened for potential protein-binding sites. Fragments of DNA, containing sequences upstream from the ATG initiation codon, were employed as probes of Southwestern blots of total cellular protein from cells grown in media promoting repression and induction of NAD-GDH. Two polypeptides interacted differentially with a promoter probe; one was present in greater abundance in repressed cells and a higher relative level of the second was witnessed in induced cells. Electrophoretic mobility shift assays with labeled promoter fragments exhibited preferential interaction with proteins in the induced cultures. The upstream sequence containing the putative protein-binding sites was fused with the coding sequence of the green fluorescent protein (GFP). The resulting plasmid was introduced into the microconidia of an albino mutant of N. crassa by electroporation. Stable integration of the plasmid and_expression of GFP in the hyphae and conidia of the transformants were demonstrated by Southern and Western blot analysis and fluorescence microscopy.  相似文献   

6.
Yeast NAD(+)-specific isocitrate dehydrogenase (IDH) is an octameric enzyme composed of four heterodimers of regulatory IDH1 and catalytic IDH2 subunits. The crystal structure suggested that the interactions between tetramers in the octamer are restricted to defined regions in IDH1 subunits from each tetramer. Using truncation and mutagenesis, we constructed three tetrameric forms of IDH. Truncation of five residues from the amino terminus of IDH1 did not alter the octameric form of the enzyme, but this truncation with an IDH1 G15D or IDH1 D168K residue substitution produced tetrameric enzymes as assessed by sedimentation velocity ultracentrifugation. The IDH1 G15D substitution in the absence of any truncation of IDH1 was subsequently found to be sufficient for production of a tetrameric enzyme. The tetrameric forms of IDH exhibited ~50% reductions in V(max) and in cooperativity with respect to isocitrate relative to those of the wild-type enzyme, but they retained the property of allosteric activation by AMP. The truncated (-5)IDH1/IDH2 and tetrameric enzymes were much more sensitive than the wild-type enzyme to inhibition by the oxidant diamide and concomitant formation of a disulfide bond between IDH2 Cys-150 residues. Binding of ligands reduced the sensitivity of the wild-type enzyme to diamide but had no effect on inhibition of the truncated or tetrameric enzymes. These results suggest that the octameric structure of IDH has in part evolved for regulation of disulfide bond formation and activity by ensuring the proximity of the amino terminus of an IDH1 subunit of one tetramer to the IDH2 Cys-150 residues in the other tetramer.  相似文献   

7.
NAD-dependent Lactobacillus bulgaricus D-Lactate dehydrogenase (D-LDHb) catalyses the reversible conversion of pyruvate into D-lactate. Crystals of D-LDHb complexed with NADH were grown and X-ray data collected to 2.2 A. The structure of D-LDHb was solved by molecular replacement using the dimeric Lactobacillus helveticus D-LDH as a model and was refined to an R-factor of 20.7%. The two subunits of the enzyme display strong asymmetry due to different crystal environments. The opening angles of the two catalytic domains with respect to the core coenzyme binding domains differ by 16 degrees. Subunit A is in an "open" conformation typical for a dehydrogenase apo enzyme and subunit B is "closed". The NADH-binding site in subunit A is only 30% occupied, while in subunit B it is fully occupied and there is a sulphate ion in the substrate-binding pocket. A pyruvate molecule has been modelled in the active site and its orientation is in agreement with existing kinetic and structural data. On domain closure, a cluster of hydrophobic residues packs tightly around the methyl group of the modelled pyruvate molecule. At least three residues from this cluster govern the substrate specificity. Substrate binding itself contributes to the stabilisation of domain closure and activation of the enzyme. In pyruvate reduction, D-LDH can adapt another protonated residue, a lysine residue, to accomplish the role of the acid catalyst His296. Required lowering of the lysine pK(a) value is explained on the basis of the H296K mutant structure.  相似文献   

8.
Erythroascorbic acid (eAsA) is a five-carbon analog of ascorbic acid, and it is synthesized from D-arabinose by D-arabinose dehydrogenase (ARA) and D-arabinono-gamma-lactone oxidase. We found an NAD+-specific ARA activity which is operative under submillimolar level of d-arabinose in the extracts of Saccharomyces cerevisiae. The hypothetical protein encoded by YMR041c showed a significant homology to a l-galactose dehydrogenase which plays in plant ascorbic acid biosynthesis, and we named it as Ara2p. Recombinant Ara2p showed NAD+-specific ARA activity with Km=0.78 mM to d-arabinose, which is 200-fold lower than that for the conventional NADP+-specific ARA, Ara1p. Gene disruptant of ARA2 lost entire NAD+-specific ARA activity and the conspicuous increase in intracellular eAsA by exogenous d-arabinose feeding, while the double knockout mutant of ARA1 and ARA2 still retained measurable amount of eAsA. It demonstrates that Ara2p, not Ara1p, mainly contributes to the production of eAsA from d-arabinose in S. cerevisiae.  相似文献   

9.
Yeast NAD(+)-specific isocitrate dehydrogenase (IDH) is an octamer containing two types of homologous subunits. Ligand-binding analyses were conducted to examine effects of residue changes in putative catalytic and regulatory isocitrate-binding sites respectively contained in IDH2 and IDH1 subunits. Replacement of homologous serine residues in either subunit site, S98A in IDH2 or S92A in IDH1, was found to reduce by half the total number of holoenzyme isocitrate-binding sites, confirming a correlation between detrimental effects on isocitrate binding and respective kinetic defects in catalysis and allosteric activation by AMP. Replacement of both serine residues eliminates isocitrate binding and measurable catalytic activity. The putative isocitrate-binding sites of IDH1 and IDH2 contain five identical and four nonidentical residues. Reciprocal replacement of the four nonidentical residues in either or both subunits (A108R, F136Y, T241D, and N245D in IDH1 and/or R114A, Y142F, D248T, and D252N in IDH2) was found to be permissive for isocitrate binding. This provides further evidence for two types of binding sites in IDH, although the authentic residues have been shown to be necessary for normal kinetic contributions. Finally, the mutant enzymes with residue replacements in the IDH1 site were found to be unable to bind AMP, suggesting that allosteric activation is dependent both upon binding of isocitrate at the IDH1 site and upon the changes in the enzyme normally elicited by this binding.  相似文献   

10.
Based on allosteric regulatory properties, NAD+-specific isocitrate dehydrogenase (IDH) is believed to control flux through the tricarboxylic acid cycle in vivo. To distinguish growth phenotypes associated with regulatory dysfunction of this enzyme in Saccharomyces cerevisiae, we analyzed strains expressing well defined mutant forms of IDH or a non-allosteric bacterial NAD+-specific isocitrate dehydrogenase (IDHa). As previously reported, expression of mutant forms of IDH with severe catalytic defects but intact regulatory properties produced an inability to grow with acetate as the carbon source and a dramatic increase in the frequency of generation of petite colonies, phenotypes also exhibited by a strain (idh1Deltaidh2Delta) lacking IDH. Reduced growth rates on acetate medium were also observed with expression of enzymes with severe regulatory defects or of the bacterial IDHa enzyme, suggesting that allosteric regulation is also important for optimal growth on this carbon source. However, expression of IDHa produced no effect on petite frequency, suggesting that the intermediate petite frequencies observed for strains expressing regulatory mutant forms of IDH are likely to correlate with the slight reductions in catalytic efficiency observed for these enzymes. Finally, rates of increase in oxygen consumption were measured during culture shifts from medium with glucose to medium with ethanol as the carbon source. Strains expressing wild-type or catalytically deficient mutant forms of IDH exhibited rapid respiratory transitions, whereas strains expressing regulatory mutant forms of IDH or the bacterial IDHa enzyme exhibited much slower respiratory transitions. This suggests an important physiological role for allosteric activation of IDH during changes in environmental conditions.  相似文献   

11.
12.
Yeast NAD+-specific isocitrate dehydrogenase (IDH) is an octamer of four IDH1 and four IDH2 subunits, and the basic structural unit of the enzyme is an IDH1/IDH2 heterodimer. To investigate one aspect of the interaction between IDH1 and IDH2, residues in a hydrophobic region at the heterodimer interface (Val-216, Ser-220, and Val-224 in IDH1; Ile-221, Val-225, and Val-229 in IDH2) were replaced by alanine residues in each and in both subunits. Gel filtration and sedimentation velocity analyses demonstrated that the residue substitutions do not disrupt the octameric structure of IDH. However, these substitutions produce novel kinetic properties including, with respect to cofactor, positive allosteric regulation by AMP and cooperativity in the absence of AMP. These allosteric properties are also apparent in NAD+-binding experiments. Despite substantial measurable activity for the mutant enzyme containing residue substitutions in both subunits, expression of this enzyme produces growth phenotypes indicative of IDH dysfunction in vivo.  相似文献   

13.
Recently, we amplified the Lactobacillus bulgaricus NAD(+)-dependent D-lactate dehydrogenase gene by the polymerase chain reaction, cloned and overexpressed it in Escherichia coli (Kochhar, S., Chuard, N., and Hottinger, H. (1992) Biochem. Biophys. Res. Commun. 185, 705-712). Polymerase chain reaction-amplified DNA fragments may contain base changes resulting in mutant gene products. A comparison of specific activities of D-lactate dehydrogenase in the crude extracts of 50 recombinant clones indicated that one of the clones had drastically reduced enzyme activity. Nucleotide sequence analysis of the insert DNA showed an exchange of A to G at position 795 resulting in substitution of Glu264 to Gly in the D-lactate dehydrogenase. The purified mutant D-lactate dehydrogenase showed a shift of 2 units in its optimum pH toward the acidic range. The dependence of kcat/Km on the pH of the mutant enzyme showed that the pKa of the free enzyme was around 4, at least 2 pH units lower than that of the wild-type enzyme. Both the wild-type and the mutant enzyme at their respective optimum pH values showed similar kcat and Km values. The data suggest that the highly conserved Glu264 is not critical for enzyme catalysis, but it must be situated within hydrogen bonding distance to amino acid residue(s) involved in substrate binding as well as in catalysis.  相似文献   

14.
15.
Inosine monophosphate dehydrogenase (IMPDH) plays an important role in the Guanosine monophosphate (GMP) biosynthesis pathway. As hIMPDH-II is involved in CML-Cancer, it is thought to be an active target for leukemic drug design. The importance of conserved water molecules in the salt-bridge-mediated interdomain recognition and loop-flap recognition of hIMPDH has already been indicated in some simulation studies (Bairagya et al., 2009, 2011a, 2011b, 2012; Mishra et al., 2012). In this work, the role of conserved water molecules in the recognition of Inosine monophosphate (IMP) and NAD+ (co-factor) to active site residues of both the isoforms has been investigated by all atoms MD-Simulation studies. During 25-ns dynamics of the solvated hIMPDH-II and I (1B3O and 1JCN PDB structures), the involvement of conserved water molecular triad (W M, W L and W C) in the recognition of active site residues (Asp 274, Asn 303, Arg 322, and Asp 364), IMP and NAD+ has been observed (Figure 1). The H-bonding co-ordination of all three conserved water molecular centers is within 4–7 and their occupation frequency is 1.0. The H-bonding geometry and the electronic consequences of the water molecular interaction at the different residues (and also IMP and NAD+) may put forward some rational clues on antileukemic agent design.  相似文献   

16.
Mitochondrial NAD(H)-specific isocitrate dehydrogenase was purified from Saccharomyces cerevisiae for analyses of subunit structure and expression. Two subunits of the enzyme with different molecular weights (39,000 and 40,000) and slightly different isoelectric points were resolved by denaturing electrophoretic techniques. Sequence analysis of the purified subunits showed that the polypeptides have different amino termini. By using an antiserum to the native enzyme prepared in rabbits, subunit-specific immunoglobulin G fractions were obtained by affinity purification, indicating that the subunits are also immunochemically distinct. The levels of NAD(H)-specific isocitrate dehydrogenase activity and immunoreactivity were found to correlate closely with those of a second tricarboxylic acid cycle enzyme, malate dehydrogenase, in yeast cells grown under a variety of conditions. S. cerevisiae mutants with defects in NAD(H)-specific isocitrate dehydrogenase were identified by screening a collection of yeast mutants with acetate-negative growth phenotypes. Immunochemical assays were used to demonstrate that one mutant strain lacks the 40,000-molecular-weight subunit (IDH1) and that a second strain lacks the 39,000-molecular-weight subunit (IDH2). Mitochondria isolated from the IDH1 and IDH2 mutants exhibited a markedly reduced capacity for utilization of either isocitrate or citrate for respiratory O2 consumption. This confirms an essential role for NAD(H)-specific isocitrate dehydrogenase in oxidative functions in the tricarboxylic acid cycle.  相似文献   

17.
18.
19.
1. The binding parameters for NADH and NAD+ to rabbit-muscle glyceraldehyde-phosphate dehydrogenase (D-glyceraldehyde-3-phosphate:NAD+ oxidoreductase (phosphorylating), EC 1.2.1.12) have been measured by quenching of the flourescence of the protein and the NADH. 2. The fact that the degree of protein fluorescence quenching by bound NAD+ or NADH, excited at 285 nm and measured at 340 nm ('blue' tryptophans), is not linearly related to the saturation functions of these nucleotides, leads to a slight overestimation of the interaction energy and an underestimation of the concentration of sites, if linearity is assumed. 3. This is also the case for NADH, but not for NAD+, when the protein fluorescence is excited at 305 nm and measured at 390 nm ('red' tryptophans). 4. The binding of NAD+ can be described by a model in which the binding of NAD+, via negative interactions within the dimer, induces weaker binding sites, with the result that the microscopic dissociation constant is 0.08 microM at low saturation and 0.18 microM for the holoenzyme. 5. The binding of NADH can be described on the basis of the same model, the dissociation constant at low saturation being 0.5 microM and of the holoenzyme 1.0 microM. 6. The fluorescence of bound NADH is not sensitive to the conformational changes that cause the decrease in affinity of bound NAD+ or NADH. 7. The binding of NAD+ to the 3-phosphoglyceroyl enzyme can be described by a dissociation constant that is at least two orders of magnitude greater than the dissociation constants of the unacylated enzyme. The affinity of NAD+ to this form of the enzyme is in agreement with the Ki calculated from product inhibition by NAD+ of the reductive dephosphorylation of 1,3-diphosphoglycerate.  相似文献   

20.
Neisseria meningitidis was found to contain at least two lactate-oxidizing enzymes. One of these was purified 460-fold from spheroplast membranes and found to be specific primarily for D-lactate, with low-affinity activity for L-lactate. The gene for this enzyme (dld) was cloned, and a dld mutant was constructed by insertional inactivation of the gene. The mutant was unable to grow on D-lactate but retained the ability to grow on L-lactate, providing evidence for a second lactate-oxidizing enzyme with specificity for L-lactate. High-affinity L-lactate-oxidizing activity was detected in intact bacteria of both the dld+ and dld mutant strains. This L-lactate-oxidizing activity was also seen in sonicated bacteria but was reduced substantially on detergent solubilization or on preparation of spheroplast membranes.  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